You are on page 1of 218

Crystals, Liquid Crystals and Superuid Helium on Curved

Surfaces.
A dissertation presented
by
Vincenzo Vitelli
to
The Department of Physics
in partial fulllment of the requirements
for the degree of
Doctor of Philosophy
in the subject of
Physics
Harvard University
Cambridge, Massachusetts
May 2006
c _2006 - Vincenzo Vitelli
All rights reserved.
Thesis advisor Author
David R. Nelson Vincenzo Vitelli
Crystals, Liquid Crystals and Superuid Helium on Curved Surfaces.
Abstract
In this thesis we study the ground state of ordered phases grown as thin layers on
substrates with smooth spatially varying Gaussian curvature. The Gaussian curvature
acts as a source for a one body potential of purely geometrical origin that controls the
equilibrium distribution of the defects in liquid crystal layers, thin lms of He
4
and
two dimensional crystals on a frozen curved surface. For superuids, all defects are
repelled (attracted) by regions of positive (negative) Gaussian curvature. For liquid
crystals, charges between 0 and 4 are attracted by regions of positive curvature while
all other charges are repelled. As the thickness of the liquid crystal lm increases,
transitions between two and three dimensional defect structures are triggered in the
ground state of the system. Thin spherical shells of nematic molecules with planar
anchoring possess four short
1
2
disclination lines but, as the thickness increases, a three
dimensional escaped conguration composed of two pairs of half-hedgehogs becomes
energetically favorable. Finally, we examine the static and dynamical properties that
distinguish two dimensional crystals constrained to lie on a curved substrate from
their at space counterparts. A generic mechanism of dislocation unbinding in the
presence of varying Gaussian curvature is presented. We explore how the geometric
potential aects the energetics and dynamics of dislocations and point defects such
as vacancies and interstitials.
iii
Contents
Title Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi
Dedication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii
1 Introduction 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Experimental motivation . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Mechanisms of coupling between defects and curvature . . . . 5
1.1.3 Cross-over between 2D and 3D physics . . . . . . . . . . . . . 13
1.1.4 Outline of the thesis . . . . . . . . . . . . . . . . . . . . . . . 14
2 Bond-orientational order on a corrugated substrate. 16
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Hexatic order on a surface . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2.1 Electrostatic analogy . . . . . . . . . . . . . . . . . . . . . . . 21
2.2.2 Defect free texture . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.3 Energetics of defect pairs on a Gaussian bump . . . . . . . . . 30
2.3 Curvature induced defect generation . . . . . . . . . . . . . . . . . . 35
2.3.1 Onset of the defect-dipole instability . . . . . . . . . . . . . . 35
2.3.2 Numerical investigation of defect-unbinding transitions . . . . 39
2.3.3 Single vortex instability . . . . . . . . . . . . . . . . . . . . . 44
2.3.4 Lattice of bumps, valleys and saddle points . . . . . . . . . . . 48
2.4 Defect Deconnement . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3 Liquid crystal textures in thick spherical shells. 60
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.2 Textures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.2.1 Tilted molecules on a sphere . . . . . . . . . . . . . . . . . . . 66
3.2.2 Nematic Texture . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.3 Stability of liquid crystal textures to thermal uctuations . . . . . . 80
iv
Contents v
3.4 Valence transitions in thick nematic shells . . . . . . . . . . . . . . . 88
3.4.1 Slab geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.4.2 Extrapolation to thin spherical shells . . . . . . . . . . . . . . 94
3.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4 Superuid lms on a curved surface. 99
5 Defects and crystalline order on curved surfaces. 112
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.2 Basic Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.3 Geometric Frustration . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.4 Geometric potential for dislocations . . . . . . . . . . . . . . . . . . 119
5.5 Dislocation unbinding and Grain Boundaries . . . . . . . . . . . . . 123
5.6 Glide suppression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
5.7 Vacancies, Interstitials and Impurities . . . . . . . . . . . . . . . . . 127
A Greens function and Isothermal Coordinates 131
B Geometric Potential 137
C Free energy on a corrugated plane 142
D Bump with a boundary 150
E Free energy of a vector eld on a sphere 162
F Liquid crystal textures and conformal mappings 170
G Vibrational spectrum of colloidal molecules 179
H Perturbation theory of curved crystals 188
I Curvature induced nite size eects 193
J Numerical Methods 199
Bibliography 203
Acknowledgments
This work originates from the inspiring mentorship of Professor D. R. Nelson
to whom goes my lasting gratitude for patiently guiding my rst steps in scientic
research while always encouraging independence of thought.
I had the privilege to have Professors B. I. Halperin and D. A. Weitz on
my thesis committee and I wish to thank them for enlightening interactions. As a
beginning graduate student, I attended some stimulating courses taught by Professors
S. Sachdev and D. S. Fisher that greatly fostered my interest in condensed matter
theory. One of the most rewarding aspects of my PhD experience at Harvard has been
the unique chance to informally interact with a number of very talented graduate
students and postdoctoral fellows from whom I learned a lot and shared moments of
fun: M. Desai, A. Delmaestro, A. Imambekov, I. Finkler, W. P. Wong, R. Barnett, D.
Podolsky, H. Chen, O. White, K. Papadodimas, E. Katifori, G. Rafael, A. Polkolnikov,
R. da Silveira, Y. Kafri, J. Qiang, C. Kilic and P. J. Lu. A special thanks goes to A.
Fernandez-Nieves, A. M. Turner and J. B. Lucks with whom I collaborated on the
work discussed in chapters I, IV and V respectively.
I wish to thank the administrative sta of the Physics Department in par-
ticular Dr. D. Norcross and S. Ferguson for providing constant help.
As an undergraduate at Imperial College London, I have been fortunate
to meet many inspiring lecturers that I wish to thank for their crucial help and
generosity particularly Professors M. P. Blencowe, J. P. Connerade, P. L. Knight, A.
McKinnon, M. B. Plenio and D. Vvedensky. An undergraduate bursary from the
Nueld foundation allowed me to try research rst hand and to spend a stimulating
vi
Acknowledgments vii
semester as a visiting student working under the supervision of Dr. T. S. Chang at
MIT.
I wish to thank my family and A. Alyahya for their support throughout.
To don Rocco, to my grandparents.
viii
Chapter 1
Introduction
1
Chapter 1: Introduction 2
1.1 Introduction
The physics of 2D condensed matter systems is a rich and mature subject
[1, 2]. In this thesis we study the eects induced in the ground state of a thin layer
of material by the Gaussian curvature of the substrate on which it is deposited. Spe-
cic examples include liquid crystalline order at the curved interface between two
immiscible uids, thin lms of He
4
wetting a corrugated substrate and two dimen-
sional crystals and liquid crystals on a lithographically prepared curved substrate.
While these systems are physically distinct, some of the consequences of the varying
Gaussian curvature of the substrate or interface can be understood within a unifying
perspective. It is the purpose of this introduction to illustrate basic concepts and
suggest possible experimental tests of the theoretical ideas presented in this thesis.
1.1.1 Experimental motivation
Recent experimental investigations of spherical crystallography have already
revealed complex ground state structures riddled with topological defects forced in by
the topology of the underlying curved space [3]. Fig. 1.1 shows a colloidosome engi-
neered in the Weitz lab with 0.5 m colloidal particles self assembled on an oil-water
interface. Besides its intriguing technological applications as a microcapsule for drug
delivery, the colloidosome provides a natural laboratory to explore (under controlled
experimental conditions) basic scientic questions of curved space crystallography rel-
evant to a larger class of materials including those of biological origin such as viruses
and clathrin-coated pits [5, 6, 7, 8, 9]. Confocal microscopy of small colloidosomes
Chapter 1: Introduction 3
Figure 1.1: Image of a colloidosome by confocal microscopy (courtesy of the Weitz
lab). The right panel shows the arrangement of 5 and 7-fold disclinations, repre-
sented as red and yellow dots respectively, in the ground state of a spherical crystal
(reproduced from Ref.[4]).
reveals 12 ve-fold coordinated lattice sites symmetrically placed approximately at
the vertices of an icosahedron inscribed in the sphere. When the droplet radius R
is larger than 5-6 times the colloidal radius, the nucleation of grain boundary scars
emanating from each disclination becomes energetically favorable [4, 10].
Liquid crystal (LC) analogs of these systems can be realized by trapping a
thin layer of thermotropic liquid crystals in a double emulsion [11]. Fig. 1.2 shows
four s =
1
2
disclination lines visible under crossed polarizers as double-brushes, which
confer to the colloidal particle an eective valence Z = 4 [12, 13]. The disclinations in
the droplet represented in Fig. 1.2 do not lie exactly at the vertices of a tetrahedron
because of the non uniform thickness of the layer. An alternative route to colloids
with a valence is provided by mono-layers of lyotropic nematogens [11]. The main
technological thrust behind these experiments is the desire to functionalize droplets
or colloids using chemical linkers. Such a tetravalent functionalization could be a
preliminary step towards the self assembly of a diamond lattice of colloids which has
Chapter 1: Introduction 4
Figure 1.2: The left panel shows an image of a double emulsion engineered (via
microuidics) by trapping a 8 m nematic layer between two spherical water in-
terfaces (picture courtesy of Alberto Fernandez-Nieves). The radius of the drop is
approximately 100 m. Under crossed polarizers each of the four disclinations with
topological charge s =
1
2
appears as a double brush. The right panel presents a
schematic illustration of the nematic texture in the mono-layer limit with the discli-
nations at the vertices of a tetrahedron. The core of the disclinations are indicated
as black dots to which chemical linkers (drawn as wiggly lines emanating from the
defects) can be attached by self assembly.
considerable potential for photonic band gap materials [13].
To characterize ground states in such systems, it is often convenient to switch
from a continuum description of the basic degrees of freedom (e.g. the angle that
indicates the local orientation of the LC molecules or the displacements of individual
atoms in the crystal) to an eective theory parameterized by the positions of a few
topological defects. As an example, the LC texture in Fig. 1.2 can be understood as
a result of the electrostatic-like repulsion between the four disclinations (see chapter
4 for more details).
Chapter 1: Introduction 5
2
FIG. 2: A surface with negative intrinsic curvature. The two principal curvatures are denoted by 1 and 2 and their product,
the intrinsic curvature, is negative (reproduced from Ref. [1]).
defect of charge q and the background curvature distribution is quadratic in q, hence both vortices and anti-vortices
are repelled from the top of the bump.
It is our hope to test the existence of this interaction in He lms by studying how it competes against the familiar
conning potential generated by the rotation of the bump. This geometric force exists in a variety of ordered phases
whose defects interact like a Coulomb gas and provides an unexpected coupling between matter and geometry.
II. BASIC THEORETICAL IDEAS
A defect at position r will feel a total potential energy, E(r), given by
E(r)

m
2
= V (r) +
A(r)

2
+ c , (3)
where m is the mass of
4
He, is the superuid density and c is an arbitrary constant. The length is dened as


m
. (4)
The rst contribution to E(r),

2

m
2
V (r), accounts for the interaction of the defect with the curvature (see Fig. 3). The
second contribution is

m
A(r), where A(r) is the area within a cup of radius r and it is multiplied by the number
density and the energy quantum . This last term generalizes (to curved space) the familiar parabolic potential in
at space that tries to conne the defect at the center of the bump as a result of the rotation (see Fig. 4). As one varies
a second order transition occurs. In fact, Fig. 5 reveals that for greater than a critical value
c
the total energy
E(r) assumes a mexican hat shape whose minima is oset with respect to the top of the bump. Taking a derivative
of Eq.(4) with respect to r leads to an implicit equation for the position of the minimum, r
m
, (or maximum):
r
m

= sin([r
m
]) . (5)
where (r) is the angle that the tangent at r to the bump forms with the xy plane in Fig. 1a. A simple graphical
construct allows to solve Eq.(5) by nding the intercept(s) of the sine curve on the RHS with the straight line of slope
Figure 1.3: A surface with negative Gaussian curvature. The two principal curvatures
(with dimensions of an inverse length) are denoted by
1
and
2
; their product is the
Gaussian curvature G(x) =
1

2
, which is negative in this example.
1.1.2 Mechanisms of coupling between defects and curvature
Defects in two dimensional monolayers can be modeled as point-like particles
that live in the local tangent plane to the surface and cannot escape in the third
dimension. The defects are sensitive to the Gaussian curvature of the substrate (see
Fig. 1.3), the latter being an intrinsic property of space that can be probed without
ever leaving the surface. Experiments such as those initiated in the Weitz lab can be
described fairly well by approximating the droplets as spheres of constant positive
Gaussian curvature. A signicant portion of this thesis addresses the less restrictive
case of condensed matter order on a substrate of varying Gaussian curvature. See
Figures 1.3 and 1.4 for examples.
The varying Gaussian curvature of the substrate acts as a source for a one
Chapter 1: Introduction 6
body potential V (x) of purely geometrical origin that controls the equilibrium dis-
tribution of the defects in the ground state and poses constraints on their dynamics.
These geometric interactions depend crucially on the scalar or vector topological
charge of the defect and on the symmetry of the order parameter. The defect poten-
tial is a non local function of the Gaussian curvature that can be explicitly determined
using the ideas and methods discussed in this thesis. To make the ideas more con-
crete, I shall compare the geometric potential felt by vortices in superuid
4
He on
curved surfaces to the one experienced by vacancies in crystalline solids grown on
a curved substrate. The former are topological defects that introduce long distance
disturbances in the superuid ow lines. The latter are point defects that arise from
locally subtracting only one atom from the curved space solid. Despite their obvious
dierences these two objects experience the same geometric potential V (x) that is
given (apart from multiplicative constants) by the solution of a Poisson-like equation
V (x) = G(x) (1.1)
where is the Laplacian and the Gaussian curvature G(x) plays the role of an elec-
trostatic like background charge. For the bump-like deformation shown in Fig. 1.4,
the qualitative form of V (x) in Eq.(1.1) can be guessed by appropriately generalizing
Gausss law of electrostatics. (An indented, dimple-like surface is equivalent within
our model to the one of Fig. 1.4 because the Gaussian curvature is the same).
Provided there is circular symmetry, the eld V (x) (proportional to the
force on a defect) can be determined as in electrostatics from integrating G(x) over
a circular patch centered around the top of the bump and with radius given by the
Chapter 1: Introduction 7
x
(b)
(a)
x
r
Figure 1.4: (a) A surface with a bump-like deformation. (b) Top view of (a) showing
a schematic representation of the positive and negative Gaussian curvature as a non-
uniform background charge distribution that switches sign at a characteristic radius
r = r
0
proportional to the size of the bump. The varying density of + and - signs is
intended to mimic the changing Gaussian curvature in the vicinity of the bump.
position of the defect. This integral is always positive and vanishes only at innity
because the Gaussian bump is topologically equivalent to the plane. Hence the defect
potential V (x) is a monotonically decreasing function of the radial distance and both
vacancies and vortices are repelled from the top of the bump where the curvature is
positive. These defects will not remain trapped in the regions where the curvature
is most negative, because the net force they experience results from the long range
(logarithmic) interaction with the Gaussian curvature everywhere on the surface. A
detailed derivation of these results along with heuristic arguments that make them
plausible is presented in the following chapters. Here we simply show how these
results t in our general conceptual scheme.
Although they share a common geometrical potential V (x), vacancies and
Chapter 1: Introduction 8
vortices are coupled to the Gaussian curvature by two distinct mechanisms that we
broadly classify as gauge and anomalous interactions respectively. The former results
from a direct coupling between the gradient of the displacement eld u
i
(x) and the
geometry of the substrate as described by gradients of its height function h(x). The
in-plane elastic free energy for a crystal embedded in a gently curved frozen substrate
can be expressed in terms of the at space Lame coecients and [14] as
F
c
=
_
dS
_
u
2
ij
+

2
u
2
kk
_
, (1.2)
where dS is the innitesimal area element. The strain tensor u
ij
given by
u
ij
(x) =
1
2
[
i
u
j
(x) +
j
u
i
(x) +A
ij
(x)] , (1.3)
contains an additional term A
ij
(x) =
i
h(x)
j
h(x) (compared to its at space coun-
terpart) that encodes informations on the geometry of the substrate. This rank two
tensor resembles the vector potential of electromagnetic theory; an appropriately de-
ned curl is equal to the Gaussian curvature [15]. A free energy like Eq.(2) has the
typical structure of the gauge theories often employed to describe condensed matter
order such as the London theory of a superconductor well below T
C
. In the super-
conductor analogy, the Gaussian curvature plays the role of a frozen and spatially
varying external magnetic eld.
The gauge coupling generates a force on each defect which is given by the
product of the eld V (x) and the charge of the
th
point defect

. The corre-
sponding defect strength

for an isotropic vacancy (interstitial) is given by the


negative (positive) area change caused by removing (adding) an atom at position x

