Você está na página 1de 36

29 Oct 2001 17:38 AR AR151-10.tex AR151-10.

SGM ARv2(2001/05/10) P1: GJC

Annu. Rev. Fluid Mech. 2002. 34:23366


Copyright
c 2002 by Annual Reviews. All rights reserved

DYNAMICAL PHENOMENA IN
LIQUID-CRYSTALLINE MATERIALS
Alejandro D. Rey
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

Department of Chemical Engineering, McGill University, Montreal, Quebec H3A 2A7,


Canada; e-mail: c3co@musica.mcgill.ca
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Morton M. Denn
The Benjamin Levich Institute for Physico-Chemical Hydrodynamics, City College of the
City University of New York, New York, New York 10031; e-mail: denn@ccny.cuny.edu

Key Words liquid crystal, nematic, Leslie-Ericksen theory, liquid-crystalline


polymer, interface
Abstract Recent progress in modeling and simulation of the flow of nematic
liquid crystals is presented. The Leslie-Ericksen (LE) theory has been successful in
elucidating the flow of low molar-mass nematics. The theoretical framework for the
flow of polymeric nematic liquid crystals is still evolving; extensions of the Doi theory
capture qualitative features of the flow of polymeric nematics in simple geometries, but
these theories have not been shown to predict texture development in flow. Mesoscopic
theories for textured materials based on spatial averaging capture only some quali-
tative features of nonrectilinear liquid-crystalline polymer flow. Interfacial effects in
liquid-crystalline systems have begun to receive attention in the context of interfacial
viscoelasticity and the dynamics of dispersed liquid-crystalline polymers in immiscible
blends.

1. INTRODUCTION

The wide range of applications of liquid-crystalline materials has created new ar-
eas of academic and industrial research. Some of the more important applications
of low molar-mass liquid crystals include displays, light valves, and temperature
and pressure sensors, whereas some of the potential uses include chromatography
and smart fluids for brakes and clutches. One of the most important new develop-
ments in display technology is the emergence of polymer-dispersed liquid crystals
for applications in flat-panel television technology and switchable windows. The
synthesis of polymer liquid crystals has enlarged the range of applications of these
materials to areas where unique properties are important, including high-strength
fibers and injection-molded parts for electronic interconnects and extreme chem-
ical and thermal environments. Some naturally existing materials, like coal and
petroleum pitches, are spun from the liquid crystalline state into high-strength
0066-4189/02/0115-0233$14.00 233
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

234 REY DENN

carbon fibers. A large number of reviews, textbooks, and monographs on the theory,
applications, and rheology of liquid-crystalline materials are available (de Gennes
& Prost 1993, Kleman 1983, Ciferri 1991, Noel & Navard 1991, Beris & Edwards
1994, Srinivasarao 1995, Marrucci & Greco 1993, Burghardt 1998, Larson 1999,
Demus et al. 1999, Sonin 1995, Rey 1993a, Tsuji & Rey 1998, Sawyer et al. 1998).
The modeling of these structured materials must take the partial positional and
orientational order into account, which necessitates adding new balance equations
to those that govern structureless fluids. For a uniaxial nematic liquid crystal,
for example, an internal momentum balance equation is required to describe the
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

average macroscopic orientation of the liquid. As is the case in other continua,


Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

constitutive equations reflecting the symmetry properties of the phases are required
to specify relations between forces and fluxes. These requirements give rise to a
variety of complex macroscopic theories applicable to different phases.
The nematic phase, which is characterized by orientational order but positional
disorder, is perhaps the most-studied liquid-crystalline phase and is the focus of
this chapter. Nematics are usually anisotropic molecules with rigid segments in
the structure, often because of the presence of parasubstituted phenyl rings. The
spatial and temporal distribution of the average orientation in liquids made up of
rod-like or disk-like molecules can often be described by a unit vector field n, the
director. The presence of order admits the possibility of defects, from which the
name nematic is derived (Kleman 1983). The spatial distribution of defects defines
the texture of a liquid crystal; texture is a fingerprint of a liquid-crystalline phase
and plays a significant role in the rheology. The imperfect molecular alignment
along the average orientation is described by one or more scalar order parame-
ters that, in polymeric nematic liquid crystals, are likely to be affected by strong
fields. The Leslie-Ericksen (LE) vector theory, which is applicable to rigid-rod and
discotic nematics in the absence of spatio-temporal variations of the scalar-order
parameter, is successful in describing the flow of low molar-mass nematics (de
Gennes & Prost 1993). The theory may also apply to some slow flows of nematic
polymers. Sufficiently strong polymer flows affect the order parameters, and mod-
els based on the dynamics of the nonequilibrium orientation distribution function,
such as the Hess theory (Hess 1975) and Doi theory (Larson 1999) as well as
its extensions (Feng et al. 2000, Kupferman et al. 2000, Lhuillier 2000), are nec-
essary. The mesoscopic Landaude Gennes tensor model, which is based on the
second moment of the orientation distribution function, generalizes the LE theory
to variable order parameters and is successful in describing flow-induced textural
transformations (Tsuji & Rey 1998) up to the defect-core length scale (Schopohl
& Sluckin 1987, 1998; Hudson & Larson 1993).

2. CLASSIFICATION OF LIQUID-CRYSTALLINE PHASES


2.1. Low Molar-Mass and Polymeric Liquid Crystals
Phase transitions to the ordered fluid state can be effected through changes in
temperature (thermotropic liquid crystals) or concentration (lyotropic liquid
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 235

crystals). The molecular orientational order can best be achieved with molecu-
lar shapes that are disc-, lath-, or rod-like. Typical thermotropic low molar-mass
liquid crystals are N-(p-methoxybenzylidene)-p0 -n-butylaniline (MBBA), 4,40 -
Di-methoxyazoxybenzene (p-azoxyanisole), 4-pentyl-40 -cyano-biphenyl (5CB),
and 4-octyl-40 -cyano-biphenyl (8CB) (de Gennes & Prost 1993). Macromole-
cules can also form similar liquid crystals (Donald & Windle 1992); a typi-
cal example is the random copolymer of 73% p-hydroxybenzoic acid and 27%
p-hydroxynaphthoic acid, sold commercially as Vectra A. The mesogenic groups
can be attached to the macromolecules as side chains or part of the main chain. So-
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

lutions of anisotropic molecules in an isotropic solvent form liquid-crystal phases


Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

for sufficiently high solute concentration; a typical example of a lyotropic liquid-


crystalline polymer is poly (1,4-phenyleneterephthalamide), sold commercially as
Kevlar. The most thoroughly characterized lyotropics are liquid-crystalline poly-
mers formed by concentrated solutions of synthetic polypeptides. Polybenzyl-
L-glutamate (PBLG) assumes a rod-like alpha-helical conformation in many
solvents. Other naturally occurring phases are obtained from certain linear viruses,
such as tobacco mosaic virus (TMV) (Lee & Meyer 1991). The natural parameter
that effects the transition is the concentration because the principal interaction
producing long-range order is the solute-solvent interaction (Donald & Windle
1992).
Symmetry considerations led Friedel (de Gennes & Prost 1993) to distin-
guish the three main classes of liquid crystals: nematic, cholesteric, and smectic.
Schematics are shown in Figure 1. (a) Uniaxial nematic order: The two main fea-
tures of the nematic phase are the long-range orientational order and the fluidity.
This phase has cylindrical symmetry and is therefore uniaxial. The direction of the
axis of cylindrical symmetry is arbitrary in space and is described by the director n.
Equilibrium biaxial nematic phases are discussed in de Gennes & Prost (1993). In
this review, we abbreviate liquid crystals by LCs, nematic liquid crystals by NLCs,
low molar-mass nematic liquid crystals by LMMNLCs, and liquid-crystalline ne-
matic polymers by LCNPs. (b) Cholesteric order: Chiral nematic molecules give
rise to the cholesteric phase. The cholesteric phase lacks positional order, as does
the nematic phase, but the director follows a helical path. The strong modulation
of the refractive index due to the helical deformation causes Bragg scattering of

Figure 1 (A) The uniaxial rod-like nematic phase, (B) the helical structure of the
cholesteric phase, and (C) the smectic A phase.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

236 REY DENN

various colors of light, a property that is exploited in the use of liquid crystals as
temperature sensors. Polymers possessing sufficient rigidity and chiral centers on
the chain backbone can exhibit cholesteric behavior. The sense of the twist is sol-
vent dependent, and the nematic phase may be formed from the cholesteric by the
compensatory effect of a binary mixture; an example is poly (benzyl-glutamate)
(PBG), which admits a nematic phase from the racemic mixture of cholesteric D
and L compounds. (c) Smectic order: The remaining liquid-crystalline phases are
all smectics, and as many as 11 are known to exist. The smectics are distinguished
by having an intermediate degree of positional order in addition to orientational
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

order. The best-known smectic phases are the A and C, which have one degree of
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

translational ordering and hence a layered structure. A schematic of the smectic A


is shown in Figure 1C.

2.2. Micellar Liquid Crystals


Lyotropic liquid-crystalline phases can be formed by amphiphilic molecules in
solution. Amphiphilic molecules consist of a hydrophilic (polar) head that attracts
water and a lipophilic tail that avoids water. The lipophilic tail is usually a long
hydrocarbon chain. Potassium stearate, phospholipids, and sodium dodecyl sulfate
are examples of amphiphilic molecules. Amphiphilic molecules can be cationic,
ionic, and nonionic surfactants, and some have double lipophilic tails. At a water-
hydrocarbon interface, the amphiphilic molecules form a smectic layer, whereas
in the two bulk phases, if the concentration is above the critical micellar concentra-
tion, microdomains known as micelles form. In the water phase, normal micelles
with the head on the outside form, whereas in the hydrocarbon phase, inverse
micelles with the head at the inner side form. The micelles may be spherical,
rod like, or disk like. If the surfactant concentration is further increased, ordered
mesophases such as nematic, smectic, hexagonal, or cubic can arise depending
on the particular surfactant(s)-solvent(s) system (Boden 1994, Larson 1999). For
in-depth discussion of the conditions for formation of the various micellar LC
phases, see Boden (1994) and Larson (1999).