.
Chapter 1: Introduction 9
Figure 1.5: An impurity in a crystalline matrix. The large shaded atom causes a local
dilation of the lattice. The square lattice has been chosen for simplicity, although the
ground state of a at two dimensional solid is typically an hexagonal lattice. By
contrast, a vacancy would correspond to removing the shaded atom from the lattice
leaving a local compression (reproduced from Ref.[16]).
The case of substitutional impurities can be treated phenomenologically by choosing

according to the characteristic size of the foreign atom introduced in the original
lattice (see Fig. 1.5);

will hence be allowed to vary continuously in our treatment


[16]. The coupling constant for this elastic interaction is controlled by the Youngs
modulus Y =
4(+)
2+
. As shown in Chapter 5, the resulting force on a vacancy (in-
terstitial) in the outward (inward) radial direction,

f
c
, has the familiar form of the
electrostatic interaction between a charge and an external eld,

f
c
=
Y
2

V (x) . (1.4)
The tensor A
ij
(x) accounts for another key feature of two dimensional crys-
tals on a curved substrate which is intimately related to the existence of a gauge
coupling: geometric frustration. In the presence of Gaussian curvature, the elastic
Chapter 1: Introduction 10
ground state will always be characterized by some strain u
G
ij
(x) and stress
G
ij
(x) which
result in a non-vanishing stretching energy, even in the absence of defects. The origin
of the long range geometric potential V (x) experienced by vacancies and interstitials
lies in the fact that the local compression or dilation (measured by

) that they
induce couples to the preexisting isostatic pressure of the stress
G
kk
(x

), which is a
non-local function of the Gaussian curvature. In fact, elastic deformations created
by the geometric constraint throughout the curved 2D solid are propagated to the
position of the point defect, x

, by force chains spanning the entire system. Vacan-


cies, interstitials or impurities atoms can all be viewed as local probes of the stress
eld that, as a rst approximation, are unaected the additional stresses induced by
their own presence (see Chapter 5). The additional self-energies which are not taken
into account by this approach are of order Y
2

G(x) and can be neglected as long as


the size of these point defects is much smaller than the radii of curvature of the
substrate.
By contrast, no geometric frustration exists for a
4
He lm on a corrugated
substrate because its wave function (x) = Ae
i(x)
is dened in a dierent space from
the one in which the superuid is conned. In the ground state of a
4
He lm, the
phase (x) is constant throughout the surface and the corresponding energy vanishes.
The free energy, F
s
, of a nonrotating superuid lm does not include a geometric
gauge eld. As discussed in Chapter 3, this free energy reads
F
s
=
K
2
_
dS g

(x)

(x) , (1.5)
where g

is the metric tensor of the substrate and the coupling constant K =


s
2
m
4
2
is
Chapter 1: Introduction 11
expressed in terms of the mass of an
4
He molecule m
4
and the superuid mass den-
sity per unit area
s
. When vortices are introduced into Eq.(4.2), these topological
charges behave somewhat like electrostatic charges when conned in a bounded ge-
ometry. Even in the absence of externally imposed supercurrents, position-dependent
self-energies originate from the broken translational symmetry implied by the presence
of the boundary or the varying Gaussian curvature. In an electrostatic analogy, each
charged particle induces polarization charges on the conducting walls (distributed
according to the shape of the boundary) with which it interacts [17]. On a curved
substrate, the induced topological charge depends both on the charge (quantum of
circulation) of the vortex and on the Gaussian curvature of the surface in which it is
imbedded. The resulting force, f
s
is given in terms of the solution of Eq.(1.1) by

f
s
= K s
2

V (x) , (1.6)
where the integer s measures the vorticity of the defect. Unlike the charge

that
describes the local area deformations in crystals, this topological charge of a vortex
or anti-vortex must be quantized. A comparison of Eq.(1.4) and Eq.(1.6) reveals that
the geometric forces experienced by vacancies (for which

< 0) and vortices or


anti-vortices in
4
He are indeed the same (apart from multiplicative constants). Note
however that the dependence on the charge of the point defect,

, is linear in
Eq.(1.4) whereas the force in Eq.(1.6) depends quadratically on the vortex charge
s. The mechanism of coupling between curvature and vortices in liquid helium is
qualitatively dierent from the gauge coupling discussed in the context of curved
space crystallography and in what follows we will refer to it as anomalous coupling.
Chapter 1: Introduction 12
Figure 1.6: Monte Carlo simulation of curvature induced unbinding of dislocations for
particles interacting with a Yukawa pair potential on two similar curved substrates
[18]. The dark and light particles in the dislocation core represent a disclination pair.
The two panels represent candidate ground states that have lower energies than the
respective congurations without defects. In the left panel, unbound dislocations (i.e.
disclinations pairs) are oppositely aligned in the radial direction towards the bump.
In the right panel, dislocation dipoles neutralize their Burgers vectors by having
their (opposite) dipole orientation alternate between being pointed towards bumps
and away from saddles. The rst scenario is favored for large bump separations.
Other defects such as disclinations in liquid crystals t in this classication with both
types of couplings playing an important role.
An interesting consequence of geometric frustration in curved crystals and
liquid crystals is the possibility of structural transitions between a defect-free ground
state to energetically favored congurations where defects nucleate to partially screen
the Gaussian curvature. Fig. 1.6 shows the result of minimizing the energy of point
particles interacting via a Yukawa potential and constrained to lie on a corrugated
geometry with two dierent values of the bump spacing relative to the bump size
[18]. The two candidate ground states characterized by a charge-neutral set of
dislocations have lower energies than a frustrated but defect-free hexagonal lattice.
Both the actual distribution of dislocations and the critical aspect ratio of the bumps
Chapter 1: Introduction 13
necessary to trigger the instability can be calculated from the geometric potential of
an isolated dislocation, which is a function of both its position and the orientation of
its Burger vector (see Chapter 5).
1.1.3 Cross-over between 2D and 3D physics
The theoretical framework described in the previous pages applies to very
thin lms of approximately constant thickness which can be treated as two dimen-
sional systems. It is interesting to study the crossover from the two to three di-
mensional regime that occurs as the thickness of the layer of material, h, increases.
Consider for concreteness the case of a nematic shell coating a solid colloidal parti-
cle on which the nematic molecules are aligned tangentially or a double emulsion
composed of a nematic liquid coating PVA enriched water droplets in a solution of
glycerol, PVA and water. Experimental investigations of this later system are cur-
rently underway in the Weitz lab [11].
In the mono-layer limit, the baseball like texture reproduced in Fig. 1.2 is
characterized by four s =
1
2
disclinations. However, for thicker shells three dimen-
sional defect congurations characterized by two pairs of half-hedgehogs at the inner
and outer surface compete with the planar texture discussed previously leading to
structural transitions above a critical value of h
c
. The escaped 3D texture is illus-
trated schematically in the right most panel of Fig. 1.7 where the pair of surface
half-hedgehogs at the south pole is highlighted by the small circle and the center of
each defect indicated by a dot. For samples between crossed polarizers, the optical
signature of this transition is a change from the quadrupolar to the bipolar defect
Chapter 1: Introduction 14
Figure 1.7: Quadrupolar and bipolar double-emulsions observed through crossed po-
larizers. In the left panel, the 4 disclinations are visible as 4 two-fold brushes as in
Fig. 1.2. Each of the two pairs of half-hegehogs is visible in the middle panel as
a 4-fold brush. Experimental conditions are similar to Fig. 1.2. The right panel
shows a schematic illustration of a candidate nematic texture escaped in the third
dimension that is consistent with the image of the bipolar droplet presented in the
middle panel. Thickness inhomogeneities in the nematic shell are believed to bring
these defect patterns into the same hemisphere, which allows easy visualization.
patterns illustrated in Fig. 1.7.
1.1.4 Outline of the thesis
This thesis is organized around the themes sketched in the previous sec-
tions. In Chapter 2 we present a detailed study of liquid crystal order on surfaces
of varying Gaussian curvature that relies on the geometric potential of an individual
disclination. The (exact) non-perturbative solution of the Poisson equation by means
of conformal mappings introduced in this context is an important mathematical in-
gredient of our approach. In Chapter 3, the crucial distinction between the gauge and
anomalous couplings to the Gaussian curvature is discussed by comparing the physics
of liquid crystal and superuid lms on a corrugated substrate. The mathematical
description of these two systems is very similar in the plane [1, 2], but this chapter
reveals remarkable dierences on a curved substrate. In Chapter 4 the ground state
Chapter 1: Introduction 15
of liquid crystal shells is explored including an analysis of the crossover between two
and three dimensional regimes. In Chapter 5, we conclude with a study of the curved
space crystallography of dislocations, vacancies and interstitials and impurity atoms.
Our approach starts from the derivations of the geometric potentials which act on
these defects and proceeds with a discussion of stress relaxation and defect unbind-
ing instabilities in two dimensional crystals on curved substrates. The dynamics of
dislocation motion on a curved surface is discussed as well.
Chapter 2
Bond-orientational order on a
corrugated substrate.
Based on V. Vitelli and D. R. Nelson PRE 70, 051105 (2004).
16
Chapter 2: Bond-orientational order on a corrugated substrate. 17
2.1 Introduction
The melting of a two dimensional crystal can occur continuously via two
second order topological phase transitions characterized by the successive unbinding
of dislocation and disclination pairs. At low temperatures, dislocations are suppressed
due to their large energy cost, but as the temperature is increased, the entropy gained
by creating defects overcomes their cost in elastic energy and dislocation unbinding
occurs to reduces the overall free energy of the system [19, 20, 21]. The quasi-long
range order of the crystal is thus destroyed leading to an hexatic phase that still
preserves quasi-long range orientational order. This phase can be characterized by
a complex order parameter with six-fold symmetry. As the temperature is increased
still further, an additional disclination-unbinding transition occurs and the hexatic
order is nally lost in an isotropic liquid phase [19].
Experimental evidence for hexatic order and defect-mediated melting has
been obtained in systems as diverse as free standing liquid crystal lms [22], Langmuir-
Blodgett surfactant monolayers [23], two-dimensional magnetic bubble arrays [24],
electrons trapped on the surface of liquid helium [25, 26, 27], two-dimensional colloidal
crystals [28, 29] and self-assembled block copolymers [30].
The unbinding of defects in the plane is entropically driven and at low tem-
perature defects are tightly bound. By contrast, on surfaces with non zero (integrated)
Gaussian curvature, excess defects must be present even at very low temperatures.
The theory of topological defects in ordered phases conned to frozen topographies
with positive or negative Gaussian curvature has been investigated previously; see,
Chapter 2: Bond-orientational order on a corrugated substrate. 18
e.g., [15, 31, 2]. As a general rule, regions of positive or negative curvature (valleys,
hills or saddles) lead to unpaired disclinations in the ground state, possibly screened
by clouds of dislocations. These clouds can in turn condense into grain boundaries
at low temperature. The predictions of recent studies of crystalline order on a sphere
[4] have been conrmed in elegant studies of colloidal particles packed on the surface
of water droplets in oil [3]. Investigations of the physics of defects in curved spaces
have also been carried out for uctuating geometries [32, 33, 34, 35]. The dynamics of
hexatic order on uctuating spherical interfaces was studied in Ref. [36]. Quenched
random topographies in the limit of small deviations from atness were investigated
in Ref. [15] .
In the present work, we investigate topography-driven generation of defects
on simple frozen surfaces with spatially varying Gaussian curvature whose topology
does not automatically enforce their presence in the ground state. We study in
particular a two-dimensional bump with a Gaussian shape and dimension large
compared to the particle spacing. For such a hilly landscape, at at innity, the
geometric control parameter is an aspect ratio given by the bump height divided by
its spatial extent. Consider a hexatic phase draped over such a bump. For small
bumps, the ideal hexatic texture is distorted, but there are no defects in the ground
state. As the aspect ratio is increased, we nd that disclination pairs progressively
unbind at T = 0 in a sequence of transitions occurring at critical values of the aspect
ratio. The defects subsequently position themselves to partially screen the Gaussian
curvature. For bumps embedded in surfaces of suciently small spatial extent, a
Chapter 2: Bond-orientational order on a corrugated substrate. 19
second instability of the smooth ground state needs to be considered. In this case,
the energy stored in the eld can be lowered by generating a single positive defect at
the center of the bump. Novel eects also arise when the hilly surface is encircled by
an aligning circular wall that insures a 2 rotation of the orientational order in the
ground state. In this case, some of the positive defects required to match the curvature
of the boundary are conned to a hemispherical cup centered on the bump, provided
the aspect ratio is larger than a critical value
D
. When is lowered below
D
,
the positive defects originally trapped in the hemispherical cup start undergoing
a series of sharp deconnement transitions, as they progressively migrate to new
equilibrium positions dictated by boundary conditions and the nite system size. We
also suggest possible ground states for periodic arrays of bumps, like those on the
bottom of an egg carton.
A natural arena to experimentally study the interplay between geometry
and defects is provided by thin copolymer lms on SiO
2
patterned substrates [37].
Flat space experiments by Segalman et al. have already demonstrated that spherical
domains in block copolymer lms form hexatic phases [30].
Our results for hexatics on frozen topographies also apply to other XY-
like models, as might be appropriate for tilted surface-active molecules on curved
substrates with interactions which favor alignment. The results are relevant as well to
two-fold nematic order on frozen topographies. In both cases, we expect qualitatively
similar defect unbinding transitions, although the equivalence becomes more exact
in the one Frank constant approximation [38]. Related results have been obtained
Chapter 2: Bond-orientational order on a corrugated substrate. 20
recently for order on a torus [39]. Even though the integrated Gaussian curvature
vanishes, defects appear in the ground state in the limit of fat torii, unless the number
of degrees of freedom is very large.
The outline of this paper is as follows. In Section II, the relevant mathemat-
ical formalism is introduced and used to highlight similarities and dierences between
defects on surfaces of varying curvature and electrostatic charges in a non-uniform
background charge distribution in at space. As an example, we calculate the dis-
torted, but defect-free, ground state texture of a hexatic conned to a surface shaped
as a Gaussian bump for aspect ratios below the rst disclination-unbinding instabil-
ity. In Section III, we investigate curvature-induced defect formation for an isolated
bump and a periodic array of bumps. In section IV, defect deconnement is discussed
and in section V various experimental issues related to our analysis are highlighted
along with some directions for future work. The development of the mathematical
formalism is largely relegated to Appendices. In Appendix A the Greens function
for the covariant Laplacian is derived by means of conformal transformations. In Ap-
pendix B, we introduce a geometric potential whose source is the Gaussian curvature.
In Appendix C, we present the general formula for the energy of textures with defects
in terms of the two functions derived in Appendix A and B. We thus explore the
existence of position-dependent defect self-interactions that arise from the varying
Gaussian curvature. Finally, boundary eects are discussed in Appendix D.
Chapter 2: Bond-orientational order on a corrugated substrate. 21
2.2 Hexatic order on a surface
2.2.1 Electrostatic analogy
The free energy for hexatic degrees of freedom embedded in an arbitrary
frozen surface can be written as
F =
K
A
2
_
dAD

(u)D

(u) , (2.1)
where u = {u
1
, u
2
} is a set of internal coordinates, n(u) is a unit vector in the tangent
plane, D

is the covariant derivative with respect to the metric of the surface and dA is
the innitesimal surface area [33, 32, 35, 40]. The generalization to systems with a p-
fold symmetry is straightforward provided that the one Frank constant approximation
is used and the consequences of the uniaxial coupling neglected [38]. This choice of
free energy implies that the minimal energy conguration will be given locally by
neighboring n(u) vectors which dier only by parallel transport. The curvature of the
surface induces frustration in the texture. In fact, by Gauss Theorema egregium
[41, 42], tangent vectors parallel transported along a closed loop are rotated by an
amount equal to the Gaussian curvature integrated over the enclosed area. On a
sphere, for example, the hexatic ground state always has twelve excess disclinations
as a result of this frustration [31, 12]. More generally, the sum of the topological
charges on any closed surface is equal to the integrated Gaussian curvature.
By introducing a local bond-angle eld (u), corresponding to the angle
between n(u) and an arbitrary local reference frame, we can rewrite the hexatic free
Chapter 2: Bond-orientational order on a corrugated substrate. 22
energy introduced in Eq.(3.1) as:
F =
1
2
K
A
_
dAg

)(

) , (2.2)
where dA = d
2
u

g, g is the determinant of the metric tensor g

and A

is the spin-
connection whose curl is the Gaussian curvature G(u) [40, 42]. The spin connection
can be viewed as a geometric vector potential. A free energy like Eq.(2) also
describes the charged Cooper pairs implicit in the London theory of a superconductor
well below T
C
. In the superconductor analogy, the Gaussian curvature plays the role
of a (spatially varying) external magnetic eld. For the problem considered here,
however, there are interesting new nonlinear eects associated with spatial variations
in the metric.
A detailed analysis of the free energy of Eq.(2.2) for a bumpy surface with
free and circular boundary conditions is presented in Appendices (C) and (D). Here
we only sketch the main steps and conclusions. The free energy can be readily con-
verted into a Coulomb gas model by using the relation

) = s(u) G(u) n(u) , (2.3)


where

is the covariant antisymmetric tensor, G(u) is the Gaussian curvature and


s(u)
1

N
d
i=1
q
i
(u u
i
) is the disclination density with N
d
defects of charge q
i
at
positions u
i
. The nal result is an eective free energy whose basic degrees of freedom
are the defects themselves [35, 4]:
F =
K
A
2
_
dA
_
dA

n(u) (u, u

) n(u

) , (2.4)
Chapter 2: Bond-orientational order on a corrugated substrate. 23
where n(u) is dened in Eq.(E.18). The Greens function (u, u

) is calculated (see
Appendix A) by inverting the Laplacian dened on the surface
(u, u