3. DEFECTS AND TEXTURES IN NEMATIC


LIQUID CRYSTALS
A nematic phase exhibits broken symmetry when compared to the isotropic phase
because it is defined by the average molecular orientation, n. Because any orien-
tation of n is permissible in principle, the degeneracy leads to the possibility of
defects, or spatial discontinuities of n. Defects are classified in terms of strength
(S) and dimensionality (D). The strength captures the degree of rotational discon-
tinuity when encircling the defect, whereas the dimension refers to points, lines,
and walls (Kleman 1983, 1989; Bouligand 1998, Noel 1998). The spatial arrange-
ment of defects is called texture, and each class of LC displays a distinguishing
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 237

number of textures; textures in many liquid-crystalline materials have been


thoroughly characterized (Slaney et al. 1998, Demus & Richter 1978, Bouligand
1998, Kleman 1991, Zimmer & White 1982). Defects play a determining role in
the suitability of a liquid crystal as a material of choice for product manufacturing,
and the defect texture is undoubtedly a factor in difficulties encountered in control-
ling spatial inhomogeneity in the industrial production of large molded parts from
liquid-crystalline polymers (Sawyer & Jaffe 1986). Moreover, measured phys-
ical properties are intrinsic material properties only when the material is a liquid
monocrystal, and the physical properties of textured liquid crystals are functions of
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

textural parameters, such as defect type and density. A large body of research has
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

been directed to understanding the physics of defect nucleation and coalescence for
low molar-mass liquid crystals (de Gennes & Prost 1993, Bouligand 1998, Kleman
1983) and for liquid crystalline polymers (Noel 1998, Kleman 1991, ORourke &
Thomas 1995, Larson 1999).
The nematic liquid monocrystal (monodomain) state appears only with com-
patible surface effects or in the presence of highly orienting external fields, such
as strong electromagnetic fields or extensional flow. Nematic textures are ubiqui-
tous and are generated, in the absence of external fields, by (a) the nematic-phase
formation process, as in the thermally induced phase transition of the isotropic
phase into the nematic phase (Bouligand 1998) and (b) by incompatible orient-
ing surface couples, arising from untreated surfaces, geometric singularities, or
suspended second phases (i.e., filled nematics, emulsions, etc.). In the absence
of external flows, texture coarsening is the result of defect-defect and defect-
surface interactions and defect reactions; the reactions are governed by the con-
servation of topological charge, such as one S = +1 disclination line decaying
into two S = +1/2 lines, or a disclination loop decaying into a point defect (Rey
1990a). Imposed external flows also create (de Gennes & Prost 1993; Rey 1993b;
Mather et al. 1996, 1997; Alderman & Mackley 1985, Rey & Tsuji 1998), trans-
form, and coarsen textures (Larson 1999, Kleman 1991, Noel 1998, Rey & Tsuji
1998).

3.1. Classification and Stability of Defects


The strength S of a defect is equal to the number of rotations the director experiences
on a path encircling the defect. Points are D = 0 defects, disclinations are D = 1
line defects, and walls are D = 2 defects. Walls and disclination ends cannot be
found in the bulk of the sample. Wedge and twist disclinations are possible ac-
cording to the rotation axis: Wedge (twist) lines are parallel (perpendicular) to the
rotation axis (Noel 1998). The cores of disclinations can be singular or nonsin-
gular; S = 1/2 disclinations have singular cores, whereas S = 1 disclinations
can have either. Figure 2 shows schematics of some representative disclinations.
In singular cores, the orientational order decreases, whereas in nonsingular cores,
the director escapes into the third dimension. Because defects have associated
core and distortion energies, short- and long-range energy calculations are used to
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

238 REY DENN


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

Figure 2 Schematics of director distributions ( full lines) around some wedge dis-
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

clinations.

establish their stability (Kleman 1991, Noel 1998); these calculations involve the
Oseen-Frank long-range elastic energy density fg:

2fg = K11 (.n)2 + K22 (n. n)2 + K33 |n n|2 , (1)

where K11, K22, and K33 are the splay, twist, and bend Frank elastic constants,
respectively, whose corresponding distortions are shown in Figure 3. Defect energy
is proportional to S2 and to the appropriate {Kii}. Anisimov & Dzyaloshinskii
(1972) showed that elastic anisotropy controls the stability of disclinations and the
relative abundance of certain types of defects.

3.2. Texture Creation During the


Isotropic Nematic Phase Transition
During the isotropic-nematic mesophase phase tranformation, when nematic sphe-
rules grow in an isotropic matrix and coalesce into larger mesophase regions, a large
number of disclinations nucleate (Bouligand 1998, Noel 1998) because the lack of
orientation registry between the uncoalesced mesophase regions is resolved by the
nucleation of disclinations. Defect nucleation during phase change is a universal
feature of phase transitions with broken symmetry; Trebin (1998) has shown the
correspondence between defects in nematics and defects in cosmology, where
the defect nucleation process is known as the Kibble mechanism (Kibble 1976).
The defect density at the end of the phase-transition stage was calculated by Kibble

Figure 3 Schematic of the three elastic modes: (A) splay (K11), (B) twist (K22), (C) bend
(K33). The segments represent the director n, and the arrows the rotation of n.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 239

within the cosmological model and by Rey & Tsuji (1998) using the Landaude
Gennes model for LCNPs. In the latter, the single parameter that controls defect
density at the phase transition is the ratio R of short-range (homogeneous) to long-
range (Frank) nematic elasticity. For relatively large R, long-range elasticity is
unable to transmit orientation information effectively, and thus, lack of orientation
registry between coalescing mesophase spherules is more prevalent than when R
is relatively smaller.
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

3.3. Texture Coarsening


Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Unstable textures undergo coarsening processes driven by, among several possi-
bilities, defect-defect reactions and annihilations and defect recombinations that
emit loops that shrink (Trebin 1998). Thus, texture coarsening is driven by a de-
fect density reduction. Texture coarsening is a self-similar process whose scaling
law gives the mean defect separation distance d in a space of dimension d as a
power law in time: d (t)1/d tn , where is the defect density. For nematic
polymers, Rojstaczer et al. (1988) and Shiwaku et al. (1990) showed experimen-
tally that for d = 2, n 0.350.37. Chuang et al. (1991) performed studies of the
isotropic rod-like nematic transition of 5CB by pressure quenches. Texture coars-
ening followed the self-similar process, with the power-law exponent n = 0.5, as
expected in dispersive systems (Trebin 1998). String recombination and decay of a
texture into monopole-antimonopole pairs was also observed. Texture coarsening
was simulated using the Landaude Gennes equations by Rey & Tsuji (1998), who
found that the coarsening-law exponent is a function of the ratio R of short-range
to long-range elasticity, such that the coarsening rate increases with increasing
R. As R , the results of the LE theory for a dispersive system are recovered.
Texture coarsening has been simulated at the vector level by Rieger (1990), using a
reaction-diffusion model, and by Greco (1989), using a phenomenological dipole
model; initial conditions for the number, type, and strength of the defects have
to be arbitrary for these models, which cannot capture texture formation. Energy
minimizations of long-range elasticity for nematic textures have been performed
by Bedford et al. (1991), Bedford & Windle (1993), and Kimura & Gray (1993);
these static simulations are unable to capture the self-similarity of the coarsening
process.
Abundant experimental data on textural transformation and texture coarsening
under flow clearly indicate that defects may be nucleated under flow and that
defect-defect annihilation processes are modified by the presence of flow (Noel
1998). Defect nucleation under flow has been characterized for rod-like nematics
(Graziano & Mackley 1984; Alderman & Mackley 1985; Mather et al. 1996, 1997;
Noel 1998; Srinivasarao 1995; Burghardt 1998; and references therein) and is also
related to the texture refinements reported for liquid-crystalline polymers (Larson
& Mead 1993, Larson 1999, Kleman 1991, Noel 1998, Srinivasarao 1995, Vermant
et al. 1994, Walker & Wagner 1994, Ugaz et al. 1997, Hongladarom & Burghardt
1998, Burghardt 1998). Simulations of texture coarsening under flow are discussed
subsequently.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

240 REY DENN

4. THEORY AND SIMULATIONS BASED


ON LESLIE-ERICKSEN THEORY
4.1. Leslie-Ericksen Equations
In the Leslie-Ericksen (LE) theory for NLCs (Leslie 1983, de Gennes & Prost 1993,
Larson 1999), the microstructure of the material is explicitly taken into account
through a director of unit magnitude. Assuming that the fluid is incompressible,
the mass, linear momentum, director, and internal energy balance equations of
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

microcontinuum mechanics are as follows:


Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

.v = 0; v = f + .T; 1 n = G + g + . ; U = T : A + : M g.N,
(2)
where f is the body force per unit volume, T the total stress tensor, v the linear
velocity, 1 the moment of inertia per unit volume, G the external director body
force (torque per unit volume), the director stress tensor, U the internal energy
per unit volume, N the angular velocity of the director relative to that of the fluid, M
the gradient of N, and g the intrinsic director body force. The kinematic measures
are
N = n + W.n; M = n + W.n; 2A = v + vT ; 2W = v vT ,
(3)
where A and W are the rate of deformation tensor and the spin tensor, respectively.
The LE constitutive equations for the stress tensor and the director body force are
given in terms of objective linear functions of N and A. Expanding in these vari-
ables and using transversely isotropic tensor coefficients that reflect the material
symmetry, the following equations are found:
fg
T = pI nT + 1 A : nnnn + 2 nN
n
+ 3 Nn + 4 A + 5 nn.A + 6 A.nn, (4)

fg fg
g = an b.n 1 N 2 n.A; = bn . (5)
n n
p is the pressure; I the unit tensor; {i }, i = 1 . . . 6, the six Leslie viscosity co-
efficients; a the director tension; b a Lagrange multiplier vector; 1 = 3 2 the
rotational viscosity; and 2 = 6 5 = 2 + 3 the irrotational torque coefficient,
where the last equality follows from the Onsager reciprocal relations.
The boundary conditions for the velocity are usually no slip at the bounding
surfaces. Boundary conditions for the director are given (Rey 2000) by the balance
of the surface elastic and the surface viscous torque:
n hse n hsv = 0. (6)
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 241

The surface elastic molecular field hes is given by


fs fg
hse = k. , (7)
n n
where k is the unit normal, and the surface free energy density fs is given by

fs (n.k, T) = o (T) + an (n.k, T); an (n.k, T) = 2 (T)[n.k]2 + 4 (T)[n.k]4 ,


(8)
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

where T is the temperature, 0 is the isotropic interfacial tension, and an is


Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

the anisotropic contribution to the surface free energy, known as the anchor-
ing energy. For LMMNLCs, the isotropic surface interfacial tension o is of the
order of 10 ergs/cm2, whereas an varies from 1041 erg/cm2 (Sonin 1995).
The director orientation that absolutely minimizes the surface free energy den-
sity is known as the easy axis of the interface. For nondeforming interfaces, the
viscous surface molecular field takes dissipation due to orientational slip into
account:
dn dn
hsv = 1/
s
/ Is . + 1
s
kk. , (9)
dt dt
where Is is the surface unit tensor and 1/
s
/ and 1 are surface viscosities associated
s

with tangential and normal rotations of the director. According to the general
boundary condition (Equation 6), two particular static-director surface conditions
are possible: (a) no-torque condition, k.fg /n = 0, corresponding to the case of
insignificant surface anchoring energy, and (b) strong-director surface anchoring,
n = nfix, corresponding to the case in which bulk gradient elasticity is insignificant
with respect to surface anchoring energy. All the terms in Equation 6 must be
retained when both anchoring and gradient energies are equally significant and
orientational slip occurs.