)
_
1

_
uu

, (2.5)
and we have suppressed for now defect core energy contributions which reect the
physics at microscopic length scales. Eq.(E.19) can be understood by analogy to two
dimensional electrostatics, with the Gaussian curvature G(u) (with sign reversed)
playing the role of a non-uniform background charge distribution and the topolog-
ical defects appearing as point-like sources with electrostatic charges equal to their
topological charge q
i
. As a result, the defects tend to position themselves so that
the Gaussian curvature is screened: the positive ones on peaks and valleys and the
negative ones on the saddles of the surface. However, this analogy does neglect
position-dependent self-interactions [43], but since these are quadratic in the charge
they are negligible for hexatics. Hence positive disclinations of minimal topological
charge q
i
=
2
6
continue to be attracted to positive curvature (see Appendix C).
More generally, we can consider p-fold symmetric order parameters with
minimum charge defects
2
p
. The case p = 1 corresponds to tilt order of absorbed
molecules and p = 2 describes 2D nematics. The cases p = 4 and p = 6 describe
tetradic and hexatic phases respectively [12]. Strictly speaking, Eq.(1) only describes
the cases p = 1 and p = 2 in the one-Frank-constant approximation [38]. Most of
our discussion focuses on topography-driven transitions on a model surface shaped
like a bell curve or Gaussian bump (see Fig. 2.1), but the same mathematical
approach can be readily carried over to study arbitrary surfaces of revolution that
Chapter 2: Bond-orientational order on a corrugated substrate. 24
Figure 2.1: (a) The vector eld n is conned to a surface shaped as a Gaussian.
(b) Top view of (a) showing a schematic representation of the positive and negative
Gaussian curvature as a background charge distribution that switches sign at r =
r
0
. Note that, according to the electrostatic analogy, a positive (negative) distribution
of Gaussian curvature corresponds to negative (positive) topological charge density.
are topologically equivalent to the plane. Furthermore, we do not expect the results
of this analysis to depend qualitatively on the azimuthal symmetry of the surface,
which is assumed purely for reasons of mathematical convenience.
Points on our model surface embedded in three dimensional Euclidean space
are specied by a three dimensional vector R(r, ) given by
R(r, ) =
_
_
_
_
_
_
_
r cos
r sin
hexp
_

r
2
2r
2
0
_
_
_
_
_
_
_
_
, (2.6)
where r and are plane polar coordinates in the xy plane of Fig. 2.1. It is useful
to characterize the deviation of the bump from a plane in terms of a dimensionless
Chapter 2: Bond-orientational order on a corrugated substrate. 25
0.5 1 1.5 2 2.5 3
1
2
3
4
5
6
7
l(r/r
0
)
r/r
0
Figure 2.2: Plot of l(
r
r
0
) as a function of the dimensionless radial coordinate
r
r
0
for
= 1, 2, 3, 4. The arrow is oriented in the direction of increasing .
aspect ratio

h
r
0
. (2.7)
The two orthogonal tangent vectors t
r

R
r
and t

can be normalized to dene


the Vierbein (orthonormal basis vectors) E
r
and E

respectively. The components of


the spin connection introduced in Eq.(2.2) are given by A

= E
r

[40, 42]. This


leads to a vanishing radial component A
r
and
A

=
1
_
l(r)
, (2.8)
where the important -dependent function l(r) (see Fig. 2.2) is dened by
l(r) 1 +

2
r
2
r
2
0
exp
_

r
2
r
2
0
_
, (2.9)
Chapter 2: Bond-orientational order on a corrugated substrate. 26
and it is equal to the radial component of the diagonal metric tensor, g

,
g

=
_
_
_
l(r) 0
0 r
2
_
_
_
. (2.10)
Note that the g

entry is equal to the at space result r


2
in polar coordinates while
g
rr
= l(r) is modied in a way that depends on but tends to the plane result g
rr
= 1
for small and large r, as illustrated in Fig. 2.2.
The Gaussian curvature for the bump is readily found from the eigenvalues
of the second fundamental form [44],
G

(r) =

2
e

r
2
r
2
0
r
2
0
l(r)
2
_
1
r
2
r
2
0
_
. (2.11)
Note that controls the order of magnitude of G(r) and that G(r) changes sign at
r = r
0
(see Fig. 2.1b). The integrated Gaussian curvature G(r) inside a cup of
radius r centered on the bump is
G(r) = 2
_
1
1
_
l(r)
_
, (2.12)
which vanishes as r . Eq.(2.12) also shows that the positive Gaussian curvature
enclosed within the radius r
0
(see Fig. 2.1) approaches 2 for 1, half the
integrated Gaussian curvature of a sphere.
2.2.2 Defect free texture
For small values of the aspect ratio , the minimal energy texture for the
hexatic will be free of defects. The ground state conguration
o
(u) satises the
dierential equation
D

0
D

= 0 , (2.13)
Chapter 2: Bond-orientational order on a corrugated substrate. 27
Figure 2.3: Projected ground state texture for an XY model on the bump, with the
boundary condition that the vector eld is parallel to the y-axis at innity. The two
insets show the defect pairs suggested by two regions of large frustration, which lie
close to a circle of radius r
0
.
which results from minimizing the free energy in Eq.(2.2) with respect to the eld (u)
for xed A

. When expressed in terms of the coordinates in Eq.(2.6), the solution of


Eq.(2.13) reads:

o
(u) = +c , (2.14)
where c is an arbitrary constant. The smooth ground state texture is thus obtained
if the director n forms an angle
o
(u) = +c with respect to the spatially varying
basis vector E
r
. Note that a solution of the form
o
(u) = c represents a defect of
charge q = 2 in this rotating system of coordinates.
As an illustration, consider the projection on the plane of the minimal en-
ergy texture of an XY model (p = 1) as shown in Fig. 2.3. The arrows represent
the orientation of tilted molecules on this surface in the one Frank constant approxi-
mation. The eld clearly displays strong frustration along a direction determined by
Chapter 2: Bond-orientational order on a corrugated substrate. 28
the choice of the constant c in Eq.(2.14). If the bump is positioned within two very
distant walls parallel to the y-axis which impose tangential boundary conditions on
the molecular tilts, the preferred direction will be along y
1
. The texture displayed
in Fig. 2.3 can be interpreted as resulting from embryonic pairs of defect dipoles
along the line x = 0. The distortion energy F
0
of this ground state is given by:
F
0
=
1
2
K
A
_
dAg

o
A

)(

o
A

) . (2.15)
This expression can be evaluated for an innitely large system by using Eq.(2.14) and
the explicit form of the spin connection derived in Section 2.2.1, with the result:
F
0
= K
A
_

0
dr
_
1
_
l(r)
_
2
r
_
l(r)
. (2.16)
It follows from Eq.(2.16) that the ground state energy is a monotonically increasing
function of the aspect ratio, proportional to
4
for small . As we shall see, for large
enough , it can be energetically preferable to reduce this energy by introducing
defect pairs into the texture. It is convenient to rewrite Eq.(2.15) in terms of the
Gaussian curvature G(r) and the Green function (u, u

) discussed in Appendix A
[40]
F
0
=
K
A
2
_
dA
_
dA

G(u) (u, u

) G(u

) . (2.17)
This result is what one obtains by setting all q
i
= 0 in Eq.(E.19). The details of the
mathematical derivation are relegated to Appendix C.
Although this result correctly represents the zero temperature limit of the
vector model, corrections may be appropriate to describe the physics of ordered phases
1
In case distant walls are present, the free solution of Eq.(2.14) is slightly modied to account for the
new boundary conditions. This is accomplished by the method of images or conformal transformations [45].
Chapter 2: Bond-orientational order on a corrugated substrate. 29
at nite temperature. Spin-wave excitations (i.e., quadratic uctuations of the
order parameter about the ground state texture) can be accounted for by integrating
out the longitudinal uctuations

(u) around the ground state conguration


o
(u).
By letting =
0
+

in Eq.(2.2) and using Eq.(E.18) we obtain [32, 35]:


F = F
0
+
1
2
K
A
_
dAg

, (2.18)
The longitudinal variable

(u) appears only quadratically in F and the trace over

(u) can be explicitly performed with the result [46]:


_
D

K
A
2
R
dA g

= e
F
L
, (2.19)
where F
L
is the Liouville action,
F
L
= c
_
dA
K
A
24
_
dA
_
dA

G(u) (u, u

) G(u

) . (2.20)
The rst term in this expression is a constant proportional to the xed surface area
of the frozen topography and will be suppressed in what follows. The remaining term
causes a shift in the coupling constant appearing in Eq.(2.17) from K
A
to K

A
=
K
A

k
B
T
12
[32, 35]. This entropic correction to the coupling constant K
A
at nite
temperature also arises when defects are present.
The energy in Eq.(2.17) represents an intrinsic, irreducible energy cost of
geometric frustration for textures without defects. As we shall see, defects can reduce
this frustration. However, for small values of the energy cost of this frustration
will still be lower than the core energies associated with the creation of the unbound
defects and the work necessary to tear them apart.
Chapter 2: Bond-orientational order on a corrugated substrate. 30
2.2.3 Energetics of defect pairs on a Gaussian bump
A quantitative understanding of the energetics of defects on a xed topog-
raphy is essential to calculate the critical value(s) of the aspect ratio above which
defect-unbinding becomes energetically favorable. The rst step is to calculate the
Greens function (u, u

) that governs the coulombic interaction among defects and


between each defect and the Gaussian curvature. The inversion of the curved space
Laplacian can be more easily accomplished by employing a set of isothermal coor-
dinates, such that the resulting Greens function reduces to the familiar logarithm
of two dimensional electrostatics. As shown in Appendix A the nal result in terms
of the original polar coordinates reads:
(u, u

) =
1
4
ln['(r)
2
+'(r

)
2
2'(r)'(r

) cos(

)] +c , (2.21)
where the function '(r) can be thought of as a radial coordinate in the conformal
plane resulting from adopting an isothermal set of coordinates (see Appendix A)
'(r) = r e

r
dr

l(r

)1

, (2.22)
and l(r) is the -dependent function introduced in Eq.(2.9). The constant c depends
on the physics at short distances, which is discussed in Appendix A.
The Green function in Eq.(2.21) corresponds to free boundary conditions
at innity and preserves the cylindrical symmetry of the metric. It diers from
the familiar result in at space by a non-linear radial stretch corresponding to a
smooth deformation of the bump into a at disk. This Greens function determines an
Chapter 2: Bond-orientational order on a corrugated substrate. 31
attractive interaction for the defect dipole pair. However, the Gaussian curvature of
the bump also generates a geometric potential that tries to pull the disclination dipole
apart. This geometric interaction arises by combining cross terms between s(u) and
G(u) in Eq.(E.19) with the position dependent self-interactions derived in Appendix
C. The resulting interaction F
G
between defects and the Gaussian curvature takes
the simple form
F
G
= K
A
N
d

i=1
q
i
_
1
q
i
4
_
V (u
i
) , (2.23)
where the geometrical potential V (u) is dened as
V (u)
_
dA G(u

) (u, u

) . (2.24)
The minus sign in front of this geometric potential insures that defects of topological
charge between zero and 4 are attracted by regions of positive Gaussian curvature
[43]. For defects with a large topological charge of q > 4, the sign of the geometric
interaction, F
G
, is reversed and, and defects of either sign are pushed away from the
bump. This scenario does not aect the geometry-driven defect formation discussed
in this paper that relies on disclinations whose charge is well below 4. However, as
a result of the position-dependent self interactions, F
G
is no longer symmetric under
the change q q, as one would expect on the basis of the electrostatic analogy. The
eect of this asymmetry is small for hexatic order but it increase for liquid crystals
with p-fold order parameter, as p decreases (see Ref. [43], and references therein).
Gauss law generalized to curved surfaces (see Appendix B) insures that the
geometric force experienced by a defect of charge q with radial coordinate r will be
determined only by the net curvature enclosed in a circle of radius r centered on the
Chapter 2: Bond-orientational order on a corrugated substrate. 32
top of the bump. The resulting electric eld is radial as expected from electrostatics
and is proportional to the gradient of the geometric potential, which for a Gaussian
bump takes the form (see Appendix B)
V (r) =
_

r
dr

_
_
l(r

) 1
_
. (2.25)
For small values of the aspect ratio the potential in Eq.(2.25) can be approximated
by
V (r)

2
e

r
2
r
2
0
4
. (2.26)
The resulting force is linear for small r (i.e. near the top of the bump) and decays like
e

r
2
r
2
0
for r r
0
. As the aspect ratio increases, the force generated by the curvature
can overcome the attractive force binding the defect pair which varies logarithmically
for short distances. As a result, oppositely charged defects that were originally tightly
bound can be separated.
This argument, however, neglects another complication resulting from the
curvature of the surface: as the aspect ratio is increased, the Greens function (u, u

)
in Eq.(2.21) and hence the force binding the defects together also increases. We
illustrate this point in Fig. 2.4 for the special case of a positive defect pinned right
on top of the bump and a negative one free to move downhill at r. Upon invoking
Equations (E.19) and (2.21), the potential binding the pair, V
pair
(r), can be written
down exactly as:
V
pair
(r) =
K
A
q
2
2
ln
_
r
a
_
+ 2q
2
E
c

K
A
q
2
2
_

r
dr

_
_
l(r

) 1
_
. (2.27)
Chapter 2: Bond-orientational order on a corrugated substrate. 33
r
R
z
(r)
-
+
+
-
dr
Figure 2.4: Eect of changing the aspect ratio of the bump on the work needed to pull
apart two oppositely charged defects. Positive and negative defects are represented by
open circles with their sign printed. The line elements corresponding to the projected
length dr for the two aspect ratios are shown.
The rst term is the at space Greens function and we have added two disclination
core energies. The last term represents a curvature correction that has the same
functional form of the geometric potential in Eq.(2.25), but it represents a distinct
contribution to the total free energy. As discussed in Appendix B, the pair potential
energy, V
pair
(r), can be understood by applying a generalized Gauss Law to the bump
to determine a force which is equal to
q
2
2r
. The potential follows by integrating
this force along the bump with the length element dr
_
l(r). Although the force is
independent of the aspect ratio, the length element grows with (see Fig. 2.4), which
makes the pair more energetically bound for larger values of . A careful calculation
of these eects (including the contribution associated with the position-dependent
self-energies) reveals that the geometric force still overcomes the binding interaction
Chapter 2: Bond-orientational order on a corrugated substrate. 34
for suciently large values of .
An estimate of the critical value of for which the dipole unbinds can
be obtained by comparing the minimal free energy of the smooth frustrated eld
arising from Eq.(2.17) with the free energy in the presence of defects. The latter, as
follows from Eq.(E.19), is composed of three contributions: the interactions among
the defects, the interaction between the defects and curvature as given by Eq.(2.23)
and the Gaussian curvature self-interaction. The latter is equal to the minimal free
energy of the smooth frustrated eld and is renormalized at nite temperature in
the same way. (Associated with the cuto is a microscopic core energy, E
c
, that we
expect to be independent of the defect position on the bump, as long as the radius
of curvature is much greater than any microscopic length scale.) The remaining two
contributions are not renormalized by thermally induced spin wave uctuations [35].
As shown in Appendix C, the dierence in free energy of a defected texture described
by Eq.(E.19) relative to the defect-free result Eq.(2.17) can be written as:
F()
K
A
=
1
2
N
d

i=1
N
d

j=i
q
i
q
j

a
(r
i
,
i
, r
j
,
j
)
+
N
d

i=1
q
i
_
1
q
i
4
_
V (r
i
) +
E
c
K
A
N
d

i=1
q
2
i
. (2.28)
where we have assumed overall charge neutrality for the defect conguration. The
subscript in
a
indicates that a constant microscopic core radius a has been absorbed
in the denition of the Green function so that the argument of the logarithm in
Eq.(2.21) becomes dimensionless, as in Eq.(C.21),

a
(r
i
,
i
, r
j
,
j
) = (r
i
,
i
, r
j
,
j
) +
1
2
ln a . (2.29)
Chapter 2: Bond-orientational order on a corrugated substrate. 35
This microscopic cuto a corresponds to a constant core radius for each defect and
it is of the order of the spacing between the microscopic degrees of freedom. The
sum of microscopic core energies in the fourth term of Eq.(2.28) needs to be xed
phenomenologically or from models that go beyond simple elasticity theory [47]. Note
that both
a
(r
i
,
i
, r
j
,
j
) and V (r) depend on . If becomes suciently large so
that F is less than zero, one or more disclination dipoles unbind in the hexatic phase
at a sequence of critical values
c
i
. The analogous defects in XY-model textures of
tilted liquid crystal molecules would be +/ vortex pairs.
2.3 Curvature induced defect generation
2.3.1 Onset of the defect-dipole instability
If a dipole is created, say, along the line r = r
0
of zero Gaussian curvature,
the positive disclination will be pulled towards the center by the positive curvature
while the negative one will be repelled into the region of negative curvature. The net
result is a reduction of the total free energy of the order of the depth of the potential
well since the logarithmic binding energy is approximately constant compared to the
geometric potential. An approximate analytical treatment is obtained by assuming
that the positive defect sits right at the center of the bump and the negative one at a
distance of the order of r
0
. The validity of this approximation scheme can be checked
by numerically minimizing the energy with respect to the position of the defects, as
discussed in Section (2.3.2). We assume charge neutrality so that the two defects
have equal and opposite topological charges of magnitude q. For order parameters
Chapter 2: Bond-orientational order on a corrugated substrate. 36
0.25 0.5 0.75 1 1.25 1.5 1.75 2
3.5
3
2.5
2
1.5
1
0.5
0
V(r/r0 )
r/r0
-


-
-
-
-
-
-
Figure 2.5: Geometric potential V (r/r
0
) as a function of the dimensionless ratio of r
and r
0
for = 1, 2, 3, 4, 5. Note that | V (r) || V (0) | for r r
0
. The arrow points
in the direction of increasing .
with a p-fold symmetry, the minimal topological charge is q =
2
p
. The approximate
free energy cost to generate this defect pair then follows from Equations (2.25),(2.27)
and (2.28):
F()
K
A

q
2
2
_
ln
_
r
a
_
+V (r)
_
+q
_
1
q
4
_
V (0)
q
_
1 +
q
4
_
V (r) + 2q
2
E
c
K
A
. (2.30)
The internal consistency of the formalism can be checked by investigating the limit
r a. As the negative defect approaches the positive one at the center of the bump,
we have V (a) V (0) and the energy tends to the expected at space result 2q
2
E
c
.
The equilibrium position of the negative defect turns out to be for r equal
to a few r
0
, so the terms containing V (r) in Eq.(2.30) can be dropped as a rst
approximation because V (r) decays exponentially (see Fig. 2.5):
F()
K
A

q
2
2
ln
_
r
a
_
+q
_
1
q
4
_
V (0) + 2q
2
E
c
K
A
. (2.31)
Chapter 2: Bond-orientational order on a corrugated substrate. 37
1 2 3 4 5
6
5
4
3
2
1
0
V(0)

-
-
-
-
-
-
Figure 2.6: Plot of the geometric potential evaluated at the center of the bump, V (0),
for aspect ratios between 0 and 5. The continuous line is plotted using the exact
form of V (r) while the dashed line is obtained from the low expansion given in
Eq.(2.26).
To estimate the critical value of the aspect ratio for which the rst dipole unbinds,
let r r
0
and solve for
c
in Eq.(2.31). Because a dierent choice for the core energy
E
c
can be accounted for by rescaling the core size a in Eq.(2.31), the condition for
unbinding is
|V (0)| >
2q
(4 q)
ln
_
r
0
a

_
, (2.32)
with
a

= ae

4Ec
K
A
. (2.33)
If we know
Ec
K
A
and
r
0
a
, the critical aspect ratio
c
can be obtained from inspection
of Fig. 2.6 where V (0) is plotted as a function of . The Taylor expansion of V (r)
derived in Eq.(2.26) gives V (0)

2
4
to leading order in . Inspection of Fig. 2.6
Chapter 2: Bond-orientational order on a corrugated substrate. 38
shows that this approximation works suciently well even for aspect ratios of order
unity. Upon substituting for V (0) into Eq.(2.32), we obtain an estimate of how
c
depends on
r
0
a

2
c

8q
(4 q)
ln
_
r
0
a

8q
(4 q)
_
ln
_
r
0
a
_
+
4E
c
K
A
_
. (2.34)
Note that a critical height h
c
=
c
r
0
for defect unbinding is predicted for xed
r
0
a

with defect charge q =


2
p
for all integer values of p. The validity of this approximate
relation is tested in Section 2.3.2.
The continuum theory adopted here is valid in the limit r
0
a. If E
c
can
be neglected compared to K
A
, the defect unbinding instability is triggered when the
energy gain derived from letting the defects screen the Gaussian curvature (approx-
imately given by qV (0)) overcomes the work needed to pull them apart a distance
r
0
. This work, of order
q
2
2
ln
_
r
0
a
_
, increases very slowly with large
r
0
a
, hence the con-
tinuum approximation can be satised while keeping the work nite. Note that the
result does not depend on the size of the system R because we assume overall discli-
nation charge neutrality and the assumption that R r
0
. In this limit, boundary
eects can be ignored provided that they do not impose a topological constraint on
the phase of the order parameter. An aligning outer wall in a circular hexatic sample,
for example, would force the bond angle eld to rotate by 2, leading to six defects in
the ground state even in at space. The interesting physics which results is addressed
in Section 2.4.
Chapter 2: Bond-orientational order on a corrugated substrate. 39
2.3.2 Numerical investigation of defect-unbinding transitions
The disclination unbinding transitions can be investigated more quantita-
tively by minimizing numerically F() in Eq.(2.28) with respect to the positions of
the defects. The aspect ratio above which F() becomes negative corresponds to
the threshold value
c
(analogous to a rst order transition) for which the singular
eld is energetically favored with respect to the smooth texture of Fig. 2.3. We
emphasize that the energy landscape can have two minima. The rst occurs when
two oppositely charged defects form a closely bound dipole (with separation of the
order of the cuto a) and hence annihilate each other leaving a smooth texture. The
second minimum corresponds to an unbound pair (with separation of a few r
0
) and
it disappears when the geometric force is too weak to overcome the binding force of
the pair. This scenario occurs for a finite value of characteristic of the geometry
of the substrate above which the formation of an unbound dipole is possible, albeit
energetically unfavored (see Fig. 2.7). As is increased above
c
, the smooth-texture
minimum becomes metastable and the unbinding of a defect pair is the most likely
scenario.
It is useful to parameterize F() in terms of the dimensionless radial co-
ordinates r
i

r
i
r
0
. The geometric potential is dened in Eq.(A.9) as a function of r
i
,
V (r)

V

(
r
r
0
), where

(x) =
_

x
dy
y
_
_
1 +
2
y
2
exp(y
2
) 1
_
. (2.35)
In order to write the defect-defect interaction in terms of the dimensionless radial
Chapter 2: Bond-orientational order on a corrugated substrate. 40
0.5
1
1.5
2
2.5
0.1
0.2
- 0.3
- 0.25
- 0.2
- 0.15
0.5
1
1.5
2
Figure 2.7: Plot of F/K
A
versus x
1
and x
2
the positions of the negative and pos-
itive defects respectively in units of r
0
, for = 1.2. A constant energy oset equal
to
q
2
2
ln
_
r
0
a

_
has been neglected. The metastable minimum at x
1
1.3 r
0
and
x
2
0.2 r
0
corresponds to an unbound pair (see Fig. 2.8a). As is decreased
further the energy barrier that separates this minimum from the smooth texture so-
lution (corresponding to x
1
approaching x
2
) disappears and the two opposite defects
annihilate.
coordinate r
i
, we introduce a new function

'( r
i
) dened by

'( r
i
)
r
i
r
0
exp[V (r
i
)] =
r
i
r
0
exp
_

_
r
i
r
0
__
=
'(r
i
)
r
0
, (2.36)
where Eq.(A.8) was used in the last step. We can now transform
a
( r
i
,
i
, r
j
,
j
) by
eliminating '(r
i
) in favor of

'( r
i
). Thus we have using Equations (C.21) and (A.12)

N
d

j=i
q
i
q
j

a
(r
i
,
i
, r
j
,
j
) =
N
d

j=i
q
i
q
j
( r
i
,
i
, r
j
,
j
)
+
1
2
N
d

i=1
q
2
i
ln(
r
0
a
) , (2.37)
where we have exploited charge neutrality and Eq.(C.22). The free energy minimized
Chapter 2: Bond-orientational order on a corrugated substrate. 41
with respect to the positions of the defects, min
_
F
K
A
_
, can now be written, according
to Eq.(2.28), as
min
_
F
K
A
_
= f() +
1
4
N
d

i=1
q
2
i
ln
_
r
0
a

_
, (2.38)
where
f() min [
1
2
N
d

j=i
q
i
q
j
( r
i
,
i
, r
j
,
j
)
+
N
d

i=1
q
i
_
1
q
i
4
_
V ( r
i
)] . (2.39)
Note that in the second term in Eq.(2.38) the core energy of each defect has been
absorbed in the modied core radius a

, as dened in Eq.(2.33). This generates an


energy cost for unbinding that can be overcome if f() assumes suciently large
negative values. Hence it is sucient to study numerically how F() varies as a
function of a single parameter, eg.
r
0
a

. As an illustration of this approach, we study


explicitly the unbinding of one and two disclinations pairs leading to the ground states
represented schematically in Fig. 2.8. The smooth ground state becomes unstable
to the formation of one defect dipole rst. The critical aspect ratio above which this
scenario occurs can be determined with the aid of Fig. 2.9 where the function f()
introduced in Eq.(2.39) is plotted as a function of the aspect ratio. From Eq.(2.38)
we see that
c1
is determined by
f(
c
1
) =
q
2
2
ln
_
r
0
a

_
. (2.40)
For
r
0
a

= 10
4
and q =
2
6
, we obtain a critical aspect ratio
c
1
3.2. As a comparison,
the approximate condition derived in Eq.(2.32) gives
c
1
3 when used in conjunction
Chapter 2: Bond-orientational order on a corrugated substrate. 42
r
0
r
0


(a)
(b)
+
+ +
-
- -

x
Figure 2.8: The equilibrium defects positions are illustrated schematically in the case
of one (a) and two dipoles (b). We assume free boundary conditions at innity, as in
Fig. 2.3, so that the eect of image charges can be neglected.
with Fig. 2.6. The rougher estimate in Eq.(2.34) leads (for q =
2
6
) to
c
1
2.6. This
discrepancy is easily understood considering that Eq.(2.34) was derived by means of
a low expansion. The critical aspect ratio
c
1
is too low for the two-dipole defect
conguration to become energetically favorable with respect to the smooth ground
state. Indeed, inspection of Fig. 2.10 reveals that the critical aspect ratio
c
2
for
which the two-dipole instability sets in is approximately equal to 3.6 for the same
choice of parameters used in the single pair case. Note that, in the presence of two
dipoles, the energy cost arising from the second term in Eq.(2.38) is twice as large
because there are four defects rather than two. However, for 4.2 generating
two dipoles becomes more energetically favored than a single dipole (see Fig. 2.11).
The approach illustrated here can be used to calculate a cascade of defect unbinding
Chapter 2: Bond-orientational order on a corrugated substrate. 43
2 3 4 5
4
3
2
- 1
0



f(!)
!
-
-
-
!c1
Figure 2.9: Plot of f() versus the aspect ratio obtained by minimizing over the
single-dipole defect conguration represented schematically in Fig. 2.8a. As discussed
in the text, the rst unbinding transition occurs when f(
c
1
) is equal to
q
2
2
ln
_
r
0
a

_
.
The value
c1
is indicated by the dashed line for
r
0
a

= 10
4
and q =
2
6
. Note that no
minimum (corresponding to an unbound pair) exists for less than 1 (approximately),
hence the curve cannot be continued to the origin (see Fig 2.7).
instabilities at critical aspect ratios
c
i
involving higher number of dipoles and their
equilibrium congurations in the ground state. Note that the unbinding eventually
stops since the integrated Gaussian curvature in the top cup cannot exceed 2. We
expect the qualitative features of our analysis to be independent of the exact shape
of the bumpy substrate although the specic values
c
i
depend on the geometry and
the choice of the microscopic parameters a and E
c
. Finally, we emphasize that the
curvature-induced unbinding is similar to a rst order transition and occurs for rather
pronounced deviations from atness (ie. large ).
Chapter 2: Bond-orientational order on a corrugated substrate. 44
2 3 4 5
7
6
5
4
3
- 2
- 1
0
f(!)
!
-
-
-
-
-
!c2
Figure 2.10: Plot of f() versus the aspect ratio obtained by minimizing the two-
dipole defect conguration represented schematically in Fig. 2.8b . The second
unbinding transition occurs when f(
c
2
) is equal to
q
2

ln
_
r
0
a

_
. The value
c
2
is
indicated by the dashed line for
r
0
a

= 10
4
and q =
2
6
(compare with Fig. 2.9).
2.3.3 Single vortex instability
The unbinding of defect pairs may not be the most likely scenario if the size
of the system R is suciently small. In this case, the creation of a single vortex at
the center of the bump may become energetically favorable for lower aspect ratios
than required by the defect dipole instability. The equation for the bond angle eld

s
(u) for a single defect of charge q at the center of the bump is given by:

s
() =
_
q
2
1
_
, (2.41)
where the bond angle is measured with respect to the rotating basis vectors corre-
sponding to the polar coordinates discussed in Section 2.2.2. Upon substituting
s
()
in Eq.(2.2) and subtracting the free energy F
0
corresponding to the defect-free texture
Chapter 2: Bond-orientational order on a corrugated substrate. 45
2 2.5 3 3.5 4 4.5 5 5.5
- 3
- 2
- 1
0
1
2
!F(")
" "c1 "c2
Figure 2.11: Plot of
F()
K
A
versus corresponding to a single dipole (continuous line)
and two dipoles (dotted line) for
r
0
a

= 10
4
. The critical aspect ratios
c
1
and
c
2
are indicated by dashed lines. Note that the aspect ratio for which the two dipole
conguration becomes energetically favored occurs for > 4.2.
we obtain:
F()
K
A
=
q
2
4
ln
_
'(R)
a
_
+q
_
1
q
4
_
V (0)
+ q
2
E
c
K
A
, (2.42)
where E
c
was added by hand. The same result is obtained by using the more general
formalism developed in Appendix D. Indeed, by letting the position of an isolated
defect tend to the center of the bump in Eq.(D.24) we obtain the energy of the singular
eld in the case of free boundary conditions and the result matches Eq.(2.42). As
discussed in Appendix D, a defect located at r
i
is attracted to the boundary at R
for free boundary conditions. One can think of this interaction as resulting from an
image defect of opposite sign behind the edge of the sample at position r

i
such that
Chapter 2: Bond-orientational order on a corrugated substrate. 46
the following relation holds in terms of the conformal radius '(r

)
'(r

i
) =
'(R)
2
'(r
i
)
. (2.43)
This result can be understood by analogy to the familiar electrostatic problem of a
charged line located a distance r
i
from the center of a cylindrical grounded conductor
whose axis is parallel to it [48]. The analogy becomes precise if one lets r
i
'(r
i
)
as explained in Appendix D.
If the geometric potential is not strong enough (as in the at space limit
= 0), the defect will migrate to the edge of the sample and annihilate with its
image leaving a smooth eld. On the other hand, when the aspect ratio is suciently
large, the defect can lower its energy by sitting at the center of the bump. Comparison
of Eq.(2.42) with Eq.(2.31) shows that, unless R r
0
, the energy of the single vortex
instability will be lower or at least comparable to the unbinding of a defect dipole.
In fact, the threshold
s
that needs to exceed to trigger the single defect instability
is easily obtained if the values of the geometric potential at the origin are tabulated
for dierent aspect ratios, as illustrated in Fig. 2.5. The condition for single vortex
generation reads
|V (0)| >
q
(q 4)
ln
_
R
a

_
. (2.44)
Using the same method adopted to derive Eq.(2.34) we obtain an estimate of how
s
depends on
R
a

(compare with Eq.(2.34)):

2
s

4q
(4 q)
ln
_
R
a

_
. (2.45)
Chapter 2: Bond-orientational order on a corrugated substrate. 47
The single vortex instability is reminiscent of vortex generation in rotating superuid
helium with playing the role of the angular speed . For a volume of helium
contained in a cylindrical vessel of radius R and rotating uniformly with constant
angular speed, the critical value
c
1
above which defect generation occurs is given by
[49]

c
1

K
2R
2
ln
_
R
a
_
, (2.46)
where K =
2
m
He
is the magnitude of the quantum of circulation and a the core radius
2
. Note that
c
1
decreases as R increases, unlike
s
which diverges logarithmically.
Thus, the single defect instability studied here is a nite size eect. In contrast, the
disclination unbinding studied earlier in this section does not depend on the system
size because of charge neutrality. Hence the thermodynamic limit can be safely taken,
provided the characteristic length over which the curvature varies (ie. r
0
) is not too
large compared to a (see Eq.(2.34)).
In considering the case of small system size, it is important to keep in mind
two assumptions implicit in the present treatment. The radius of curvature
r
0