4.2. Scaling and Dimensionless Numbers


The characteristic static and dynamic behavior of NLCs is found by using the time
scales, length scales, and dimensionless numbers of the LE equations. The set of
material parameters in the generalized LE theory are {i , i = 1, . . . , 6; Kii , ii =
11, 22, 33; 1/
s
/ ; 1 ; 0 ; 2 ; 4 }. If the characteristic geometric length scale is L, the
s

extrinsic timescale of the LE theory is the orientation diffusion time or = 1 L 2 /K,


where K is a characteristic Frank constant. In the generalized LE theory, there
is an intrinsic length scale, given by the ratio of gradient to anchoring energy:
`ga = K/2 . In the absence of flow, orientation gradients can exist only if L > `ga .
In the presence of a flow whose time constant is flow , flow-induced orientation is
possible only if flow < or . This last condition is expressed in terms of the Erick-
sen number, which for a shear flow of rate is E = 1 L2 /K. Thus it is found
that, if E 1, flow-induced orientation will be homogeneous only if L < `ga .
Comparing the transient apparent shear viscosity of a NLC in simple shear
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

242 REY DENN

between two plates with separation L, L , shear rates and , and shear
strains and , it follows that ( L2 , ) = ( L2 , ) if L2 = L2
and = .

4.3. Flow-Induced Orientation


According to the LE theory, in the absence of elastic effects due to director gra-
dients, a flow orients the director such that the total viscous torque density v
vanishes:
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

v = 1 n [n + (A W).n] = 0,
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

(10)

where the first term represents the effect of transient rotation and the second
the effect of the deformation. The temperature-dependent reactive parameter
establishes the relative magnitude of the rotational (W) to the aligning (A) flow
effects.
Both A and W are nonzero for simple shear. For rod-like NLCs, > 0 and
two modes are possible: (a) 0 < < 1, called the tumbling or nonaligning mode,
and (b) > 1, called the aligning mode. Consider a simple-shear flow along the z
direction, with v = (0, 0, x). We assume the director orients in the shear plane, so
we can write n = (sin , 0, cos ), in which case Axz = /2, Nx = sin /2,
Nz = cos /2. The viscous torque along the y axis computed from Equa-
tion 10 gives the following director angle equation:

/t + 1 [1 cos 2 ]/2 = 0, (11)

with solutions
(r p )
1+ t
(a) nonaligning mode: 0 < < 1, = arctan tan 1 2
1 2

(12)

1 1
(b) aligning mode: > 1, L = arc cos ; (13)
2
L is known as the Leslie angle. The possible stable alignment angles are restricted
to 0 < < /4. For nonaligning materials the shear flow attempts to rotate the
director with a period equal to [2 (1 2 )1/2 / ].
The orienting behavior of uniaxial nematics in shear-free flows (Rey & Denn
1988) is simpler because N = 0, and A has a simple diagonal form at steady state:

e1 0 0

A = 0 e2 0 ; e1 + e2 + e3 = 0. (14)
0 0 e3
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 243

According to Equation 10, the steady-state orientation occurs when n is colin-


ear with the eigenvectors of A, and the orientation is stable when n is normal
(parallel) to the compression (extension) axis or plane. For uniaxial extensional
flows, e1 = e, e2 = e3 = e/2, rod-like nematics align with the director along
the stretching direction (1-axis). For equibiaxial stretching flows, e1 = 2e,
e2 = e3 = e, and n aligns normal to the compression direction (1-axis).
The generalization of the expression for viscous torques in a binary miscible
NLC mixture was derived by Rey (1996). The concentration-dependent reactive
parameter of the mixture at low deformation rate mix 0 is given by
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

8C 18
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

mix = n + m , (15)
0
8(C 1) + 1 8(C 1) + 1
where 8 is the relative volume fraction, C(8) is the concentration-dependent
ratio of the rotational viscosities of the pure components, and {i ; i = n, m} are
the reactive parameters of the pure components. If C is independent of 8, then
Equation 15 is in agreement with low deformation-rate experimental results (Liu
& Jamieson 2000, Ternet et al. 1999).

4.4. Shear Flows of Nonaligning LE Fluids


The presence of long-range elasticity and possible strong anchoring of the director
at the bounding surfaces modifies the planar (two-dimensional orientation) director
dynamics of nonaligning NLCs (0 < < 1) under simple-shear flow. The resulting
orientation must satisfy the torque balance equation

fg fg
+ = 0;
e v
= n
e
. , (16)
n n
where e is the elastic torque and v is the viscous torque defined in Equation 10.
The elastic torques arising from in-plane orientation gradients stabilize the flow
and result in steady-state planar solutions, with a series of limit-point bifurcations
known as tumbling instabilities, but at high shear rates, stationary solutions exist
only if the director escapes from the shear plane. Studies using the full nonlinear
LE equations with fixed orientation at the walls (n = nw = constant) (Luskin &
Pan 1992, Zuniga & Leslie 1989, Han & Rey 1993) found that out-of-plane (OP)
solutions bifurcate supercritically from in-plane (IP) solutions at a critical Ericksen
number; the denotes the fact that the director can escape the shear plane in
two equally possible directions. This out-of-plane transition generally precedes
the two-dimensional tumbling instability. Both OP and partially IP modes of
sheared nonaligning NLCs have been experimentally observed in 8CB. Pieranski
& Guyon (1974) reported that OP orientation exists throughout the sample thick-
ness, whereas Cladis & Torza (1975) detected OP orientation at the interface
between the center and wall regions, with IP orientation at the centerline region.
These two sets of experimental observations were predicted by Han & Rey (1994),
using 8CB material data at 35 C and homeotropic director anchoring, as shown
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

244 REY DENN


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 4 Computed visualizations of director orientation of sheared 8CB with homeo-


tropic anchoring. (a) E = 133, (b) E = 130.

in Figure 4. The results of Cladis & Torza (1975) correspond to a chiral OP


solution branch that displays IP orientation at the center gap region but has ori-
entation along the vorticity close to the two bounding surfaces, when E = 133,
as shown in Figure 4a. Pieranski & Guyons (1974) results were replicated when
E = 130, when the stable solution branch is an achiral OP mode, as shown in
Figure 4b.
The thermal dependence of the transient and steady rheological material func-
tions of 8CBP has been studied experimentally by Gu & Jamieson (1994) and
simulated by Han & Rey (1995a) using the full nonlinear LE equations. This NLC
compound becomes isotropic at T = TNI = 40.8 C, exhibits a lower temperature
smectic-A phase at T = TNSm = 33.2 C, and undergoes the aligning nonaligning
rheological transition at T = Tan = 38.45 C. Differences between the experimental
cone-and-plate geometry and the parallel-plate geometry of the model lie in the Er-
icksen number, which is position dependent in the former and constant in the latter.
For T > Tan (T < Tan) the behavior is aligning (nonaligning). For shear steps, the
simulations reproduce the emergence and oscillatory dissipation of characteristic
double peaks in the transient shear viscosity, driven by the IP OP transition.
The simulations also reproduce smectic-A pretransitional phenomena. Quantita-
tive agreement between simulation and experiments was observed for the highly
nonlinear temperature dependence of the apparent steady-shear viscosity , in-
cluding the viscosity jump at Tan and the cross-over = 4 /2 at T = 36.2 C, as
shown in Figure 5a. The computed steady-state nematodynamic phase diagram,
in terms of the Ericksen number and temperature, and characteristic orientation
structure visualizations for sheared 8CB with homeotropic director anchoring, are
shown in Figure 5b. Close to the isotropic phase, the NLC displays the aligned
boundary-layer IP mode for all E, whereas at T < Tan, the splay-bend distorted,
IP mode exists only if E < 100. At higher E, only the twisted OP mode with few
splay or bend distortions is stable.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 245


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 5 (a) Temperature dependence of the apparent steady-shear viscosity (large


dots) as a function of 1T = TNI-T. At 1Tan = TNI-Tan, the viscosity exhibits a jump, and
close to 1TMV = TNI-TMV, = a = b. (b) Nematodynamics phase diagram of sheared
8CB with homeotropic anchoring. The phase fields correspond to the basin of attractions
for initial conditions ny = 1. The vertical line at T = 38.5 C corresponds to the aligning-
nonaligning transition (a first-order transition), whereas the slanted dashed line gives the
E (T) at which the IP-OP transition occurs (a second-order transition). From Han & Rey
(1995a), with permission of the Journal of Rheology.

4.5. Nonviscometric Flows


Simulations of microstructure evolution in nonviscometric flows of aligning ne-
matics using the LE equations have been performed for Jeffrey-Hamel (Rey &
Denn 1988) and radial outflow between disks (Rey 1990b). Viscous torques in-
clude shear and elongation, and the stable orientation is achieved by balancing
the three torques. The Jeffrey-Hamel flow geometry is best described in (r, , z)
cylindrical coordinates. The shear plane is spanned by (r, ), and z is along the
vorticity axis. The IP director angle is , and the director and velocity fields are
n() = (cos , sin , 0); v() = (u, 0, 0). For converging (diverging) flow u < 0
(u > 0). The stable centerline director is along the extension direction, which for
converging (diverging) flow is along radial (azimuthal) direction. The viscous
torque balance yields, at steady state, the following director equation:
u0
[ cos 2 1] + [ sin 2 ]u = 0. (17)
2
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

246 REY DENN

The stable solutions, consistent with the centerline orientations for diverging and
converging flows are, respectively,
" p #
+ 1 2u 42 u2 + (2 1)u02
9 > 0, a = tan ; 9 < 0, a = a+
u 0 (1 + )

(18)

" p #
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

2u + 42 u2 + (2 1)u02
9> 0, a+ 1
= tan ; 9 < 0, a = a+ .
u 0 (1 + )
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