must
be much larger than the core radius everywhere for the continuum approach to be
valid, that is r
0
a. Additionally, the Gaussian curvature must be vanishing small
at the edge of the system which requires R to be larger than a few r
0
.
2
A similar mechanism applies to superconductors in a uniform magnetic eld.
Chapter 2: Bond-orientational order on a corrugated substrate. 48
2.3.4 Lattice of bumps, valleys and saddle points
In some experimental realizations perhaps modelled on those of Ref. [30]
the topography will be periodic. In this Section we discuss qualitatively how the
results described above generalize to a 2-dimensional lattice of bumps with variable
aspect ratio for both square and triangular lattices. A more quantitative approach to
this problem would involve nding conformal set of coordinates for periodic boundary
conditions. This is possible in principle but more involved since cylindrical symmetry
is now lost. Nonetheless, the intuition gained by studying the single bump allows us
to make some guesses for the ground state. We rst note that the geometric potential
generated by the lattice of bumps is not simply the superposition of results for single
bump potentials. This is caused by the non linear relation between the surface height
and the Gaussian curvature acting as a source for the geometric potential. To explore
this point further, consider what happens when four bumps are placed at the vertices
of a square. At the center of the square a minimum of the height function occurs
corresponding to a new region of positive Gaussian curvature. This eect is particu-
larly acute for r
0
L, where L is the bump spacing. In general interference between
bumps creates a dual lattice of valleys. A similar breakdown of the superposition
principle arises for triangular lattices.
As the aspect ratio of hilly landscapes such as those shown in Fig. 2.12 is
increased, defects can be created to screen the Gaussian curvature. Their positions
can be guessed by considering a unit cell of the lattice such that the integrated Gaus-
sian curvature vanishes. For a square lattice, we conjecture that the rst topography
Chapter 2: Bond-orientational order on a corrugated substrate. 49
-6 -4 -2 0 2 4 6
-6
-4
-2
0
2
4
6
-6 -4 -2 0 2 4 6
-6
-4
-2
0
2
4
6
-5
0
5
-5
0
5
-1
0
1
2
3
-5
0
5
-3 -2.5 -2 -1.5 -1 -0.5 0
0
0.5
1
1.5
2
2.5
3
-3 -2.5 -2 -1.5 -1 -0.5 0
0
0.5
1
1.5
2
2.5
3
-3
-2
-1
0
0
1
2
3
0
0.25
0.5
0.75
1
-3
-2
-1
Figure 2.12: (Top) Ground states for square (left) and triangular (right) arrays of
bumps. The rst and second rows correspond to moderate values of the aspect ratio
respectively. For simplicity, we assume that r
0
, the bump width is comparable to
the lattice spacing. Positive defects (red dots) screen regions of positive Gaussian
curvature while negative ones (blue dots) are located on the saddles of the hilly
landscape.
Chapter 2: Bond-orientational order on a corrugated substrate. 50
induced transition is associated with the appearance of positive defects at the top
of the bumps and negative ones half way between them in the vertical or horizontal
direction (see Fig. 2.12). This two-fold degeneracy is compatible with the symme-
try of the lattice and analogous to the freedom in choosing the axis along which the
rst disclination-dipole appears on the single bump. The negative defects are shared
between two adjacent cells while the positive ones are shared among four cells thus
ensuring overall charge neutrality. As the value of increases even more, one might
expect an additional positive defect appears in the valley located at the center of
each cell and two additional negative defects shared with the adjacent cells are cre-
ated between the bumps at right angles to the direction discussed above (see Fig.
2.12).
For the triangular lattice, we conjecture that the rst transition corresponds
to positive defects on top of the bumps and negative ones between the bumps along
one of the three axis of symmetry of the unit cell. As the value of the aspect ratio
is increased, additional positive defects appear on the six minima of the surface and
negative ones are generated along the remaining two axes of symmetry of the unit
cell (see Fig. 2.12). A simple count of the total defect charges enclosed in the unit
cell shows that this scenario also satises the requirement of defect charge neutrality.
2.4 Defect Deconnement
With potential experiments in mind [37], it is interesting to consider the
case of hexatic order on a bump encircled by a circular wall of radius R r
0
which
Chapter 2: Bond-orientational order on a corrugated substrate. 51
aligns the hexatic bond angles. As a simple model, imagine an array of hexagons
which locally achieve a common orientation tangential to the wall (see Fig. D.1b).
The hexatic order parameter will thus rotate by 2 upon making a circuit of the
wall, insuring that at least six defects of charge
2
6
must be included in the ground
state for all values of the aspect ratio. These boundary-condition induced defects will
interact with the Gaussian curvature of the bump and with the wall. The N
d
defects
contribute large (constant) self energies of the form

N
d
i=1
K
A
q
2
i
ln
_
R
a
_
that dominate
the total energy for suciently large systems. Since

N
d
i=1
q
i
must be equal to 2, the
energy is minimized when the defects split up into the smallest possible charges.
The equilibrium defect conguration must minimize the free energy taking
into account the conning potential generated by the Gaussian curvature and the
interactions of the defects with the boundary and among themselves. The repulsive
force exercised by the wall on a defect located at '(r
i
) in the conformal plane can
be computed by placing an image defect of same charge outside the wall at position
(R)
2
(r
i
)
. The mathematics resembles the problem of nding the magnetic eld of a line
current located at a given distance r
i
from the center of a cylinder of high permeability
material and whose radius R is greater than r
i
[48]. The analogy is complete upon
performing the change of coordinates r '(r) and identifying the gradient of the
bond angle

(u) with the magnetic eld. This is explained in detail in Appendix C


and D where we introduce a conjugate function (u) analogous to the vector potential
that simplies the analysis of this problem. Thus, each of the N
d
defects will also
interact with an equal number of image defects. This situation can be described
Chapter 2: Bond-orientational order on a corrugated substrate. 52
mathematically by deriving an appropriate Greens function
N
that includes the
images, as discussed in Appendix D (see Eq.(D.19)). The resulting free energy F
N
reads:
F
N
K
A
=
1
2
N
d

j=i
q
i
q
j

N
(x
i
; x
j
) +F
0
+
N
d

i=1
q
i
(1
q
i
4
)V (r
i
) +
N
d

i=1
q
i
2
4
ln
_
'(R)
a
_

N
d

i=1
q
i
2
4
ln
_
1 x
2
i

, (2.47)
where F
0
is dened in Eq.(2.17) and the Greens function
N
(x
i
; x
j
) is given by

N
(x
i
; x
j
) =
1
4
ln
_
x
2
i
+x
2
j
2x
i
x
j
cos (
i

j
)

1
4
ln
_
x
2
i
x
2
j
+ 1 2x
i
x
j
cos (
i

j
)

.
(2.48)
The last term accounts for the interaction with the image defects and the superscript
N indicates Neumann boundary conditions on an appropriate potential function.
Here, we use scaled coordinates in the conformal plane x
i

(r
i
)
(R)
. The interaction of
the defects with the curvature is not aected by the presence of the distant wall.
To provide an illustration of the combined eect of curvature and boundary
conditions on tangential vector order, we rst consider the simpler case of a nematic
order parameter with periodicity equal to . This simplied model neglects dierences
in the elastic constants for bend and splay and does not incorporate any eect due to
the uniaxial coupling of the nematogens to the curvature. In this case, minimization
of the logarithmically diverging part of the free energy (fourth term in Eq. 2.47)
Chapter 2: Bond-orientational order on a corrugated substrate. 53
suggests that there will be only two disclinations of charge q = displaced along a
radial direction (see Fig. 2.13b). By applying Eq.(2.47), we can parameterize the
energy of the system in terms of the scaled radial coordinates, x
1
and x
2
of the two
disclinations. The resulting energy landscape is plotted in Fig. 2.13 (for = 2 and
R = 7r
0
) and clearly reveals two minimal-energy congurations. The rst minimum
corresponds to one disclination conned at the top of the bump (slightly shifted
from the center) and the other at a radial distance approximately 70% of R (see
Fig. 2.13b bottom panel). The second minimum corresponds to a fully deconned
state with both disclinations placed symmetrically at approximately 67% of R (see
Fig. 2.13b top panel). As the aspect ratio is raised even further, the saddle in the
energy landscape of Fig. 2.13a becomes a minimum corresponding to a conguration
in which both disclinations are conned in the cup of positive Gaussian curvature by
the geometric potential.
As illustrated in Fig. 2.14, there is a critical value of the aspect ratio,

D
1.5, above which it is energetically favorable for the system to have one discli-
nation conned at the top of the bump. For <
D
the fully deconned congu-
ration becomes energetically favorable, but the two minima can still coexist. As
is decreased even further, the repulsion between the two disclinations overcomes the
conning force of the geometric potential and makes the second minimum in Fig.
2.14a (corresponding to the partially conned conguration) disappear altogether.
This spinodal point occurs for 0.9 on the Gaussian bump. The specic values
of the critical aspect ratios are geometry dependent, but the generic mechanism of
Chapter 2: Bond-orientational order on a corrugated substrate. 54
2
4
6
2
4
6
0
1
2
2
4
6
F/k
A
x
1
x
2
(a)
(b)
Figure 2.13: (a) The free energy for a nematic (double headed vector eld) living
on a Gaussian bump surrounded by an aligning circular wall is plotted for = 2 as
a function of the scaled radial coordinates x
1
and x
2
of the two disclinations. The
radial coordinates have been scaled by r
0
and the size of the system is R = 7r
0
. Note
that the energy plot is symmetric with respect to the line x
1
= x
2
. (b) Schematic
illustration of the positions of the two disclinations (black dots) corresponding to the
deep energy minima at positions x
1
= 0.04 and x
2
= 4.9 (or viceversa) and to a
shallow minimum at x
1
= x
2
= 4.7. The two defects are on opposite sides of the
bump. The continuous line corresponds to the circular boundary while the dashed
one to the circle of zero Gaussian curvature and radius r
0
(drawing not to scale).
Chapter 2: Bond-orientational order on a corrugated substrate. 55
0 0.25 0.5 0.75 1 1.25 1.5 1.75
0.6
0.4
0.2
0
0.2
0.4
-
-
-
F/k
A
!

!
D
Figure 2.14: Plot of the free energy of a nematic (double headed vector eld) on a
Gaussian bump encircled by an aligning wall as a function of . The dotted line
represents the energy of the fully deconned conguration in Fig. 2.13b top panel
while the continuous line corresponds to the defect pattern illustrated in the bottom
panel of Fig. 2.13b. The energy of the fully deconned conguration is approximately
independent of because the two disclinations are far away from the bump.
deconnement depends only on a large separation of the length scales r
0
and R that
control the interaction with the curvature and the boundary respectively.
The analysis for the hexatic case is complicated by the fact that more defect
congurations are possible when six defects are present. We start by noting that even
in at space ( = 0) there are two natural low energy defect congurations with high
symmetry: the ground state corresponding to the six defects sitting at the vertices of
an hexagon and a higher energy state given by a pentagonal distribution of defects
with the sixth defect sitting at the center of the circular sample (see Fig. 2.15b). As
the aspect ratio is raised, the pentagonal arrangement becomes energetically favored
since it pays to have a defect conned in the (geometric) potential well at the origin
(see Fig. 2.15a). To study the transition, it is useful to derive expressions for the
energy of the two defect congurations as a function of the radius of the outer defect
Chapter 2: Bond-orientational order on a corrugated substrate. 56
ring r. Every defect (except the one at the origin, possibly) has the same scaled
coordinate x
i
= x and the angles between two defects are integer multiples of
2
n
where n is the number of defects in the outer ring (n = 5 for the pentagon and n = 6
for the hexagon). In this case, the sums involved in the rst (interaction) term in
Eq.(2.47) can be eciently evaluated using the following identity:
1
2
n1

i
ln
_
p
2
+ 1 2p cos(
2i
n
)
_
= ln (1 p
n
) ln (1 p) . (2.49)
Upon using Eq.(2.49) with p = 1 and p = x
2
to evaluate the sums arising from the
rst and the second term of the Greens function in Eq.(2.48) respectively, we obtain
the free energy F
H
for the hexagonal conguration:
F
H
()
K
A
=

6
ln
_
_
'(r)
'(R)
_
5

_
'(r)
'(R)
_
17
_


6
ln 6
+
11
6
V (r) +

6
ln
_
'(R)
a
_
. (2.50)
The free energy for the pentagonal conguration F
P
is readily obtained after similar
manipulations
F
P
()
K
A
=
5
36
ln
_
_
'(r)
'(R)
_
6

_
'(r)
'(R)
_
16
_


2
ln 2
+
11
36
V (0) +
55
36
V (r) +

6
ln
_
'(R)
a
_
. (2.51)
Note that these manipulations are very similar to the ones necessary to describe
superuid helium in a cylinder of radius R [50]. In fact, the superuid problem is
analogous to the case of hexatic order with free boundary conditions on a circular
boundary of radius R (see Appendix D). The rather unusual form of the argument
Chapter 2: Bond-orientational order on a corrugated substrate. 57
0 0.5 1 1.5 2
- 1
- 0.8
- 0.6
- 0.4
F/k
A
!
!
D
R
R
r0
r0
(a)
(b)
Figure 2.15: Plot of the free energy of an hexatic phase (draped on the Gaussian bump
encircled by a wall) as a function of . The dotted line represents the energy of the
hexagonal conguration illustrated in the top panel on the left while the continuous
line corresponds to the pentagonal arrangement in the bottom panel. The outer defect
rings in both congurations are approximately 90% of R. The critical aspect ratio

D
corresponding to the deconnement transition discussed in the text is indicated
by the dashed line.
of the logarithm in Equations (2.50) and (2.51) arises from the sum over the image
defects whose positions depend non-linearly on the position of the defects themselves.
Minimization of Equations (2.50) and (2.51) with respect to r xes the
distance of the outer defects. The resulting minimal energies F
P
and F
H
are plotted
as a function of in Fig. 2.15a. The critical value
D
, for which F
P
< F
H
can be
easily estimated by realizing that F
H
is approximately independent of because the
disclinations are far from the bump. On the other end, F
P
decreases with because
the conned disclination is trapped in a potential well whose depth is approximately
given by
11
144

2
(see the second term of Eq.(2.51) and the low expansion for
V (0) derived in Eq.(2.26)). The critical aspect ratio,
D
, for which the deconnement
transition occurs can be estimated by setting the depth of this potential well equal to
the energy dierence between the hexagon and pentagon congurations in at space.
Chapter 2: Bond-orientational order on a corrugated substrate. 58
The latter can be read o from the energy diagram in Fig. 2.15a and the result is
approximately 0.3K
A
which leads
D
1.1 in agreement with the value indicated in
Fig. 2.15a.
As the aspect ratio is raised even further, less symmetric defect congura-
tions become energetically favored corresponding to a larger number of disclinations
conned in the cup of positive Gaussian curvature. For example, when two discli-
nations are conned within r = r
0
, the outer defect ring is given by four defects
approximately located at the vertices of a square. We note that these defect congu-
rations cease to exist at low aspect ratios because they require the geometric potential
to overcome the strong repulsive interaction between the conned defects. As dis-
cussed earlier for Fig.(2.14), the actual values of the aspect ratios involved depend
on the specic geometry of the substrate. However the basic mechanism behind the
deconnement transition is more general.
Note that, as increases, the geometric mechanism of defect dipole unbind-
ing discussed in the last section may also set in. Because of the presence of one or
more positive defects at the top of the bump, the critical aspect ratio necessary to
unbind one dipole will be larger than what was calculated before. If dipole unbinding
does occur, the new defects will decorate the existing patterns by adding new posi-
tively charged disclinations in the region of positive Gaussian curvature and expelling
the negative ones in the external region of the bump (r > r
0
) where the Gaussian
curvature is also negative (see Fig.2.1).
Chapter 2: Bond-orientational order on a corrugated substrate. 59
2.5 Conclusion
We have discussed how the varying curvature of a surface such as a Gaus-
sian bump can trigger the generation of single defects or the unbinding of dipoles,
even if no topological constraints or entropic arguments require their presence. This
mechanism is independent of temperature if the system is kept well below its Koster-
litz Thouless transition temperature. It would be interesting to revisit Kosterlitz-
Thouless defect unbinding transitions on surfaces of varying Gaussian curvature in
the presence of a quenched topography [15] in the light of the present work. One
might also explore the dynamics of the delocalization transition that occurs when a
bump is conned by a circular edge and the aspect ratio is lowered until the defects,
initially conned on top of the bump by the geometric potential, are forced to slide
towards the boundary. Quantitative studies of periodic arrangements of bumps would
be interesting and could be inspired by fruitful analogies with methods and ideas from
solid state physics.
We also hope to extend this work by considering crystalline order on bumpy
topographies and taking explicitly into account the screening of clouds of dislocations
and possible generation of grain boundaries [4]. Such an analysis would facilitate com-
parison with experiments performed with a single grain of block copolymer spherical
domains
3
on a suitably patterned substrate [30, 37].
3
Block copolymers are formed by blocks of the same monomer unit (labeled by A) covalently bound to
sequences of an unlike type (labeled by B). By controlling the relative volume fraction of the two blocks,
one can engineer a self-assembled ordered phase composed of spheres of block A in a sea of B monomers.
The typical radius of the resulting block-copolymer spherical cores is of the order of a few nanometers and
their spacing tens of nanometers. These values can be tuned by suitably choosing the block-copolymers and
varying their volume fraction.
Chapter 3
Liquid crystal textures in thick
spherical shells.
Based on V. Vitelli and D. R. Nelson, cond-mat/0604293
60
Chapter 3: Liquid crystal textures in thick spherical shells. 61
3.1 Introduction
The study of liquid crystal phases benets from geometrical reasoning in
two important ways. Firstly, liquid crystal elasticity can often be cast in terms of the
curvature of equipotential lines (or surfaces) that map out the corresponding textures.
Second, the observed textures are strongly aected by geometric and topological
constraints imposed by the presence of boundaries conning the system. The liquid
crystal ground state results from the competition between the energetic requirement
of minimizing the curvature of the texture and the geometric frustration introduced
by boundaries that impart a preferred curvature at the edge of the sample that often
cannot propagate across the system [38, 47, 51].
The boundary conditions can be controlled experimentally with the possibil-
ity of designing molecular systems with intriguing technological applications [52]. For
example, colloidal particles coated by a very thin nematic layer have in their ground
state four disclinations sitting at the vertices of a tetrahedron. Each coated colloidal
particle can then be viewed as the fundamental building block of a self assembled
lattice with tetravalent coordination. The bonds between the colloidal particles
could be provided by chemical linkers attached at the four bald spots at the cores
of the four disclinations present in each colloid [53]. A second example, is provided by
self-assembled systems of block copolymers [30] which are a promising tool for soft
lithography on both at and curved substrates [54]. In addition, liquid crystals in
conned geometries provide an arena for physicists and mathematicians interested in
applications of geometrical and topological ideas to material science [42, 55, 56].
Chapter 3: Liquid crystal textures in thick spherical shells. 62
In this work we present a theoretical study of liquid crystal phases (focusing
on vector, nematic and hexatic order) conned in a spherical shell of varying thickness
with the director assumed to be tangent to the two interfaces. We rst consider the
two dimensional regime where a nematic lm coats a quenched spherical surface
such as a colloidal particle in solution or the interface of, say, a water droplet in
oil. The presence of topological defects in the ground state for ordered states on
spherical surfaces is unavoidable [31, 57, 12] . Recent experimental and theoretical
investigations of spherical crystallography have provided an alternative context to
study the constraints posed by the compactness of the underlying curved space [4,
3]. More recent explorations have concentrated on 2D ordered phases conned to
interfaces of varying Gaussian curvature [39, 58, 59] as well as dynamically uctuating
surfaces [35, 36].
As the thickness of a nematic lm increases, an escaped three dimensional
texture, also strongly inuenced by the spherical topology and the boundary condi-
tions, become energetically favored with respect to planar textures. This instability
destabilizes the tetravalent nematic texture on colloids. In this paper, we estimate
the thickness of the nematic lm above which a texture with four radial disclination
lines of charge s =
1
2
becomes unstable to four half-hedgehogs. The two competing
textures studied in this paper are shown in Fig. 3.1. We also discuss the possibility
of hysteresis between the two textures.
The organization of this paper is as follows. In section 3.2 we derive exact
solutions for the ground state of spherical lms of tilted molecules and nematogens
Chapter 3: Liquid crystal textures in thick spherical shells. 63
1mm
Middle
Liquid
Inner Liquid
Outer
Liquid
0.05mm
Figure 3.1: (Top panel) Two-dimensional texture characterized by four short disclina-
tion lines at the vertices of a tetrahedron inscribed in the sphere. The surface texture
shown (inscribed on the surface of a baseball) is invariant throughout the thickness
of the shell. (Bottom panel) Cut view of the escaped three dimensional texture given
by two pairs of half hedgehogs located at the north and south poles of the sphere.
Chapter 3: Liquid crystal textures in thick spherical shells. 64
within isotropic elasticity by using the method of conformal mappings. In the nota-
tion of References [53, 12], these situations correspond to order parameters described
by a bond angle with p = 1, 2 fold symmetry in the tangent plane of the sphere (see
Appendix A). A mathematical justication for our approach is provided in Appendix
B, where the same technique is illustrated in the context of a more familiar at space
problem. In section 3.3 we study the stability of liquid crystal textures to thermal
uctuations by means of a normal mode analysis whose details are relegated to Ap-
pendices A and C. The stability of the valence-four texture against escaped solutions
is considered in section 3.4 where a phase diagram is derived with the thickness as a
control parameter. The texture distortions caused by the elastic anisotropy between
bend and splay deformations are briey considered in section 3.5.
3.2 Textures
The liquid crystal free energy for molecules embedded in an arbitrary frozen
surface can be written in the one constant approximation as
F =
K
2
_
dAD
i
n
j
(u)D
i
n
j
(u) , (3.1)
where u = {u
1
, u
2
} is a set of internal coordinates, n(u) is the liquid crystal director
dened in the tangent plane, D
i
is the covariant derivative with respect to the metric
of the surface and dA is the innitesimal surface area [33, 32, 35, 40]. The free energy
of Eq.(3.1) is invariant upon rotating each molecule n(u) by the same (arbitrary)
angle with respect to any axis of rotation perpendicular to the local tangent plane.
The treatment of systems with a p-fold symmetry is straightforward provided that the
Chapter 3: Liquid crystal textures in thick spherical shells. 65
one Frank constant approximation is used for p = 1 and p = 2 and the consequences
of any additional couplings to curvature neglected [32]. This choice of free energy
implies that the minimal energy conguration will be given locally by neighboring
n(u) vectors which dier only by parallel transport. The curvature of the surface
induces frustration in the texture. In fact, by Gauss Theorema egregium [42],
tangent vectors parallel transported along a closed loop are rotated by an amount
equal to the Gaussian curvature integrated over the enclosed area. On a sphere, this
theorem insures that the nematic ground state always has four excess disclinations
[31, 12]. More generally, the sum of the topological charges on any closed surface
is equal to the integrated Gaussian curvature, implying a minimum of 2 and 12
disclinations in the ground state of tilted molecules and hexatics, respectively.
We introduce a local angle eld (u), corresponding to the angle between
n(u) and an arbitrary local reference frame, we can rewrite the free energy introduced
in Eq.(3.1) as:
F =
1
2
K
_
dS g
ij
(
i
A
i
)(
j
A
j
) , (3.2)
where dS = d
2
u

g, g is the determinant of the metric tensor g


ij
and A
i
is the spin-
connection whose curl is the Gaussian curvature G(u) [40, 42]. On a sphere of radius
R parametrized by polar coordinates (, ), the only non vanishing components of the
(inverse) metric tensor are g
rr
=
1
Rsin
and g

=
1
R
. A convenient choice of of the
spin connection (which plays the role of the vector potential) is discussed in Appendix
A. The simplied free energy in Eq.(4.3) is the starting point of our analysis.
Chapter 3: Liquid crystal textures in thick spherical shells. 66
3.2.1 Tilted molecules on a sphere
The orientational order of molecules tilted by a constant angle with respect
to a spherical interface can be modelled by a vector eld n(, ) dened in the local
tangent plane on which the molecule has a xed length projection [57]. To determine
the ground state of the liquid crystal texture, we minimize the Frank free energy of
Eq.(4.3). As discussed above, the topological charges must sum up to 4, the inte-
grated Gaussian curvature of the sphere [40, 42]. For a vector eld (p = 1) the texture
with only two defects of charges +2 minimizes the Frank free energy and satises
the topological constraint. Since the defects repel each other they preferentially sit
at two antipodal points that we can designate as the north and south pole of the
sphere. If the splay and bend coupling constants of the nematic are equal, then there
is a large degeneracy in the ground state arising from the invariance of the vector
free energy in Eq.(4.3) under global rotations (u)(u)+c, where u , . One
representative texture is a sink and a source of n(u) at the two poles. In this
splay rich texture n(u) is parallel to the lines of longitude on a sphere. In a bend rich
texture, related to the previous by a

2
rotation about the local normal to the sur-
face, n(u) is everywhere parallel to the lines of latitude. Any other rotation of n(u)
that makes an arbitrary constant angle with respect to this texture is an acceptable
solution for the ground state of the molecules.
As we now show, this degeneracy is lifted when K
3
,= K
1
. Indeed the eect
of distinct splay and bend elastic constants K
1
and K
3
(the twist elastic constant
K
2
is absent in two dimensions) is to select the bend-rich texture if K
1
> K
3
or
Chapter 3: Liquid crystal textures in thick spherical shells. 67
the splay-rich one if K
3
> K
1
. The intermediate congurations obtained by a global
rotation of the director are now unstable. Assume for simplicity that K
3
> K
1
. In
this case, it is convenient to recast the Frank free energy (see Appendix A) as follows:
F =
1
2
_
d
2
x

g [K
1
_
D
i
n
j
) ( D
i
n
j
_
+(K
3
K
1
)(Dn)
2
] , (3.3)
where the covariant derivatives is expressed in terms of the Christoel connection,

j
i t
,
D
i
n
j
=
i
n
j
+
j
i t
n
t
, (3.4)
and the covariant form of the curl squared is [40, 60],
(

D n)
2

_
D
i
n
j
D
j
n
i
) ( D
i
n
j
D
j
n
i
_
. (3.5)
The rst term in Eq.(3.3) resembles the Frank free energy in the one coupling constant
approximation and is minimized by choosing the sink-source (or lines of longitude)
solution. The second term (which is positive denite) will vanish for this texture since
the sink-source texture is bend free. All other textures have a higher energy.
A similar argument can be used to prove that the two vortex-conguration
which follows the lines of latitude is the minimum of the free energy when K
1
> K
3
by rewriting the Frank free energy as
F =
1
2
_
d
2
x

g [K
3
_
D
i
n
j
) ( D
i
n
j
_
+(K
1
K
3
) (D n)
2
] , (3.6)
Chapter 3: Liquid crystal textures in thick spherical shells. 68
where the covariant form of the divergence reads
D n
1

i
_

g n
i
_
. (3.7)
The latitudinal texture minimizes the rst term of Eq.(3.6) while the second vanishes
because this texture is splay free. Any deviation from the splay-free latitudinal texture
will only increase the energy.
The energy of both textures can be expressed as a function of the anisotropy
parameter, , and the mean of the elastic constants, K,

K
3
K
1
K
3
+K
1
, (3.8)
K
K
3
+K
1
2
, (3.9)
and the radius of the sphere, R, scaled by the short distance cuto a. The resulting
free energy for arbitrary reads (see Eq.(E.16))
F = 2K (1 | |)
_
ln
_
R
a
_
0.3
_
. (3.10)
The conclusions of this section are summarized in Fig. 3.2 which suggests that there
is a discontinuous rst order transition when passes through zero. This analysis
mirrors similar arguments valid in the plane [61].
3.2.2 Nematic Texture
The nematic texture of very thin spherical shells of nematic liquid crystal
with tangential boundary conditions can be analyzed within the one Frank constant
approximation by using the method of conformal mappings whose mathematical jus-
tication is illustrated in Appendix B by means of a simpler example.
Chapter 3: Liquid crystal textures in thick spherical shells. 69
F
-1 1

Figure 3.2: Schematic illustration of the phase diagram of the texture of tilted
molecules on a sphere as a function of the anisotropy parameter superimposed
on a plot of the free energy, F, stored in the texture versus . The two competing
ground states (top view) are characterized by either pure bend (lines of longitude
conguration at left) or pure splay (lines of latitude conguration at right).
Chapter 3: Liquid crystal textures in thick spherical shells. 70
An elegant argument introduced by Lubensky and Prost in Ref.[12] shows
that the ground state of nematogens on a sphere is given by 4 disclinations of topo-
logical charge s = 1/2 sitting at the vertexes of a tetrahedron. The energy of single
disclinations is proportional to the square of its strength. As a result, the longitudi-
nal and latitudinal textures derived for tilted molecules in section 3.2.1 are unstable
since their energies can be lowered by splitting each s = 1 defect at the north and
south pole into two s = 1/2 disclinations and letting them relax to their equilibrium
positions at the vertexes of a tetrahedron where they are as far away from each other
as possible. According to a calculation in Ref.[12], the energy F
s
of a sphere of radius
R with in plane orientational order and 2p interacting minimal disclinations for a
p-fold order parameter is given by
F
s
= 2K h
_
1
p
ln
_
4p
2
R
a
_
+c
p
_
. (3.11)
where the {c
p
} are constants depending on the symmetry of the order parameter and
the defect core energy while h is the thickness of the liquid crystal layer
1
. When
p = 2 is chosen in Eq.(3.11) the elastic energy is indeed smaller than the corresponding
value for p = 1 in the limit R a, in agreement with related arguments given in
Ref.[53].
To obtain an algebraic expression for the texture we proceed as illustrated in
Appendix B and seek a function (x, y, z) = (x, y, z)+i(x, y, z) which is harmonic
on the sphere except for two arcs connecting the defects in pairs. The calculation
for nematogens described below was suggested to us by F. Dyson [62]. The function
1
The numerical values of the relevant constants are c1 = 0 and c2 0.2.
Chapter 3: Liquid crystal textures in thick spherical shells. 71
Figure 3.3: Schematic illustration of the baseball texture of a thin nematic shell. The
same texture is reproduced from a dierent perspective in the top panel of Fig. 3.1.
(x, y, z), which we can interpret as an electrostatic potential, takes equal and op-
posite values on the two arcs and is equal to zero on a baseball-like seam (see Fig.
3.3) which divides the sphere into two congruent regions. The nematic director is
then oriented (up to a global rotation) along the contour lines of (x, y, z), that is
the equipotential lines of this curved space capacitor. In this analogy, the contour
lines of (x, y, z) are electric eld lines, hence they correspond to a valid texture
where the director is rotated locally by

2
with respect to the equipotential lines. The
arcs can be either great-circle arcs extending more than half way round the sphere or
short great-circles arcs connecting the same pair of defects along the shortest path.
The rst choice leads to equipotential lines whose seam resembles in shape that of a
Chapter 3: Liquid crystal textures in thick spherical shells. 72
N
Conformal plane
S
Figure 3.4: Graphical construction of the stereographic projection. Regions close to
the north pole have larger images in the conformal plane than regions of equal areas
close to the south pole. The stereographic projection preserves the topology of the
surface provided all points at innity are identied with the north pole.
baseball. If the second choice is made the pattern of equipotential lines would not
deviate much from concentric circles and the seam would look more like the seam of
a cricket ball. We will explicitly show that the two choices are equivalent since the
equipotential lines of the rst solution are eld lines of the second and vice versa.
We choose the arcs connecting the defect pairs along great circles and we
take the four defects labelled by A, B, C, D to lie at the vertices of a tetrahedron
inscribed on a sphere of radius 1 and whose north and south poles are N = (0, 0, 1)
and S = (0, 0, 1) respectively
A =
1

3
(1, 1, 1) , B =
1

3
(1, 1, 1) ,
C =
1

3
(1, 1, 1) , D =
1

3
(1, 1, 1) . (3.12)
We now perform a stereographic projection (see Fig. A.1) that maps every point on
Chapter 3: Liquid crystal textures in thick spherical shells. 73
a unit sphere centered on the origin onto the plane z = 1 according to the rule
_
_
_
_
_
_
_
x
y
z
_
_
_
_
_
_
_

_
_
_
_
_
_
_
a
b
1
_
_
_
_
_
_
_
. (3.13)
The coordinates of the image points (connected to points on the sphere by dashed
lines in Fig. A.1) are given by
a =
2x
1 z
,
b =
2y
1 z
. (3.14)
Upon transforming to a complex coordinate w = a + ib, the four tetrahedral points
of Eq.(3.12) are mapped onto
A

= p (1 +i) , B

= p (1 i) ,
C

= q (1 +i) , D

= q (1 i) , (3.15)
where
p =

3 + 1 , q =

3 1 . (3.16)
(In this section p does not refer to the symmetry of the order parameter). The
great arc passing through the south pole (corresponding to one capacitor plate in the
electrostatic analogy) maps onto the segment A

, as illustrated schematically in the


top panel of Fig. 3.5, while the great arc through the north pole maps onto the two
semi-innite segments of the line a +b = 0 which bracket C

. We can fold the two


cuts in the w plane back on top of each other by mapping the w plane onto the u
Chapter 3: Liquid crystal textures in thick spherical shells. 74
A'
B'
C'
D'
-2q
2
2p
2
w
u
v
k -k 1 -1
C
D
A
B
Figure 3.5: Illustration of the change in the branch cut structures of the complex
function (v) describing the nematic texture after performing a series of conformal
transformations. In the top panel we have simply performed a stereographic pro-
jection from the sphere. The middle panel shows the fold up transformation of
Eq.(3.17) whereas the bottom panel corresponds to the transformation in Eq.(3.21)
that symmetrizes the positions of the cuts.
Chapter 3: Liquid crystal textures in thick spherical shells. 75
plane via
u = iw
2
. (3.17)
As shown in the middle panel of Fig. 3.5, the images

A and

B of A and B now both
lie on the real axis at 2p
2
while the images of C and D now lie at 2q
2
. The two cuts
in the u plane are both on the real axis, running from zero to 2p
2
and from 2q
2
to
minus innity. On the sphere, these correspond to geodesics connecting defects which
stretch more than halfway around the sphere. In order to make the cuts symmetric
with respect to the imaginary axis (see the bottom panel of Fig. 3.5) we search for a
conformal transformation that maps the following four points in the complex u-plane
to four points on the real axis of a complex v-plane
u
0
= 0 v
0
= k ,
u
1
= 2q
2
v
1
= k ,
u
2
= v
2
= 1 ,
u
3
= 2p
2
v
3
= 1 . (3.18)
In order to fully determine the conformal transformation we need to determine the
value of k. This can be done by using a standard relation in the theory of conformal
transformations [45]
u
0
u
1
u
0
u
2
u
3
u
2
u
3
u
1
=
v
0
v
1
v
0
v
2
v
3
v
2
v
3
v
1
. (3.19)
Upon inserting the points of Eq.(3.18) into Eq.(3.19), we determine the value of k
(less than one)
k =
2

2 p
p + 2

2
. (3.20)
Chapter 3: Liquid crystal textures in thick spherical shells. 76
Equation (3.18) contains four independent relations so we are still left with three
conditions to determine the three independent coecients {, , } of the bilinear
conformal transformation that implements the mapping illustrated pictorially in the
bottom plate of Fig. 3.5
v =
u +
u +
. (3.21)
The required coecients needed to implement the mapping in Eq.(3.18) are
= 1 ,
= 2p (2

2 +p) ,
= 2p (2

2 p) , (3.22)
To solve Laplaces equation, we desire a function (v) which is analytic except on
the two cuts on the real axis, and whose real part takes constant values on the cuts.
By symmetry, (v) is an odd function of v, and its real part (v) is zero when the
real part of v is zero. Therefore the image of the seam in the v plane is simply the
imaginary axis 'e v = 0. Upon substituting for v using Equations (3.21) and (3.17),
the condition 'e v = 0 becomes
16 + 4p
2
m(w
2
) |w|
4
= 0 , (3.23)
With the help of Equations (3.14) and (3.13), we can now write down the equation
of the seam explicitly in the original cartesian coordinates [62]
z = (2 +