(19)
In the centerline (wall) region, elongation (shear) dominates. Both shear and ex-
tension promote aligning close to the radial direction for converging flow, whereas
for diverging flow, shear promotes radial alignment and elongation azimuthal align-
ment. Thus, at high E, converging (diverging) flows have weak (strong) orienta-
tion gradients. These results are a consequence of the fact that extension and
compression are co-planar with shear, and their magnitudes have the same radial
dependence.
Two-dimensional converging flow of an aligning LE fluid with rectilinear up-
stream and downstream regions has been simulated by Chono et al. (1994). Director
alignment in the converging region is essentially along the streamlines, as expected.
The recirculating region near the corner is larger than that in a Newtonian fluid
because of the entropic elasticity terms in the constitutive equation.
Nonviscometric radial outflows of aligning NLCs between two parallel disks
have been characterized experimentally by Hiltrop & Fisher (1976) and simulated
with the LE equations by Rey (1990), assuming homeotropic (orthogonal) director
anchoring. For flow-aligning NLCs, the net effect of shear is to try to keep the
director aligned close to the flow direction in the shear plane, whereas the effect of
elongation is to align the director along the extension direction. Inhomogeneous
shear-elongational flows will exhibit transitions from IP modes to OP modes as a
result of these competing orienting effects. The increase in cross-sectional area in
the pressure-driven radial outflow between parallel disks introduces a stretching
deformation normal to the shear plane and drives the emergence of three possible
orientation modes, shown as side-view schematics in Figure 6. For homeotropic
director anchoring there is a critical pressure drop for each radial distance at
which the original IP bow mode becomes unstable and any of two OP screw
modes appear within a cylinder centered at the axis of the disks. The radius of the
cylinder containing the OP modes is pressure dependent; increasing the pressure
drop increases the radius. Further increase in pressure drop above another critical
value produces the emergence of a radially aligned IP peak mode.
There have been few simulations of nonviscometric flows of tumbling nematics
using the LE equations. Rey & Denn (1989) studied planar converging flow of a
tumbling nematic and found a cascade of transitions. Burghardt & Fuller (1990)
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 247


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 6 Side view of the sequence of orientational modes of a flow-aligning NLC


in radial outflow, with homeotropic anchoring, as the pressure drop increases: (A) in-
plane bow mode, (B) left-screw mode, (C) right-screw mode, (D) in-plane peak mode.
The arrows represent changes due to increasing pressure drop.

studied transients in plane shear. Chono et al. (1998) computed velocity and texture
development in pressure-driven planar channel flow. A typical result is shown in
Figure 7 for parameters characteristic of 8CB at 34 C and for Ericksen numbers
ranging from 10 to 70. The velocity profile is essentially independent of the director
orientation. The number of tumbling transitions in the fully developed region
increases with Ericksen number. Regions of different orientation are separated by
splay-bend walls; singularities are not possible at steady state in a LE fluid because
of the diffusive nature of the director equations. It is likely that some of these states
are unstable to out-of-plane perturbations.
Figure 8 shows the coating thickness e deposited on a fiber of radius r dragged
through a bath of a LMMNLC mixture (E7) at 25 C. The coating flow consists of
mixed extension and shear. The thickness follows a scaling e/r Ca 0.94 , where Ca
is the capillary number, rather than the classical Landau-Levich-Derjaguin scaling
e/r Ca 2/3 for Newtonian fluids. The 2/3 power scaling is observed for coating
above the nematic-isotropic transition temperature. This scaling is predicted by
the LE equations in the absence of elasticity (Park et al. 2001) and originates in
the coupling between orientation and stress.

4.6. Banded Textures


Light transmission patterns of sheared LCPs under crossed polars show a banded
texture that arises during and/or after cessation of flow (Elliott & Ambrose 1950,
Toth & Tobolsky 1970, Kiss & Porter 1980, Donald et al. 1983, Navard 1986,
Srinivasarao & Berry 1991, Larson & Mead 1992, Vermant et al. 1994, Larson
1999, Harrison & Navard 1999). The formation of banded textures under a mag-
netic field has been simulated with the full LE theory under fixed anchoring con-
ditions (Rey & Denn 1990) and under orientational slip (Rey 1991), with good
agreement with experiments. Simulations of banded textures after cessation of
flow using extensions of the LE and Landaude Gennes equations (Rey & Tsuji
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

248 REY DENN


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 7 Developing flow between parallel plates for a LE fluid with parameters char-
acteristic of 8CB at 34 C. The wall is at y = 0 and the midplane at y = 0.5. E = 10 (top),
50 (middle), and 70 (bottom). From Chono et al. (1998), with permission from Elsevier
Science.

1998) indicate that molecular and defect elastic storage drive the pattern-formation
process.
Yan & Labes (1992, 1994) observed a clear banded-texture image during weak-
shear start-up flow of PBG, using strong homeotropic director anchoring and a
monodomain initial condition; the relations between the banded-texture charac-
teristics, the molecular weight, and the shear rates were experimentally determined.
These studies have established that the banded texture is a result of a periodic spa-
tial modulation of the orientation, leading to a spatial variation of the effective
birefringence. Han & Rey (1995b) simulated the experiments of Yan & Labes
(1994) by computing the transient nonplanar orientation pattern formation and
corresponding light transmission during shear flow of PBG, using the full LE
equations and experimentally measured viscoelastic material constants. Figure 9a
shows a visualization of the nz director component, where x is horizontal (velocity
direction), y vertical (gradient direction), and z (vorticity direction) into the page.
The figure shows the presence of an array of tubular orientation inversion walls
immersed in a matrix of planar (nz 0) orientation, whose axes are orthogonal to
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 249


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 8 Reduced-film thickness for wetting of a polymer fiber by a nematic liquid-


crystal mixture (E7). The data follow a power law of Ca0.94, whereas Newtonian fluids
follow the Landau-Levich-Derjaguin 2/3 scaling. From Park et al. (2001).

the x direction. Figure 9b shows the corresponding light transmission intensity as


a function of dimensionless distance along the x axis under crossed polars, clearly
showing periodic behavior along the flow direction. Yan & Labes (1994) found that
as the shear rate increases the wavelength decreases from infinity to a saturation
value equal to the half-gap thickness. The infinite wavelength occurs at a critical
shear rate, or equivalently at a critical Ericksen number, corresponding to the IP-OP
orientation transition. The transition for periodic pattern formation is found to be
very close to the critical Ericksen number for the nonperiodic planar nonplanar
orientation transition (Han & Rey 1995b). Figure 9b also shows that there are ap-
proximately eight major peaks in the light transmission distribution, which means
that the wavelength of the banded texture for the Ericksen number shown is close
to the gap thickness. The simulations show that the transmission intensity is peri-
odic and that the maxima occur between two adjacent tubular orientation inversion
walls.

4.7. Variable Order Parameter


The LE theory assumes that the nematic structure is defined only by the director, n,
and that the degree of molecular alignment is uniform. Ericksen (1991) relaxed the
second assumption by introducing a scalar order parameter into the free energy in
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

250 REY DENN


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 9 (a) Visualization of the nz director component for E = 2918 and an applied
shear strain of 240. The long and short axes represent the flow and thickness direc-
tions, respectively. Dark and bright stripes represent nz 1 and nz +1 (orientation
inversion walls), respectively, and a midgray area represents a planar orientation ma-
trix (nz 0). (b) Computed light transmission intensity as a function of dimensionless
length along the flow direction. The intensity is periodic, and the wavelength is of the
order of the shear-cell thickness. From Han & Rey (1995b), with permission of the
American Chemical Society.

an attempt to replicate the behavior of liquid-crystalline polymers. The extended


Ericksen theory has been explored in a variety of rectilinear flows (e.g., Calderer &
Mukherjee 1998, Calderer & Liu 2000), but it is unlikely to offer as much insight
into the flow of LCNPs as extensions of the Doi theory, discussed in the following
sections.

5. DOI THEORY AND EXTENSIONS

Liquid-crystalline polymers exhibit behavior that cannot be described by the LE


theory. There are qualitative differences between lyotropic LCNPs, which un-
dergo an isotropic-nematic transition in solution at a critical concentration, and
thermotropic LCNPs, which undergo an isotropic-nematic transition in the melt
at a critical temperature (but which, in fact, often decompose before reaching the
transition temperature). The backbone chemistry that enables thermotropic LCNPs
to be processed in the melt typically results in molecules that are less stiff than
those that form lyotropic LCNPs, and it is likely that entanglements occur in ther-
motropic LCNPs. Lyotropes often exhibit a three-region viscosity curve like
that shown in Figure 10 (Walker & Wagner 1994): The viscosity in Region I, at
low rates, is shear thinning, with a power-law dependence that is close to 0.5; the
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 251


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 10 Three-region viscosity curve in two samples of a liquid-crystalline solution of


60% hydroxypropylcellulose in water. Data from Walker & Wagner (1994), reprinted with
permission from the Journal of Rheology.

viscosity is insensitive to shear rate in Region II, at intermediate rate, whereas


the viscosity is again shear thinning in Region III. Lyotropes often exhibit an in-
termediate interval, near the transition between Regions II and III, in which the
first normal-stress difference is negative. Three-region viscosity curves are less
obvious in thermotropes, and negative normal stresses are not observed.
Polymeric liquid crystals are highly textured, with a correlation length for ne-
matic order that is typically a few microns in size. The submicron defect structures
are not well understood, but they are believed to play a major role in the rheology at
low and intermediate stresses. Texture cannot be removed in thermotropic LCNPs,
either by shearing or by the application of strong fields. The texture is sensitive
to the intensity and magnitude of shear, as illustrated in Figure 11 for Vectra B,
a co-poly(ester amide) consisting of 60 mole % 6-hydroxy-2-naphthoic acid, 20
mole % hydroxybenzoic acid, and 20 mole % aminophenol, but the initial texture
is recovered upon cessation of shear. Although the free energy minimum should
be a monodomain, LCNPs, especially thermotropes, appear to have a glass-like
energy landscape with deep minima that retain texture. Texture development in the
flow of LCNPs has been probed by a variety of methodologies in the work of Riti
and coworkers (Riti et al. 1997, Riti & Navard 1998) and references cited therein.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

252 REY DENN


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 11 Development of texture in Vectra B at 5000 d/cm2 as viewed under crossed


polarizers at (a) 0, (b) 5, (c) 115, and (d) 500 strain units (Kim 1996).

5.1. Constitutive Equation


The Doi theory is based on an evolution equation for the probability distribution of
the orientation of a suspension of monodisperse rods. (u, R, t) is the probability
of finding a rod at position R in an orientation u. The Smoluchowski equation for
(u) includes the rotational diffusivity D and a nematic mean-field potential V(u),
as follows:

D V(u) Du
=D + . (20)
Dt u u kB T u u Dt
The nematic potential is usually taken as the Maier-Saupe form,

V(u) = 3/2UkB Thuui : uu, (21)

where U is a nondimensional potential, kB the Boltzmann constant, T the temper-


ature, and h. . .i denotes an average over the distribution function :
Z
hy . . . zi = (u, R, t) y(u) . . . z(u) du. (22)
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 253

The evolution equation for u for infinite rods is


Du
= u.v v : uuu. (23)
Dt
The theory is usually written in terms of the order parameter tensor S, defined

1
S = uu I , (24)
d
where d is the dimension of the space (two or three). The evolution equation for S
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

and the equation for the stress involve the fourth moment huuuui, and all variants
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

of the theory in use explicitly or implicitly require a closure.