3) xy , (3.24)
Chapter 3: Liquid crystal textures in thick spherical shells. 77
a
b
c
d
Figure 3.6: Dierent views of a track of parallel nematogens which partitions the
sphere into two equal areas, each containing two s =
1
2
disclination defects. It resem-
bles a fattened version of the seam of a baseball.
or in spherical polar coordinates {, } as
cos
sin
2

=
_
1 +

3
2
_
sin 2 . (3.25)
The seam dened by the line of zero potential, is represented for dierent orientations
of the sphere in Fig. 3.6. Its contour length l, on a unit radius ball, is readily
calculated upon integrating the expression for the innitesimal arc of the seam
dl =

sin
2

_
d
d
_
2
+ 1 d , (3.26)
from
min
0.69 radians to
max
2.44 radians and multiplying the result by four
in view of the symmetry of the seam. The values of
min
and
max
are obtained
from Eq.(3.25) by setting equal to

4
and
3
4
respectively. Upon using Eq.(3.25) to
substitute () in Eq.(3.26), we obtain l 9.09 for a sphere of unit radius. The seam
Chapter 3: Liquid crystal textures in thick spherical shells. 78
is longer than the equatorial circumference by slightly less than 50%.
The branch cut structure in the v plane is suciently simple to allow a guess
of the corresponding analytic function (v). A function with cuts from k to 1 and
k to 1, whose real part is equal and opposite on the two cuts and with a single
imaginary period around any curve separating the cuts is easily identied to be a
standard elliptic integral
(v) =
_
v
0
_
(k
2
t
2
)(1 t
2
)

1
2
dt . (3.27)
with v given in terms of w by Equations (3.17) and (3.21). The nematic director is
oriented (up to a global rotation) along the contour lines of the imaginary or (real
part) of (v). The equipotential (red) and eld lines (black) of (v) are conveniently
plotted using the stereographic projection plane w = a+ib in Fig. 3.7 along with the
positions of the disclinations (green dots). It is easy to switch from the stereographic-
projection plane w = a +ib of Fig. 3.7 to spherical polar coordinates (, ) by using
the relation (reviewed in Appendix B),
w = 2Rcot
_

2
_
e
i
, (3.28)
where R is the radius of the sphere.
If we had constructed the base-ball with cuts along the short geodesics
connecting the defects, then the form of the texture given in Eq.(3.27) would be the
same, but the parameter k in the elliptic integral would be given by
k = (

2)
2
(

2 + 1)
2
, (3.29)
Chapter 3: Liquid crystal textures in thick spherical shells. 79
-4 -2 0 2 4
-4
-2
0
2
4
Figure 3.7: Illustration of the nematic texture in stereographic projection (Color
online). The red and black eld lines correspond to two energetically degenerate (in
the one Frank constant approximation) families of bend and splay rich textures. The
dots indicate the four tetrahedral s =
1
2
disclination defects.
Chapter 3: Liquid crystal textures in thick spherical shells. 80
instead of Eq.(5.13). The equation for the seam becomes
z = (2

3) xy (3.30)
and the corresponding equipotential and eld lines are plotted in black and red re-
spectively in Fig. 3.7. The expression reads
cos
sin
2

=
_
1

3
2
_
sin 2 , (3.31)
in spherical polar coordinates, and leads to the same contour length l as Eq.(3.25).
Thus the two choices of arcs lead to two equivalent textures diering only by a

2
about the local normal. As in Section 3.2.1, the degeneracy in energy between the
red and black ow lines in Fig. 3.7 is lifted upon considering the eect of elastic
anisotropy generated by a dierent energy cost for bend and splay.
3.3 Stability of liquid crystal textures to thermal uctuations
In this section, we study the stability of liquid crystal ground states to
thermal uctuations [53]. To explore the delity of directional bonds at nite tem-
peratures, we employ a Coulomb gas representation of the liquid crystal free energy
(in the one Frank constant approximation) obtained by substituting in Eq.(4.3) the
relation

) = s(u) G(u) n(u) , (3.32)


where

is the covariant antisymmetric tensor [40], G(u) is the Gaussian curvature


and s(u)
1

N
d
i=1
q
i
(u u
i
) is the disclination density with N
d
defects of charge
Chapter 3: Liquid crystal textures in thick spherical shells. 81
q
i
at positions u
i
. The nal result is an eective free energy whose basic degrees of
freedom are the defects themselves [35, ?, 33]:
F =
K
2
_
dA
_
dA

n(u) (u, u

) n(u

) . (3.33)
The Greens function (u, u

) is calculated (see Appendix A) by inverting the Lapla-


cian dened on the sphere
(u, u

)
_
1

_
uu

, (3.34)
and we have suppressed for now defect core energy contributions which reect the
physics at microscopic length scales. Equations (3.32) and (3.33) can be understood
by analogy to two dimensional electrostatics, with the Gaussian curvature G(u) (with
sign reversed) playing the role of a uniform background charge distribution and the
topological defects appearing as point-like sources with electrostatic charges equal
to their topological charge q
i
2
. On a generic surface, the defects tend to position
themselves so that the Gaussian curvature is screened: typically, the positive ones are
attracted to peaks and valleys while the negative ones to the saddles of the surface
[58, 59]. This geometric potential is ruled out by symmetry on an undeformed sphere
since the Gaussian curvature is constant. The Gaussian curvature plays the role of
a uniform background charge xing the net charge of the defects consistent with the
topological constraint imposed by the Poincare-Hopf theorem (see section 3.2 and
References [42, 63]). The equilibrium positions of the defects are then determined
only by defect-defect interactions which are proportional to the logarithm of their
2
The charge q can be dened by the amount increases along a counterclockwise path enclosing the
defects core.
Chapter 3: Liquid crystal textures in thick spherical shells. 82
chordal distance (see Appendix A) according to
F =
K
2p
2

i=j
n
i
n
j
ln [1 cos
ij
] , (3.35)
where the integers n
i
and n
j
describing the singularities associated with each defect,
and the integer p controls the period
2
p
of the orientational order parameter. The
topological charge describing the rotation of the order parameter around each defect
is given by s
j
=
n
p
. The geodesic angle
ij
subtended by the two defects at positions
u
i
= {
i
,
i
} and u
j
= {
j
,
j
} can be conveniently recast in terms of their spherical
polar coordinates
cos
ij
= cos
i
cos
j
+ sin
i
sin
j
cos [
i

j
] . (3.36)
As a simple example, we rst consider the case of a Z = 2 colloidal particle
(p=1) with two antipodal defects of index 1. We can study the eect of thermal
disruption of the ground state by setting
ij
= + in Eq.(3.35) and expanding in
the bending angle . The resulting free energy, apart from an additive constant, reads
F
K
4

2
. (3.37)
Upon applying the equipartition theorem we obtain in the limit K k
B
T [53]
cos
ij
) 1 +
1
2

2
) ,
1 +
k
B
T
K
, (3.38)
which describes the delity of antipodal bonds of a divalent colloidal particle.
The eect of thermal uctuations on the tetrahedral ground state of nematic
molecules conned on the sphere (p=2 and all s
j
=
1
2
) can be studied by means of a
Chapter 3: Liquid crystal textures in thick spherical shells. 83
normal mode analysis. The basic results sketched in [53] were obtained by a slightly
dierent method. Here we describe an alternative treatment in some detail and extend
our analysis to hexatic and tetratic defect arrays (see Appendix C).
We start by dening a generalized array of defect coordinates {q
i
} as a 2N
dimensional vector, where N is the number of defects in the ground state (or, equiv-
alently, the valence of the colloidal molecule). N = 4 in the case of the tetrahedron.
The rst N entries of the vector q are the longitudinal deviations of the N defects
from a perfect tetrahedral conguration while the remaining N components describe
defect displacements along the lines of latitude of a sphere of unit radius. As a re-
sult, the deviations of the i
th
defect from its equilibrium conguration {
i
0
,
i
0
} are
parameterized by the two independent components of the vector q
q
i
=
i
,
q
N+i
=
i
sin(
i
0
) . (3.39)
The relations in Eq.(3.39) can be used to reexpress Eq.(3.36) in terms of the compo-
nents of the displacements vector q
i
, with the result,
cos
ij
= cos
_

i
0
+q
i
_
cos
_

j
0
+q
j
_
+
sin
_

i
0
+q
i
_
sin
_

j
0
+q
j
_
cos
_

i
0

j
0
+
q
N+i
sin
i
0

q
N+j
sin
j
0
_
. (3.40)
Upon substituting Eq.(3.40) in Eq.(3.35), the free energy F can be expanded around
the equilibrium conguration to quadratic order in q
i
with the result (apart from an
additive constant)
F
1
2

ij
M
ij
q
i
q
j
, (3.41)
Chapter 3: Liquid crystal textures in thick spherical shells. 84
where the matrix, M
ij
, describing the deformation of the tetrahedral molecule is
naturally dened as
M
ij
=
_

2
F
q
i
q
j
_
q
i
,q
j
=0
. (3.42)
The eigenvalues of this matrix can be classied according to the irreducible represen-
tation of the symmetry group of the tetrahedron; their degeneracies can be determined
purely from the group theoretical relation [64, 65]
n
()
=
1
g

i
g
i

()
i

()
i
, (3.43)
where n
()
is the number of frequency degenerate normal modes that transform like
the irreducible representation labeled by , g
i
is the number of symmetry operations
of the tetrahedral point group in the i
th
class, g =

i
g
i
= 24 is the total number of
symmetry operation in the group,
()
i
is the character of the i
th
class in the irreducible
representation labelled by while
()
i
is the corresponding character for the reducible
representation formed by the defects displacements.
The information necessary to apply Eq.(3.43) to a tetravalent colloid is col-
lected in Table 3.1. The top row contains the ve symmetry class {E, C
3
, C
2
, S
4
,
d
}
contained in the tetrahedral point group
d
, corresponding respectively to the iden-
tity, three and two fold rotations, four fold rotatory-reections and reection through
a plane of symmetry [64]. The number of symmetry operations g
i
included in the i
th
class also appears in the top row: thus, {g
i
} = {1, 8, 3, 6, 6} where the same ordering
used above to list the classes has been adopted.
The left most column of Table 3.1 lists the one, two and three dimensional
irreducible representations of the tetrahedral group {A
1
, A
2
, E, F
1
, F
2
}, along with
Chapter 3: Liquid crystal textures in thick spherical shells. 85
Table 3.1: Character for the irreducible representations of the tetrahedral point group
together with the character of the eight-dimensional representation generated by
the defect displacements of a tetravalent colloid.

d
E 8C
3
3C
2
6S
4
6
d
A
1
1 1 1 1 1
A
2
1 1 1 1 1
E 2 1 2 0 0
F
1
3 0 1 1 1
F
2
3 0 1 1 1
8 1 0 0 0
the eight dimensional representation generated by the defect displacements. The
entries of the table list the characters corresponding to each class of the ve irreducible
representations,
()
i
, and in the last row the corresponding characters,
()
i
, for the
eight dimensional representation. The former are tabulated from standard group
theoretical treatments while the latter needs to be worked out from the traces of the
transformation matrices that describe how the displacement coordinates q
i
transform
under the action of each symmetry element in the group. These manipulations are
rather cumbersome, especially for the icosahedral molecule arising when a spherical
surface is coated with a pure hexatic layer (see Appendix C).
In the rich literature on molecular vibrations a set of empirical rules has
been developed to write down the characters by examining only the transformation
of the three dimensional cartesian displacements of the few atoms whose equilibrium
Chapter 3: Liquid crystal textures in thick spherical shells. 86
positions are not altered by the symmetry operation. In Appendix C we provide
analogous rules that simplify the task of nding the
()
i
characters by incorporating
the constraint that each atom is conned on a sphere and hence only two orthogonal
displacements need to be considered as shown in Eq.(3.39).
The interested reader is referred to Appendix C for a more comprehensive
mathematical justication of the normal mode analysis applied to the tetrahedral
colloid and to the more complicated cases of hexatic Z = 12 and tetratic order
Z = 8. Here, we simply summarize the results of applying Eq.(3.43) in conjunction
with Table 3.1 to nd the degeneracies of the eigenvalue spectrum of the matrix M
ij
.
The representation contains (only once) the three dimensional representations F
2
and F
1
as well as the two dimensional representation E.
= F
2
+F
1
+E . (3.44)
The three normal coordinates with vanishing frequency correspond to the three rigid
body rotations and belong to the F
1
irreducible representation [64, 65]. We are left
with a doublet (E) and a triplet (F
2
) corresponding respectively to two shear-like
twisting deformations of the tetrahedron and to three stretching and bending modes
of the cords joining neighboring defects.
This symmetry analysis is conrmed by direct diagonalization of the matrix
M
ij
which leads the following set of eigenvalues
i
{
i
} =
3 K
8
{0, 0, 0, 1, 1, 2, 2, 2} . (3.45)
In Section C, we also list the eigenvectors w
i
of M
ij
. The displacement coordinates
Chapter 3: Liquid crystal textures in thick spherical shells. 87
are readily expressed in terms of the eigenvectors
q
i
= U
1
ij
w
j
, (3.46)
where the unitary matrix U diagonalizes M and hence the free energy of Eq.(3.41)
and is dened by
U M U
1
= Diag (
i
) . (3.47)
Its construction is easily achieved by the standard Gram-Schmidt orthogonalization
procedure to the eigenvectors {w
i
} = {w
1
, ..., w
8
}, where the same ordering chosen in
listing the eigenvalues in Eq.(3.45) is implicitly assumed. The resulting orthogonal
basis vectors are the rows of the 8 8 matrix U.
We are now in a position to evaluate cos
ij
) where the thermal average is
performed with the Boltzman weight obtained from the free energy in Eq.(3.41) which
is now diagonal. Note that for the tetrahedron any choice of pair of defects labelled
by i and j (where i ,= j) will lead to the same answer, unlike the less symmetric cases
of the twisted cube (p=4) and the icosahedron (p=6) considered in Appendix C. The
bending angle cos
ij
in Eq.(3.40) can be Taylor expanded in the q
i
. The resulting
expression is rather cumbersome, but once the displacements {q
i
} are reexpressed in
terms of the normal coordinates {w
i
} (by means of Eq.(3.46)), cos
ij
reduces to
cos
ij
=
1
3
+
2
9
_
w
2
6
+w
2
7
+w
2
8
_
, (3.48)
where the only eigenmodes {w
6
, w
7
, w
8
} appearing in Eq.(3.48) correspond to the
bending triplet of Eq.(3.45).
It is now easy to perform the thermal average by Gaussian integration of
Chapter 3: Liquid crystal textures in thick spherical shells. 88
the energetically degenerate eigenmodes, with the result [53]
cos
ij
) =
1
3
+
2
9
_
w
2
6
) +w
2
7
) +w
2
8
)
_
=
1
3
+
16k
B
T
9K
. (3.49)
3.4 Valence transitions in thick nematic shells
In this section, we study the crossover from a two to three dimensional
regime as the thickness of the spherical shell, h, increases. For thicker shells, three di-
mensional defect congurations (escaped in the third dimension) compete with the
planar textures described in the previous sections, leading to a structural transition
and a change in valence from Z = 4 to Z = 2 beyond a critical value of h.
We rst consider the case of a cylindrical slab (or disk) of radius R and
thickness h lled with a nematic whose director is tangent to the two circular faces
[66] (see Fig. 3.8). This simpler geometry captures the essential features of the
problem and provides a suitable starting point for understanding thin spherical shells
(see Fig. 3.9).
3.4.1 Slab geometry
To estimate the energy stored in the texture of Fig. 3.8, we coarse grain
the system to blobs of size h. The elastic energy arises from two sources: a long
distance contribution from a radial texture associated with an s = 1 disclination and
a local energy cost for the elastic deformations inside the spherical blob in Fig. 3.8.
In the one Frank constant approximation, the former can be estimated as the energy
Chapter 3: Liquid crystal textures in thick spherical shells. 89
Figure 3.8: Side view of the 2D nematic texture ansatz solution in Eq.(3.54). The
cylindrical slab has height h and radius R. The arrows can be interpreted as ow
lines of a uid entering a narrow and long channel from a point source located on
the top plate. To obtain the nematic texture the vectors need to be normalized and
viewed as rods with both ends identied as shown in Fig. 3.9.
Chapter 3: Liquid crystal textures in thick spherical shells. 90
1mm
Middle
Liquid
Inner Liquid
Outer
Liquid
0.05mm
Figure 3.9: Nematic texture in a thin spherical shell. The nematic director near the
pair of half hedgehogs indicated in the gure is well described by the one calculated
for the slab in Fig. 3.8.
Khln
_
R
h
_
of a disclination whose enlarged core of size h is given by the spherical
blob while the latter is roughly 4Kh, the energy of two half hedgehogs living inside
the blob [67]. In view of the azimuthal symmetry of our conguration, the director,
n(r, z), can be parameterized by the angle (r, z) formed with respect to the z axis
of the circular slab along the centers of the two half hedgehogs shown in Fig. 3.8
n(r, z) = sin (r, z)e
r
+ cos (r, z)e
z
. (3.50)
Chapter 3: Liquid crystal textures in thick spherical shells. 91
The energy density f() =
1
2
K
1
( n)
2
+
1
2
K
3
(n)
2
expressed in terms of the
bond angle reads
f() =
K
1
2
_
sin
r
+
r
cos
z
sin
_
2
+
K
3
2
(
z
cos +
r
sin )
2
. (3.51)
where K
1
and K
3
are the splay and bend constants and
r


r
and
z


z
. In
the one Frank constant approximation, minimization of the free energy leads to a non
linear partial dierential equation for the bond angle (r, z)
1
r

r
_
r

r
_
+

2

z
2
=
sin 2
r
2
. (3.52)
The operator on the left arise from the Laplacian in cylindrical coordinates.
3
.
Instead of solving this partial dierential equation, we follow a route anal-
ogous to that in Ref.[68], that is we construct an exact 2D solution for the liquid
crystal problem and then rotate it along the axis z to retrieve an ansatz for the 3D
director conguration
4
.
To solve the 2D problem we adopt the method of conformal mappings that
simplies the study of many complicated boundary problems in uid dynamics and
electrostatics [45]. Think of Fig. 3.8 as a source of uid conned to ow in a narrow
and long channel (R h). The spherical half-hedgehog corresponds to the source
and the hyperbolic one at the top to the stagnation point of the ow. The complex
3
Note there is no need to explicitly consider the azimuthal angle as an additional independent variable
in view of the symmetry of the problem.
4
Note that the 2D solution is the true minimum of the two dimensional liquid crystal elastic free energy
while the 3D Ansatz does not minimize f() nor satisfy Eq.(3.52)
Chapter 3: Liquid crystal textures in thick spherical shells. 92
potential (w) of the desired ow is
(w) = Log
_
sin
_
w
h
_
1
_
, (3.53)
where the complex variable is denoted by w = r + iz to distinguish it from the z
coordinate along the axis of the cylinder. The velocity eld is given by