This set of equations has been solved without closure by Faraoni and coworkers
(1999) for unbounded simple shear by employing an expansion in spherical har-
monics, using a Galerkin approximation with truncation at 10 terms. The solution
shows rich dynamics above a critical value of U, where the isotropic solution for
S becomes unstable. A typical result is shown in Figure 12 for U = 5.33, where
b2,2 is a coefficient in the expansion for and G is the shear rate. Zero values of
the real part of b2,2 correspond to isotropic solutions. The nomenclature describing
the various regimes is intended to be descriptive of the motion of the rods. Log
rolling is an orientation orthogonal to the plane of shear, and it has been observed
for some thermotropic LCNPs (Romo-Uribe & Windle 1996), but the transition
seems to vanish with increasing molecular weight. Grosso et al. (2001) have shown
that chaotic behavior is possible.
The continuum equation for S is obtained by multiplying Equation 20 by uu
and by averaging over the distribution. Using the closure approximation huuuui =
huuihuui results in the equation
DS/Dt = F(S) + G(S), (25)
where

1 1
F(S) = 2DS + 6DU S S + S (S : S)S (S : S)I (26)
d d
and

2 1
G(S) = A + v .S + S.v 2v : S S + I .
T
(27)
d d
The corresponding stress tensor has the form
T = pI (ckB T/2D)F(S) + 2s A, (28)
where c is the concentration of rods and s the viscosity of the suspending fluid.
The rotational diffusion coefficient D is sometimes taken to depend on S. Closure
approximations have been studied by Chaubal et al. (1995), Chaubal & Leal (1998),
and Feng et al. (1998), who recommend the Bingham closure in preference to
quadratic closures.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

254 REY DENN


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 12 Solution diagram for the Doi theory with U = 5.33. Solid lines represent stable
stationary solutions, dashed lines unstable stationary solutions, filled circles stable periodic
solutions, open circles unstable periodic solutions, filled squares stable periodic solutions
with double period, and open squares unstable periodic solutions with double period. From
Faraoni et al. (1999), with permission by the Journal of Rheology.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 255

The Doi theory, which implicitly assumes a monodomain, cannot describe tex-
tures seen in liquid-crystalline polymers. This weakness has motivated a number
of authors to incorporate longer-range interactions through enhanced potentials.
Marrucci & Greco (1991), for example, suggested a potential of the form

1 2 2
V = 2UkB T S + R S + L (uu : S) : uu,
2
(29)
24

where R is a characteristic interaction distance between molecules and L is the


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

molecular length. This potential neglects inhomogeneous terms associated with


the shape of the molecule, which is equivalent to taking the three Frank elas-
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

tic coefficients to be equal. R is expected to be much larger than L. Kupferman


and coworkers (2000), Feng and coworkers (2000), and Lhuillier (2000) have de-
veloped similar continuum formulations based on incorporation of the Marrucci-
Greco potential into Equation 20. Their results, in turn, are essentially equivalent
to an equation obtained by Rey & Tsuji (1998) using the Landaude Gennes ap-
proach and to an equation obtained by Beris & Edwards (1994) using a Poisson
bracket formalism. [The Poisson bracket formalism for a tumbling parameter dif-
ferent from 1 does not obey the Jacobi identity (Kats & Lebedev 1994), and
the equations cannot be rewritten in terms of canonically conjugated variables.]
This formalism introduces terms in 2S into both the evolution equation for S and
the equation for the stress. The continuum theories reduce to the LE theory in the
limits of weak flows and small distortions (Feng et al. 2000), and all predict
the orientation dynamics of the LE equations and of the Doi equation, including
the steady tumbling mode of the LE model (Carlsson 1984) and the tumbling-
wagging-aligning cascade predicted by the Doi equation (Rey & Tsuji 1998,
Kawaguchi 1998).

5.2. Solved Flows


Solutions to the continuum Doi equations have been obtained for viscometric
flows in many of the studies cited above and in extensional flows (Forest et al.
1999). Negative first normal stresses are predicted in steady shear, and the flow
transitions shown in Figure 11 are retained in the averaged equations, although
transition parameters are dependent on the specifics of the closure approxima-
tion used (e.g., Feng et al. 1998). Ramalingam & Armstrong (1993) and Forest
et al. (1997) have solved the thin-filament equations for the isothermal spinning
of liquid-crystalline polymer fibers, whereas Mori et al. (1997b) have solved the
two-dimensional axisymmetric equations. Mori and coworkers found an unusual
concavity in the computed velocity profile that had been observed experimentally
(Mori et al. 1997a). Feng and coworkers have solved the Doi equations with a vari-
ety of closure approximations for time-dependent flow between rotating eccentric
cylinders, and they report the development of director singularities of order 1/2
in some cases. There are significant quantitative differences between the various
closures, although all give qualitatively similar results.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

256 REY DENN


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Figure 13 Simulation of transient shear flow for a fluid with a Marrucci-Greco potential
at De = 1, E = 400. (a) The structure-tensor component Sxx(y, t), where y is the coordinate
between the plates; (b) the velocity component vx(y, t); (c) visualization of the structure tensor
during one cycle; and (d) the velocity profile over one cycle. From Kupferman et al. (2000),
reproduced with permission by Elsevier Science.

Kupferman and coworkers (2000) have reported the development of structure


in two-dimensional transient shear flow using the continuum formulation with
the Marrucci-Greco potential. Figure 13 shows some results for the velocity and
orientation fields at a Deborah number of unity and an Ericksen number of 400.
Under these conditions there is no steady state. The director tumbles in the middle
section of the channel, generating smooth orientational waves. The outermost
defect line is periodically annihilated as it is squeezed between the propagating
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 257

wave and the anchoring wall. Velocity gradients are very large. Their results show
that it is essential to incorporate coupling between the velocity and structure fields.
One disappointing feature of this study was that structure with a length scale
independent of the macroscopic dimension did not emerge, and the transverse
orientational patterns are reminiscent of those obtained from the LE theory (Chono
et al. 1998).

5.3. Textured Systems


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

Larson & Doi (1991) attempted to deal with the presence of texture in LCPs by
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

formulating a mesoscopic theory in which the textured fluid is assumed to be


composed of domains in which the director orientation is uniform. Field equa-
tions for the stress and orientation are obtained by spatial averaging of the LE
equations over regions large relative to the domain size, with phenomenological
evolution equations for mesoscopic parameters. Kawaguchi & Denn (1999) de-
veloped a mesoscopic theory for textured fluids that reduces to the Larson-Doi
theory at lowest order, but with a different meaning for the model parameters.
Ugaz et al. (2001) developed a polydomain Ericksen model that is similar to
the Larson-Doi model when contributions from director elasticity are neglected.
Rey (1996) developed a coarse-grained model using the tensor order parameter
that predicts entropy-driven stable textures and showed that under certain condi-
tions it converges to the Larson-Doi model. Burghardt and coworkers (Bedford &
Burghardt 1996; Burghardt 1998; Cinader & Burghardt 2000a,b) have carried out
a series of experimental studies of LCP structure development in channel flows
with expansions and contractions, together with simulations using the Larson-Doi
theory. The general conclusion is that the theory captures some of the qualitative
features of the orientation development but is quantitatively inadequate.
Several other models aim at explaining the rheological implications of collective
defect behavior in nematic polymers (Wissbrun 1985, Marrucci 1991, Yamakazi
et al. 1991). Detailed descriptions of defect annihilations, reactions, defect core
transitions, and other local phenomena are not captured because the averaging is
over distances orders of magnitude greater than defect cores. Rey (1993b) incor-
porated the loop emission model of de Gennes (1976) into population balance
equations to describe the birth, death, and deformation of the loop population in
the presence of arbitrary flows. The predictions for steady and oscillatory shear
were shown to be consistent with the experiments of Graziano & Mackley (1984)
and Alderman & Mackley (1985). Rey & Tsuji (1998) used the Landaude Gennes
model to study defect nucleation and coarsening in the presence of shear flow. The
sequence of phase transition texture creation texture coarsening was simu-
lated when shear flows of various strengths were activated at different stages of
the process cascade. They found that flow speeds up coarsening by optimizing the
defect annihilation process.
The only three-dimensional flow visualization study of a LCP in a complex
geometry seems to be that of Kawaguchi & Denn (1997), who used tracers to
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

258 REY DENN

follow the velocity profile of thermotropic Vectra A through a conical contraction


between two cylindrical channels. They observed a breakdown of simple streamline
flow, with an asymmetric flow in the contraction and regions that appear to be
stagnant. It is likely that substantial changes in microstructure and material occur
over a small spatial region, possibly because of changes in texture. Gentzler et al.
(1998) used solid-state nuclear magnetic resonance (NMR) spectroscopy to image
director orientation in frozen samples from a contraction flow of this polymer. The
technique showed orientational differences in different parts of the samples, but the
volumes were too large to resolve meaningful structural information. A recent study
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

by Gentzler et al. (2000) used NMR to study velocity and orientation development
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

in situ for tube flow of a lyotropic LCNP, and it is likely that investigators will find
uses for NMR in geometries that cannot be probed by optical techniques because
of the opacity of the liquids.

6. INTERFACIAL EFFECTS

Interfaces involving NLCs are characterized by nematic ordering as well as by the


geometry, with the result that even the interfacial tension of a planar interface may
change owing to changes in the nematic ordering in textured materials. The study
of nematic interfaces is just beginning.

6.1. Interfaces in LC Blends


Liquid-crystalline materials are employed as the dispersed phase in immiscible
blends in a variety of technologies. Low molar-mass liquid crystals are used in
display technology (Drzaic 1995), whereas polymeric liquid crystals are used as
flow modifiers to reduce the pressure drop in extrusion (Cogswell 1983) and as
the fibrous phase in self-reinforced composites (Weiss et al. 1987, Handlos &
Baird 1995, Qin 1996). Liquid-crystalline polymers are also of interest as barrier
layers in multilayer films (Sawyer et al. 1998). The interface between LCs and
isotropic phases has not been extensively studied, but limited observations suggest
a strong effect of the LC anisotropy. Droplets smaller than the correlation length
for nematic order do not seem to contribute to the linear viscoelasticity in poly-
mer blends where the LCNP is a dispersed phase, for example, whereas larger
droplets contribute as expected (Lee & Denn 1999, Riise et al. 1999). The relax-
ation dynamics of LCNP droplets following a step strain scale differently from
those of a flexible polymer in the same suspending fluid (Lee & Denn 2000) and
show an unexpected dependence on the initial strain. Kernick & Wagner (1999)
have suggested that the interfacial tension might be affected by morphological
changes induced during droplet deformation, and they employ a theory of van
Oene (1972) regarding the effect of normal stresses on interfacial tension to argue
that the parallel orientation induced by droplet stretching may reduce the interfacial
tension.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 259

Li & Denn (2001) have carried out a Monte Carlo simulation of the inter-
face between a liquid-crystalline polymer and a flexible polymer using the three-
dimensional Bond Fluctuation Method (Carmesin & Kremer 1988, Deutsch &
Binder 1990), from which the equilibrium interfacial tension, the interfacial bend-
ing elasticity, and the chain configurations can be extracted. Two far-field orienta-
tions were assumed in the calculations, parallel to the interface and orthogonal to
the interface (homeotropic). The interface was more diffuse for the homeotropic
orientation because of the easier penetration of chain ends into the isotropic phase.
Some chains in the LCNP phase moved out of plane and adopted a parallel ori-
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

entation near the interfacial plane, and a parallel orientation was induced in the
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

isotropic polymer over a length of the order of one radius of gyration. The computed
equilibrium interfacial tension for the parallel far-field orientation was consider-
ably larger than that for the homeotropic orientation. This computation suggests
an effect opposite to that proposed by Kernick & Wagner (1999).

6.2. Interfacial Viscoelasticity


Interfacial viscoelastic theories of nematic liquid crystals using the LE and Landau
de Gennes formalisms have recently been formulated (Rey 2000, 2001) and used
to analyze static and dynamical phenomena. Static interfacial phenomena include
disjoining pressures and contact angles in nematic films. Nonequilibrium phenom-
ena include wetting and spreading, Marangoni flows, dynamic interfacial tension,
interfacial Miesowiczs viscosities, and interfacial dilational viscosities. The ne-
matic Rayleigh fiber instability has also been analyzed (Rey 1997). We limit the
discussion here to the disjoining pressure in nematic films and Marangoni surface
flows predicted by the interfacial LE equations.
The interfacial LE equations consist of the interfacial linear momemtum balance
equation and the interfacial torque balance equation, the latter given by Equations 6
and 7. The interfacial linear momentum balance equation for a nematic (N)-viscous
fluid (A) interface is

k.(tA tN ) = s .tse + s .tsv , (30)

where tN is the total stress tensor in the nematic phase at the N/A interface, tA
is the total stress tensor in the isotropic fluid phase, tse is the elastic surface stress
tensor, tsv is the viscous extra surface stress tensor, s is the surface gradient, and
k is the unit normal. The elastic stress tensor tse is a 2 3 tensor whose gradients
represent tangential and normal elastic forces. The surface elastic stress tensor is
given by the usual 2 2 symmetric interfacial tension contribution tsen (normal
stresses) and the 2 3 anisotropic contribution tseb (bending stresses):

an dan
tse = tsen + tseb ; tsen = Fs Is ; tseb = Is . k = Is . nk . (31)
k d(n.k)
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

260 REY DENN

The physical significance of the normal and bending components of the surface
stress tensor is clearer when one considers the net surface forces engendered by
their surface gradients:

dan
f = s .t = k.(s n)T .Is + {2Hfs }k
d(n.k)
| {z }
surface gradients in normal stresses

an an
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

+ 2H .k s . k , (32)
k k
| {z }
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

surface gradients in bending stresses

where H = s .k/2 is the average curvature. The first term on the right is
the tangential nematic Marangoni force, which arises from orientation gradients
and drives interfacial flows. The remaining terms comprise the surface normal
force, which contributes to the disjoining pressure and plays a fundamental role
in thin-film stability. For a planar (H = 0) interface, the normal force may tend
to stabilize or destabilize a nematic thin film, depending on the sign of surface
gradients and the nature of the anisotropic anchoring energies. Stabilization of thin
films by anisotropic surface energy has practical utility in the use of liquid-crystal
surfactants (Larson 1999). The LE and Landaude Gennes interfacial theories and
the equilibrium molecular simulations have not yet been integrated.

7. CONCLUDING REMARKS
The LE theory has been successful in elucidating the flow of low molar-mass
nematics in simple geometries. The theory has not been tested against relevant
experiments in complex flow fields, but complex flow fields are not likely to be
encountered in applications of LMMLCs. The theoretical framework for the flow
of nematic liquid crystals is still evolving. Various extensions of the Doi theory that
differ in final form only in detail capture many of the qualitative features of the flow
of LCNPs in simple geometries. These theories have not been shown to predict
texture development in flow; hence, they cannot be expected to predict behavior
in processing flows. Mesoscopic theories based on spatial averaging capture only
some qualitative features of nonrectilinear LCNP flow. Interfacial effects in liquid-
crystalline systems have just started to receive attention. The outstanding fluid-
dynamical challenge is the development of a more complete theoretical basis for the
description of texture evolution in flow and its consequences in complex flow fields.
The development of a rational theory of blends for immiscible systems containing a
liquid-crystalline phase, incorporating new developments in describing interfacial
effects, is another important challenge, but real progress for polymeric systems
will require an adequate description of the dynamic structure distribution in the
LC phase.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 261

ACKNOWLEDGMENTS
A.D.R. acknowledges support from the Natural Sciences and Engineering Re-
search Council (Canada), Air Force Office of Scientific Research-Mathematical
Directorate, the Donors of The Petroleum Research Fund (American Chemical So-
ciety), and the NSF Center for Advanced Fibers and Films at Clemson University.

Visit the Annual Reviews home page at www.AnnualReviews.org


Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

LITERATURE CITED
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Alderman NJ, Mackley MR. 1985. Optical tex- Burghardt WR, Fuller GG. 1990. Transient
tures observed during the shearing of ther- shear of nematic liquid crystals: manifesta-
motropic liquid-crystal polymers. Faraday tions of director tumbling. J. Rheol. 34:959
Discuss. Chem. Soc. 79:14960 92
Anisimov SI, Dzyaloshinskii IE. 1972. A new Calderer MC, Liu C. 2000. Poiseuille flow of
type of disclination in liquid crystals and nematic liquid crystals. Int. J. Eng. Sci. 38:
the stability of disclinations of various types. 100722
Sov. Phys. JETP 36:77479 Calderer MC, Mukherjee B. 1998. Some
Bedford BD, Burghardt WR. 1996. Molecu- mathematical issues in the modeling of flow
lar orientation of a liquid-crystalline polymer phenomena of polymeric liquid crystals. J.
solution in mixed shear-extensional flows. J. Rheol. 42(6):151936
Rheol. 40:23557 Carlsson T. 1984. Theoretical investigation of
Bedford SE, Nicholson TM, Windle AH. the shear flow of nematic liquid crystals with
1991. A supra-molecular approach to the the Leslie viscosity 3 > 0: hydrodynamic
modeling of textures in liquid crystals. Liq. analogue of first order phase transitions. Mol.
Cryst. 10:6372 Cryst. Liq. Cryst. 104:30734
Bedford SE, Windle AH. 1993. Modeling of Carmesin I, Kremer K. 1988. The bond fluctu-
microstructure in mesophases. Liq. Cryst. ation method: a new effective algorithm for
15:3163 the dynamics of polymers in all spatial di-
Beris AN, Edwards BJ. 1994. Thermodynam- mensions. Macromolecules 21:281923
ics of Flowing Systems with Internal Mi- Chaubal CV, Leal LG. 1998. A closure approx-
crostructure. New York: Oxford Univ. Press. imation of liquid crystalline polymer mod-
683 pp. els based on parametric density estimation.
Boden N. 1994. Micellar liquid crystals. In J. Rheol. 42:177201
Micelles, Membranes, Microemulsions, and Chaubal CV, Leal LG, Fedrickson GH. 1995.
Monolayers, ed. WM Gelbart, A Ben-Shaul, A comparison of closure approximations for
D Roux, pp. 153211. New York: Springer- the Doi theory of LCPs. J. Rheol. 39:79103
Verlag. 608 pp. Chono S, Tsuji T, Denn MM. 1994. Numeri-
Bouligand Y. 1998. Defect and textures. In cal simulation of planar contraction flow of
Handbook of Liquid Crystals, ed. D Demus, nematic liquid crystals (in Japanese). Trans.
J Goodby, GW Gray, H-W Spiess, V Vill, Jpn. Soc. Mech. Eng. 60:194450
1:40653. Chichester, NY: Wiley-VCH Chono S, Tsuji T, Denn MM. 1998. Spa-
Burghardt WR. 1998. Molecular orientation tial development of director orientation of
and rheology in sheared lyotropic liquid crys- tumbling nematic liquid crystals in pressure-
talline polymers. Macromol. Chem. Phys. driven channel flow. J. Non-Newton. Fluid
199:47188 Mech. 79:51527
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

262 REY DENN

Chuang I, Durrer R, Turok N, Yurke B. 1991. Ericksen JL. 1991. Liquid crystals with variable
Cosmology in the laboratory: defect dynam- degree of orientation. Arch. Rat. Mech. Anal.
ics in liquid crystals. Science 251:133642 113:97120
Ciferri A, ed. 1991. Liquid Crystallinity in Faraoni V, Grosso M, Crescitelli S, Maffet-
Polymers: Principles and Fundamental Pro- tone PL. 1999. The rigid-rod model for ne-
perties. New York: VCH. 438 pp. matic polymers: an analysis of the shear flow
Cinader DK Jr, Burghardt WR. 2000a. Mole- problem. J. Rheol. 43:82943
cular orientation in channel flows of main- Feng J, Chaubal CV, Leal LG. 1998. Closure
chain thermotropic liquid crystalline poly- approximations for the Doi theory: Which to
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

mers. Rheol. Acta 39:24758 use in simulating complex flows of liquid-


Cinader DK Jr, Burghardt WR. 2000b. Poly- crystalline polymers? J. Rheol. 42:1095
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

domain model predictions of liquid crys- 119


talline polymer orientation in mixed shear/ Feng J, Sgalari G, Leal LG. 2000. A theory for
extensional channel flows. Rheol. Acta 39: flowing nematic polymers with orientational
25970 distortion. J. Rheol. 44:1085101. Erratum.
Cladis PE, Torza S. 1975. Stability of ne- 2000. J. Rheol. 44:1435
matic liquid crystals in Couette flow. Phys. Forest MG, Wang Q, Bechtel SE. 1997. One
Rev. Lett. 35:128386 dimensional isothermal spinning models for
Cogswell FN. 1983. Compositions of melt- liquid crystalline polymer fibers. J. Rheol.
processible polymers having improved pro- 41:82150
cessibility. US Patent No. 4386174 Forest MG, Wang Q, Zhou H. 1999. Non-
de Gennes PG. 1976. Nematodynamics. In homogeneous patterns with core defects in
Molecular Fluids, ed. R Balian, G Weil, pp. elongational flows of liquid crystal polymers.
377400. London: Gordon & Breach. 459 pp. J. Rheol. 43:157382
de Gennes PG, Prost J. 1993. The Physics of Gentzler M, Patil S, Reimer JA, Denn MM.
Liquid Crystals. London: Oxford Univ. 1998. Molecular motion and orientation dis-
Press. 597 pp. 2nd ed. tributions in melt-processed, fully aromatic
Demus D, Goodby J, Gray GW, Spiess HW, liquid crystalline polymers from 1H NMR.
Vill V. 1999. Physical Properties of Liquid Solid State Nucl. Magn. Reson. 12:97112
Crystal. Weinheim: Wiley-VCH. 503 pp. Gentzler M, Song YQ, Muller SJ, Reimer JA.
Demus D, Richter L. 1978. Textures of Liq- 2000. Quantitative NMR velocity imaging of
uid Crystals. Leipzig, Ger.: VEB Dtsch. Verl. a main-chain liquid crystalline polymer flow-
Grundst. ing through an abrupt contraction. Rheol.
Deutsch HP, Binder K. 1991. Interdiffusion Acta 39:112
and self-diffusion in polymer mixtures: a Graziano DJ, Mackley MR. 1984. Disclina-
Monte Carlo study. J. Chem. Phys. 94:2294 tions observed during the shear of MBBA.
304 Mol. Cryst. Liq. Cryst. 106:10319
Donald AM, Viney C, Windle AH. 1983. Greco F. 1989. Model predictions of small-
Banded structures in oriented thermotropic angle light scattering from films of nematic
polymers. Polymer 24:15559 liquid crystalline polymers. Macromolecules
Donald AM, Windle AH. 1992. Liquid Crys- 22:462227
talline Polymers. Cambridge: Cambridge Grosso M, Keunings R, Crescitelli S, Maffet-
Univ. Press. 310 pp. tone PL. 2001. Prediction of chaotic dynam-
Drzaic PS. 1995. Liquid Crystal Dispersions. ics in sheared liquid crystalline polymers.
Singapore: World Sci. Phys. Rev. Lett. 86:318487
Elliott A, Ambrose EJ. 1950. Evidence of chain Gu D-F, Jamieson AM. 1994. Shear deforma-
folding in polypeptides and proteins. Dis- tion of homeotropic monodomains: tempe-
cuss. Faraday Soc. 9:24651 rature dependence of stress response for
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 263

flow-aligning and tumbling nematics. J. modeling of liquid crystalline polymers. PhD


Rheol. 38:55571 thesis. Univ. Calif., Berkeley
Han WH, Rey AD. 1993. Supercritical bifur- Kawaguchi MN, Denn MM. 1997 Visualiza-
cations in simple shear flow of a non-aligning tion of the flow of a thermotropic liquid
nematic: reactive parameter and director an- crystalline polymer in a tube with a coni-
choring effects. J. Non-Newton. Fluid Mech. cal contraction. J. Non-Newton. Fluid Mech.
48:181210 69:20719
Han WH, Rey AD. 1994. Orientation sym- Kawaguchi MN, Denn MM. 1999. A meso-
metry breakings in shearing liquid crystals. scopic theory of liquid crystalline polymers.
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

Phys. Rev. E 50:168891 J. Rheol. 43:11124


Han WH, Rey AD. 1995a. Simulation and Kernick WA III, Wagner NJ. 1999. The role
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

validation of temperature effects on the ne- of liquid-crystalline polymer rheology on the


matorheology of aligning and nonaligining evolving morphology of immiscible blends
liquid crystals. J. Rheol. 39:30122 containing liquid-crystalline polymers. J.
Han WH, Rey AD. 1995b. Theory and simu- Rheol. 43:52149
lation of optical banded textures of nematic Kibble TWB. 1976. Topology of cosmic do-
polymers during shear flow. Macromolecules mains and strings. J. Phys. A 9:138798
28:84015 Kim EG. 1996. Shear rheology and texture of a
Handlos AA, Baird DG. 1995. Processing and thermotropic liquid crystalline polymer. MS
associated properties of in situ compos- thesis. Univ. Calif., Berkeley
ites based on thermotropic liquid crystalline Kimura T, Gray DG. 1993. Annealing method
polymers and thermoplastics. J. Macromol. for modeling liquid crystal textures. Macro-
Sci. Rev. Macromol. Chem. C 35:183238 molecules 26:345556
Harrison P, Navard P. 1999. Investigation of Kiss G, Porter RS. 1980. Rheo-optical studies
the band texture occurring in hydroxypropyl of liquid crystalline solutions of helical poly-
cellulose solutions using rheo-optical, rheo- peptides. Mol. Cryst. Liq. Cryst. 60:26780
logical and small angle light scattering tech- Kleman M. 1983. Points, Lines and Walls in
niques. Rheol. Acta 38:56993 Liquid Crystals, Magnetic Systems, and Var-
Hess S. 1975. Irreversible thermodynamics ious Ordered Media. Chichester, NY: Wiley.
of nonequilibrium alignment phenomena in 322 pp.
molecular liquids and liquid crystals. Z. Kleman M. 1989. Defects in liquid crystals.
Naturforsch. Teil A 30:72838 Rep. Prog. Phys. 52:555654
Hiltrop K, Fisher F. 1976. Radial poiseuille Kleman M. 1991. Defects and textures in liq-
flow of a homeotropic nematic LC layer. Z. uid crystalline polymers. See Ciferri 1991,
Naturforsch. Teil A 31:8007 pp. 36594
Hongladarom K, Burghardt WR. 1998. Mole- Kupferman R, Kawaguchi MN, Denn MM.
cular orientation, Region I shear thinning 2000. J. Non-Newton. Fluid Mech. 91:255
and the cholesteric phase in aqueous hy- 71
dropropylcellulose under shear. Rheol. Acta Larson RG. 1999. The Structure and Rheology
37:4653 of Complex Fluids. New York: Oxford Univ.
Hudson SD, Larson RG. 1993. Monte Carlo Press. 663 pp.
simulation of a disclination core in nematic Larson RG, Doi M. 1991. Mesoscopic domain
solutions of rodlike molecules. Phys. Rev. theory for textured liquid-crystalline poly-
Lett. 70:291619 mers. J. Rheol. 35:53963
Kats EI, Lebedev VV. 1994. Fluctuational Larson RG, Mead DW. 1992. Development
Effects in the Dynamics of Liquid Crystals. of orientation and texture during shearing
New York: Springer-Verlag. 70 pp. of liquid-crystalline polymers. Liq. Cryst.
Kawaguchi MN. 1998. Flow visualization and 12:75168
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

264 REY DENN

Larson RG, Mead DW. 1993. The Ericksen Jamieson AM. 1997. The origin of stress-
number and Deborah number cascades in oscillation damping during start-up and re-
sheared polymeric nematics. Liq. Cryst. 15: versal of torsional shearing of nematics.
15169 Rheol. Acta 36:48597
Lee HS, Denn MM. 1999. Rheology of a visco- Mori N, Hamaguchi Y, Nakamura K. 1997a.
elastic emulsion with a liquid crystalline Measuremenmt of velocity profile develop-
polymer dispersed phase. J. Rheol. 43:1583 ment in the spinning flow of liquid crystalline
98 polymer solutions. J. Rheol. 41:23747
Lee HS, Denn MM. 2000. The deformation Mori N, Hamaguchi Y, Nakamura K. 1997b.
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

and retraction of thermotropic LCP droplets Numerical simulation of the spinning flow
in a flexible polymer matrix. J. Non-Newton. of liquid crystalline polymers. J. Rheol. 41:
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Fluid. Mech. 93:31523 1095104


Lee S-D, Meyer RB. 1991. Elastic and vis- Navard P. 1986. Formation of band textures
cous properties of lyotropic polymer nemat- in hydroxypropyl-cellulose liquid crystals. J.
ics. See Ciferri 1991, pp. 34364 Polym. Sci.: Phys. Ed. 24:43542
Leslie FM. 1983. Some topics in continuum the- Noel C. 1998. Defects and textures in nematic
ory of nematics. Philos. Trans. Soc. London main-chain liquid crystalline polymers. In
Ser. A 309:15565 Handbook of Liquid Crystals, ed. D Demus,
Lhuillier D. 2000. Continuum Thermodynam- J Goodby, GW Gray, H-W Spiess, V Vill,
ics: The Art and Science of Modeling Mat- 3:93120. Chichester, NY: Wiley-VCH
ters Behaviour, ed. GA Maugin, R Drouot, Noel C, Navard P. 1991. Liquid crystal poly-
FS Sidoroff, pp. 23746. Dordrecht: Kluwer mers. Prog. Polym. Sci. 16:55110
Li X, Denn MM. 2001. Influence of bulk ne- ORourke MJE, Thomas EL. 1995. Morphol-
matic orientation on the interface between a ogy and dynamic interaction of defects in
liquid crystalline polymer and a flexible poly- polymer liquid crystals. MRS Bull. 20(No.
mer. Phys. Rev. Lett. 86:65659 9):2938
Liu PY, Jamieson AM. 2000. Twist viscosity of Park JO, Srinivasarao M, Rey AD. 2001.
mixtures of low molar mass nematics. Rheol. Forced wetting of nematic fluids on cylindri-
Acta 39:53241 cal objects. 72nd Annu. Meet., Soc. Rheol.,
Luskin M, Pan TW. 1992. Non-planar shear Hilton Head, Pap. PF9:26
flows for non-aligning nematic liquid crys- Pieranski P, Guyon E. 1974. Two shear flow
tals. J. Non-Newton. Fluid Mech. 42:369 regimes in nematic p-n-hexyloxybenzili-
84 dene-p0 -aminobenzonitrile. Phys. Rev. Lett.
Marrucci G. 1991. Rheology of nematic poly- 32:92426
mers. See Ciferri 1991, pp. 395422 Qin Y. 1996. A literature review on the in situ
Marrucci G, Greco F. 1991. The elastic con- generation of reinforcing fibers. Polym. Adv.
stants of Maier-Saupe rodlike molecular Tech. 7:15159
nematics. Mol. Cryst. Liq. Cryst. 201:17 Ramalingam S, Armstrong RC. 1993. Analysis
30 of isothermal spinning of liquid crystalline
Marrucci G, Greco F. 1993. Flow behavior polymers. J. Rheol. 37:114169
of liquid crystalline polymers. Adv. Chem. Rey AD. 1990a. Defect controlled dynamics of
Phys. 86:331404 nematic liquids. Liq. Cryst. 7:31534
Mather PT, Pearson DS, Larson RG. 1996. Rey AD. 1990b. Radial creeping flow of rod-
Flow patterns and disclination-density mea- like nematic liquid crystals. J. Rheol. 34:425
surements in sheared nematic liquid crystals. 67
1: Flow-aligning 5CB. Liq. Cryst. 20:527 Rey AD. 1991. Periodic textures of nematic
38 polymers and orientational slip. Macro-
Mather PT, Pearson DS, Larson RG, Gu DF, molecules 24:445056
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

LIQUID-CRYSTALLINE MATERIALS 265

Rey AD. 1993a. Macroscopic modeling of ling alignment in main-chain thermotropic


dynamic phenomena in liquid crystalline ma- liquid crystalline polymer melts under shear:
terials. Adv. Transp. Process. 9:185229 an in-situ WAXS study. Macromolecules
Rey AD. 1993b. Analysis of shear flow effects 29:624655
on liquid crystalline textures. Mol. Cryst. Liq. Sawyer LC, Jaffe M. 1986. The structure of
Cryst. 225:31335 thermotropic polyesters. J. Mater. Sci. 21:
Rey AD. 1996. Phenomenological theory of 18971913
textured mesophase polymers in weak flows. Sawyer LC, Linstid HC, Romer M. 1998. Em-
Macromol. Theory Simul. 5:86376 erging applications for neat LCPs. Plast. Eng.
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

Rey AD. 1997. Theory of break-up dynamics of 54(12):3741


liquid crystalline fibers. J. Phys. II 7:1001 Schopohl N, Sluckin TJ. 1987. Defect core
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

11 structure in nematic liquid crystals. Phys.


Rey AD. 2000. Viscoelastic theory for nematic Rev. Lett. 59:258284
interfaces. Phys. Rev. E 61:154049 Schopohl N, Sluckin TJ. 1988. Hedgehog
Rey AD. 2001. Irreversible thermodynamics structure in nematic and magnetic system. J.
of liquid crystal interfaces. J. Non-Newton. Phys. Fr. 49:1097101
Fluid Mech. 96:4562 Shiwaku T, Nakai A, Hasegawa H, Hashi-
Rey AD, Denn MM. 1988. Jeffrey-Hamel moto T. 1990. Ordered structure of ther-
flows of Leslie-Ericksen liquids. J. Non- motropic liquid-crystal polymer. 1. Charac-
Newton. Fluid Mech. 27:375401 terization of liquid-crystal domain texture.
Rey AD, Denn MM. 1989. Converging flow of Macromolecules 23:159099
tumbling nematic liquid crystals. Liq. Cryst. Slaney AJ, Takatoh K, Goodby JW. 1998.
4:25372 Defect textures in liquid crystals. In The
Rey AD, Denn MM. 1990. Analysis of transi- Optics of Thermotropic Liquid Crystals, ed.
ent periodic textures in nematic polymers. S Elston, R Sambles, pp. 30772. London:
Liq. Cryst. 4:40919 Taylor & Francis
Rey AD, Tsuji T. 1998. Recent advances in the- Sonin AA. 1995. The Surface Physics of Liq-
oretical liquid crystal rheology. Macromol. uid Crystals. Amsterdam: Gordon & Breach.
Theory Simul. 7:62339 180 pp.
Rieger J. 1990. Dynamics of disclinations in Srinivasarao M. 1995. Rheology and rheo-
nematic liquid-crystal. Main-chain polymer optics of polymer liquid crystals. Int. J. Mod.
films. Macromolecules 23:154547 Phys. B 9:251572
Riise BL, Mikler N, Denn MM. 1999. Rheol- Srinivasarao M, Berry GC. 1991. Rheo-optical
ogy of a liquid crystalline polymer dispersed studies on aligned nematic solutions of a rod
in a flexible polymer matrix. J. Non-Newton. like polymer. J. Rheol. 35:37997
Fluid Mech. 86:314 Ternet DJ, Larson RG, Leal LG. 1999. Flow-
Riti JB, Cidade MT, Godhino MH, Martins aligning and tumbling in small-molecule liq-
AF, Navard P. 1997. Shear induced textures uid crystals: pure components and mixtures.
of thermotropic acetoxypropylcellulose. J. Rheol. Acta 38:18397
Rheol. 41:124760 Toth WJ, Tobolsky AV. 1970. The synthetic
Riti JB, Navard P. 1998. Textures during recoil chemistry of Bis-5-oxazolones. Polym. Lett.
of anisotropic hydroxypropylcellulose solu- 8:53740
tion. J. Rheol. 42:22537 Trebin HR. 1998. Defects in liquid crystals and
Rojstaczer S, Hsiao BS, Stein RS. 1988. Tex- cosmology. Liq. Cryst. 24:12730
ture formation in liquid crystalline polymers. Tsuji T, Rey AD. 1998. Long range order in
Div. Polym. Am. Chem. Soc. Polym. Preprints sheared liquid crystalline materials. Macro-
29:48687 mol. Theory Simul. 7:62339
Romo-Uribe A, Windle AH. 1996. Log rol- Ugaz V, Cinader DK, Burghardt WR. 1997.
29 Oct 2001 17:38 AR AR151-10.tex AR151-10.SGM ARv2(2001/05/10) P1: GJC

266 REY DENN

Origins of region I shear thinning in model talline polymer. Polym. Eng. Sci. 27:684
lyotropic liquid crystalline polymers. Macro- 91
molecules 30:152730 Wissbrun KF. 1985. A model for domain flow
Ugaz VM, Burghardt WR, Zhou W, Kornfield of liquid-crystal polymer. Faraday Discuss.
JA. 2001. Transient molecular orientation Chem. Soc. 79:16173
and rheology in flow aligning thermotro- Yamazaki Y, Holz A, Edwards SF. 1991. Shear
pic liquid crystalline polymers. J. Rheol. thinning in polymeric liquid crystals. Phys.
45:102963 Rev. A 43:546382
van Oene H. 1972. Modes of dispersion of vis- Yan J, Labes MM. 1992. Control of the aniso-
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

coelastic fluids in flow. J. Colloid Interface tropic mechamical properties of liquid crys-
Sci. 200:8694 tal polymer films by variation in their banded
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

Vermant J, Moldenaers P, Picken SJ, Mewis texture. Macromolecules 25:579093


J. 1994. A comparison between texture and Yan NX, Labes MM. 1994. Critical behav-
rheological behavior of lyotropic liquid crys- ior of shear-induced transient periodic struc-
talline polymers during flow. J. Non-Newton. tures in a lyotropic liquid crystalline polymer
Fluid Mech. 53:123 as a function of molecular weight. Macro-
Walker L, Wagner N. 1994. Rheology of re- molecules 27:784345
gion flow in a lyotropic liquid-crystal poly- Zimmer JE, White JL. 1982. Disclination
mer: the effects of defect texture. J. Rheol. structures in the carbonaceous mesophase.
38:152547 Adv. Liq. Cryst. 5:157211
Weiss RA, Huh W, Nicolais L. 1987. Novel Zuniga I, Leslie FM. 1989. Shear-flow insta-
reinforced polymers based on blends of bilities in non-aligning nematic liquid crys-
polystyrene and a thermotropic liquid crys- tals. Europhys. Lett. 9:68993
P1: FDS
November 5, 2001 10:30 Annual Reviews AR151-FM

Annual Review of Fluid Mechanics


Volume 34, 2002

CONTENTS
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

FRONTISPIECE xii
MILTON VAN DYKE, THE MAN AND HIS WORK, Leonard W. Schwartz 1
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

G.K. BATCHELOR AND THE HOMOGENIZATION OF TURBULENCE,


H.K. Moffatt 19
DAVID CRIGHTON, 19422000: A COMMENTARY ON HIS CAREER AND
HIS INFLUENCE ON AEROACOUSTIC THEORY, John E. Ffowcs Williams 37
SOUND PROPAGATION CLOSE TO THE GROUND, Keith Attenborough 51
ELLIPTICAL INSTABILITY, Richard R. Kerswell 83
LAGRANGIAN INVESTIGATIONS OF TURBULENCE, P.K. Yeung 115
CAVITATION IN VORTICAL FLOWS, Roger E.A. Arndt 143
MICROSTRUCTURAL EVOLUTION IN POLYMER BLENDS, Charles L.
Tucker III and Paula Moldenaers 177
CELLULAR FLUID MECHANICS, Roger D. Kamm 211
DYNAMICAL PHENOMENA IN LIQUID-CRYSTALLINE MATERIALS,
Alejandro D. Rey and Morton M. Denn 233
NONCOALESCENCE AND NONWETTING BEHAVIOR OF LIQUIDS, G. Paul
Neitzel and Pasquale DellAversana 267
BOUNDARY-LAYER RECEPTIVITY TO FREESTREAM DISTURBANCES,
William S. Saric, Helen L. Reed, and Edward J. Kerschen 291
ONE-POINT CLOSURE MODELS FOR BUOYANCY-DRIVEN TURBULENT
FLOWS, K. Hanjalic 321
WALL-LAYER MODELS FOR LARGE-EDDY SIMULATIONS, Ugo Piomelli
and Elias Balaras 349
FILAMENT-STRETCHING RHEOMETRY OF COMPLEX FLUIDS, Gareth H.
McKinley and Tamarapu Sridhar 375
MOLECULAR ORIENTATION EFFECTS IN VISCOELASTICITY, Jason K.C.
Suen, Yong Lak Joo, and Robert C. Armstrong 417
THE RICHTMYER-MESHKOV INSTABILITY, Martin Brouillette 445
SHIP WAKES AND THEIR RADAR IMAGES, Arthur M. Reed and Jerome
H. Milgram 469

vi
P1: FDS
November 5, 2001 10:30 Annual Reviews AR151-FM

CONTENTS vii

SYNTHETIC JETS, Ari Glezer and Michael Amitay 503


FLUID DYNAMICS OF EL NINO VARIABILITY, Henk A. Dijkstra and
Gerrit Burgers 531
INTERNAL GRAVITY WAVES: FROM INSTABILITIES TO TURBULENCE,
C. Staquet and J. Sommeria 559

INDEXES
Subject Index 595
Access provided by INFLIBNET N-LIST Colleges Programme on 04/04/17. For personal use only.

Cumulative Index of Contributing Authors, Volumes 134 627


Cumulative Index of Chapter Titles, Volumes 134 634
Annu. Rev. Fluid Mech. 2002.34:233-266. Downloaded from www.annualreviews.org

ERRATA
An online log of corrections to the Annual Review of Fluid Mechanics chapters
may be found at http://fluid.annualreviews.org/errata.shtml

Você também pode gostar