(w) and
the corresponding complex nematic director n(w) = cos + i sin is obtained by
normalizing this vector eld. Note that the prime in

(w) denotes the derivative


with respect to w of the complex function = +i and we dene i. A
straightforward but tedious calculation (see Appendix B) leads the functional form
of (r, z) for a source at z =
h
2
and a stagnation point at z = +
h
2
, namely
tan (r, z) = sec
_
z
h
_
sinh
_
r
h
_
. (3.54)
This trial solution respects the boundary conditions of the problem, has the correct
short and long distance behavior in r and the expected functional form near the
source
5
.
Upon inserting (r, z) in Eq.(3.51) and integrating the energy density f()
over the cylindrical volume of the slab, we obtain the total elastic energy E
1
stored
in the eld
6
E
1
= Kh
_
ln
_
R
h
_
+c
_
, (3.55)
where c 4.2. Note that the functional dependence of E
1
on R and h matches
the expectations from the blob argument. Indeed the prefactor of the logarithm is
5
Similar problems were recently investigated in the context of complex magnetic patterns of ferromagnetic
nanostructures shaped as strips and rings. See O. Tchernyshyov and Gia-Wei Chern, cond-mat/0506744 and
references therein.
6
The resulting energy E1 does not rely on the assumption R h, as can be explicitly checked numerically.
Chapter 3: Liquid crystal textures in thick spherical shells. 93
universal in the sense that it does not depend on the details of the trial solution near
the hedgehogs, but only on its long distance behavior. By contrast, we expect the
result quoted above for the coecient c to be only an estimate (an upper bound)
since it relies on a trial solution for (r, z) that is not the absolute minimum of the
free energy. Numerical studies carried out in Ref.[66] for the slab geometry reported
that c 4.19.
A competing energy minimum for the nematic is given by a planar texture
with two s =
1
2
disclination lines. Note that +
1
2
lines cannot escape in the third
dimension [61]. A tedious but straightforward calculation shows that the energy E
p
of the pair of disclination lines is given by
E
p
=

2
Kh
_
ln
_
R
a
_
0.06 +
4 E
c
Kh
_
. (3.56)
where a is a macroscopic cuto typically of the order of the molecular length. These
two disclination repel each other, and are repelled by the circular boundary, leading
to a separation of order R. The second term of Eq.(3.56), corresponding to the
interaction of the two disclinations with the boundary and among themselves, is
negligibly small. The third term accounts for the core energies of the two disclination
lines 2E
c
. The combination Kh is equal to the two dimensional coupling constant
K
2D
. The relevant dimensionless ratio is
Ec
K
2D
.
By setting E
p
= E
1
we obtain the critical thickness h

above which the


escaped half-hedgehogs become energetically favored compared to a single s =
1
2
disclination line:
h

= e
c

Ra . (3.57)
Chapter 3: Liquid crystal textures in thick spherical shells. 94
The core energy terms in Eq.(3.56) reduce the previous estimate of c to c

= c
2Ec
Kh
.
A similar analysis applies to spherical shells which we now discuss.
3.4.2 Extrapolation to thin spherical shells
In principle, one could proceed along the same route as in the previous
section and nd a conformal mapping that provides a trial function for the bond
angle (, r)
7
corresponding to the texture shown in Fig. 3.9. This is possible but
rather cumbersome. In the case of very thin shells one can adapt the slab calculation
by noting again that the energy is composed of two parts. There is a long distance
piece arising from combing the hair of the nematic texture in the tangent plane of
the sphere that we can read o from a suitable 2D calculation (see Appendix A) and
the short distance contribution arising from the short distance contribution arising
from the two pairs of half-hedgehogs at the north and south pole.
The energy E
d
of two short +1 disclination lines placed at antipodal points
on a sphere (the north and south pole, say) can be estimated by performing a 2D
calculation on the curved surface and simply multiplying the result by the thickness
of the layer h
E
d
= 2Kh
_
ln
_
R
a
_
0.3 +
E
c
Kh
_
, (3.58)
where the middle term accounts for the interaction between the two disclinations in
their equilibrium positions. Note that this result is accurate only up to factors of
the order of
_
h
R
_
since the explicit integration over the volume of the thin shell was
7
In this case spherical coordinates are adopted for the independent variables.
Chapter 3: Liquid crystal textures in thick spherical shells. 95
bypassed. To obtain the energy of the escaped solution, the core size a in Eq.(3.58) is
rescaled to h. This will account for the integration of the energy density at distances
of the order of a few h R from the two hedgehogs
8
.
The energy stored in the remaining portions of the thin shell is approx-
imately given by twice the energy 4.2Kh of the yellow blob of Fig. 3.8. This
estimate neglects curvature corrections of the order of
_
h
R
_
and arises because at dis-
tances of the order h the spherical shell looks locally like a at circular slab as long
as h R. The resulting energy E
2
of the escaped conguration reads
E
2
= 2Kh
_
ln
_
R
h
_
0.3 + 4.2
_
. (3.59)
Although the prefactor of the sub-leading term linear in h has only been estimated, we
expect that the coecients of the logarithm, which arises from large scales compared
to h, is exact. For a spherical shell whose radius R is a hundred times its thickness,
the corrections from higher powers of
h
R
are indeed negligible. However for reasonable
values of
R
h
, the logarithmic term of Eq.(3.59) is still comparable in magnitude to the
sub-leading one linear in h.
The energy E
4
of the tetravalent conguration can be evaluated using similar
considerations, with the result
E
4
= Kh
_
ln
_
R
a
_
0.4 +
4E
c
Kh
_
. (3.60)
Upon setting E
4
= E
2
we obtain the critical thickness h

below which the tetravalent


8
In these portions of the shell the integrand reduces to the energy density of the two disclination problem
and hence the integration can be easily carried out leading the result in Eq.(3.58) with a lower cuto of the
order h.
Chapter 3: Liquid crystal textures in thick spherical shells. 96
conguration becomes energetically favored
h

= e
(4.1
2Ec
Kh
)

Ra . (3.61)
The exponential prefactor arises from the terms linear in h in Eq.(3.59), which cannot
be ignored in estimating h

even in the limit R h. Note that an accurate determi-


nation of the argument in the exponent would require knowledge of the core energies
of the disclination lines
9
.
The energy barrier between these two coexisting minima of the free energy
f() can be estimated by splitting the path connecting them in space in two steps.
First, consider a continuous deformation of the escaped texture of Fig. 3.9 obtained
by appropriately rotating each nematigen until the non escaped solution (with two
disclinations lines of index one at the north and south pole) is recovered. This part
of the path must be uphill in energy if the escaped solution was allowed to escape in
the rst place. The corresponding energy barrier E is given approximately by the
dierence between E
d
as calculated in Eq.(3.58) and E
2
in Eq.(3.59)
E = 2Kh
_
ln
_
h
a
_
4.2
_
. (3.62)
The second step consists in letting each of the unstable disclination lines
split in two +
1
2
defects and subsequently separate them until they sit at the vertexes
of a tetrahedron inscribed in the sphere. This portion of the path is downhill because
the non-escaped texture of valence 2 is unstable
10
. As a result the energy barrier
is simply the energy dierence calculated in Eq.(3.62).
9
In fact, the exponential prefactor can be interpreted as a numerically signicant rescaling of the core
radius.
10
This can be proved by writing down the energy of the pair and show that it decreases monotonically as
one separates them because of the electrostatic-like repulsion [12, 58].
Chapter 3: Liquid crystal textures in thick spherical shells. 97
Upon inserting K 10
6
dyn in Eq.(3.62) and taking k
B
T 4 10
14
erg,
as in Ref.[69, 67], we obtain
E
k
B
T

15h
nm
_
ln
_
h
a
_
4.2 +
E
c
Kh
_
. (3.63)
For shells with critical thickness h

Eq.(3.63) reduces to
E
k
B
T
10
3
_
R
a
ln
_
R
a
_
, (3.64)
where a core size of the order of 10 nm was assumed and the core energies were set
to zero [12]. This estimate indicates that the energy barrier is very high around h

suggesting that exchange between the two minima is unlikely to happen by thermal
activation. In a monodisperse solution of shells with thickness h, the ratio between
shells of the two valence will be given by their Boltzman factors as long as equilibrium
is reached. If one engineers shells with thickness below h

, the Z=4 conguration


would be more likely.
3.5 Conclusion
We have studied the crossover from the two dimensional regime of liquid
crystals conned on a spherical surface to the full three dimensional problem in a
spherical shell. For very thin shells, the nematic ground state has four disclination
lines sitting at the vertices of a tetrahedron inscribed in the ball and whose texture
approximately track the seam of a baseball. As the thickness increases, a competing
three dimensional defected texture characterized by two pairs of half hedgehogs at
the north and south pole becomes energetically favorable. For ultra-thin shells this
Chapter 3: Liquid crystal textures in thick spherical shells. 98
instability is suppressed and one expects a defected ground state with tetravalent
symmetry. Estimates of the stability of this texture to thermal uctuations indicate
that the vibrations around the equilibrium congurations of the defects should not be
signicant. The present analysis has been carried out primarily in the limit in which
the elastic anisotropy parameter, =
K
3
K
1
K
3
+K
1
1.
We hope to extend our investigation with a systematic study of the eect of
elastic anisotropy on the nematic texture. It is interesting to note that in the case of
pure bend or splay (ie. = 1) the ground state is given by only two disclinations of
unit index at the north and south pole. This suggests the possibility that the eect
of the elastic anisotropy may not be limited to locally adjusting the orientation of the
director but may induce a change in the inter-defect interaction and hence a distortion
of the tetravalent equilibrium conguration. The limit of strong elastic anisotropy is
also relevant to studies of the nematic to smectic transition in a spherical geometry
for which the ratio of the bend to splay coupling constants
K
3
K
1
is expected to diverge.
Additional experimental complications include the possibility of having a
nematic layer of non constant thickness that would induce trapping of the defects in
the regions where the layer is thinner. This eect may also induce a local transition
to an escaped texture where the layer thickens in just one hemisphere so that two
disclination lines of index
1
2
are traded for a pair of half hedgehogs. If that happens
shells with three-fold symmetry could be observed.
Chapter 4
Superuid lms on a curved
surface.
Based on V. Vitelli and A. M. Turner PRL 93, 215301(2004).
99
Chapter 4: Superuid lms on a curved surface. 100
The physics of topological defects on curved surfaces plays an increasingly
signicant role in the engineering of devices based on coated interfaces. [52, 53,
3]. Defects also aect the mechanical properties of some biological systems, such as
spherical viruses, whose shape is dependent on the presence of disclinations in their
protein shell [70]. Furthermore, the eects induced by a curved substrate on the
distribution of defects are not fully understood even in well studied systems such as
thin superuid or superconducting lms. In this paper, we study simple continuum
generalizations of the plane XY model to frozen surfaces of varying curvature to gain
a broad understanding of the interaction between topological defects and curvature.
The XY model is a simple setting in which particle-like objects emerge from
a more fundamental theory. The basic degree of freedom is an angle-valued function
on the plane whose values vary from 0 to 2. These angles could represent the
orientations of interacting arrows. The interaction, which tends to align neighboring
arrows, results from the continuum free energy F given by
F =
K
2
_
d
2
u ( (u))
2
, (4.1)
where the set of coordinates u = (x, y) label points on the plane. Despite its simplicity,
this model captures the main properties of vortices in layers of superuid
4
He or thin
superconducting lms when the eld (u) is identied with the phase of the collective
wave function. In addition, the elastic energy of Eq.(4.1) correctly describes liquid
crystalline phases for which the bond angle, (u), has periodicity
2
p
with p 3.
For a solution of nematigens (p = 2) and tilted molecules in a a Langmuir lm
(p = 1) two dierent elastic constants are necessary to account for bend and splay
Chapter 4: Superuid lms on a curved surface. 101
deformations [38], but these are renormalized to the same value at nite temperatures
[71]. Besides its experimental signicance, the XY model is the cornerstone of our
conceptual understanding of topological defects, singular congurations of the eld
(u).
Like particles, defects have charges and a characteristic Coulomb-like inter-
action. The charge q, a multiple of
2
p
, can be dened by the amount increases along
a path enclosing the defects core. The force between two defects located at positions
u
i
and u
j
is determined by the energy stored in the eld, Kq
i
q
j
U(u
i
, u
j
), where
the inter-particle potential U(u
i
, u
j
) is proportional to the logarithm of the distance
in the plane. On a at surface, thin layers of superuids, superconductors and liquid
crystals can all be analyzed within the framework of Eq.(4.1) [72]. However, there
is a crucial dierence between, say, the phase of the superuid order parameter and
the angle that describes the local orientation of liquid crystal molecules. The for-
mer transforms like a scalar since the quantum mechanical phase does not change
when the system is rotated, while the latter represents a vector aligned to the local
direction of the molecules. Thus, a common boundary condition for a liquid crystal
is for the director to be tangent to the boundary of the substrate. By contrast, no
such constraint exists for a
4
He lm because its wave function is dened in a dierent
space from the one in which the superuid is conned. This distinction is crucial
on a curved surface. In the ground state of a
4
He lm, the phase (u) is constant
throughout the surface and the corresponding energy vanishes. The free energy F
s
to
Chapter 4: Superuid lms on a curved surface. 102
be minimized is a scalar generalization of Eq.(4.1):
F
s
=
K
2
_
d
2
u

g g

(u)

(u) . (4.2)
Here the set of coordinates u = (u
1
, u
2
) label points on the surface while

g is the
determinant of the metric tensor g

. On the other hand, a constant bond angle (u)


is not the ground state of the liquid crystal because it is measured with respect to
an arbitrary basis vector E

(u) with = 1, 2. Indeed, it is not possible to make the


directions of the molecules parallel everywhere on a curved space; the lowest energy
state is attained by optimally distributing the unavoidable bend and splay of the
vectors over the whole surface. The free energy functional F
v
to be minimized is a
vector generalization of Eq.(4.1) [40]:
F
v
=
K
2
_
d
2
u

gg

(u)

(u))(

(u)

(u)) , (4.3)
where

(u), the connection, compensates for the rotation of the 2D basis vectors
E

(u) in the direction of u

. Since the curl of

(u) is equal to G(u) [33], the


integrand in Eq.(4.3) cannot be made to vanish on a surface with non-zero Gaussian
curvature
1
. As the substrate becomes more curved, the energy cost of this geometric
frustration can be lowered by generating defects in the ground state even in the
absence of topological constraints [73, 74].
In this letter, we introduce a novel coupling between a defect and the varying
curvature of the substrate which originates in a conformal anomaly of the free energies
of Equations (4.2) and (4.3). This anomaly arises, even at zero temperature, from
1
(u) is a non-conservative eld and hence cannot be equal to everywhere.
Chapter 4: Superuid lms on a curved surface. 103
imposing a constant cuto, a, localized at the core of each defect
2
. By contrast,
nite temperature conformal anomalies [46] are generated by the presence of a short
wavelength cuto for the uctuations in (u) at every point on the surface. A physical
consequence of the anomalous coupling is that topological defects in superuids and
superconductors interact with the curvature in a radically dierent way from the case
of liquid crystal order
3
.
For thin layers of superuids and superconductors, we prove that the geo-
metric interaction E
s
(u
i
) is given by:
E
s
(u
i
) =
K
4
q
2
i
V (u
i
) , (4.4)
where u
i
and q
i
are respectively the position and topological charge of the defect. The
geometric potential V (u) satises a covariant version of Poissons equation where the
negative of the Gaussian curvature G(u) plays the role of the charge density:
D

V (u) = G(u) . (4.5)


For an azimuthally symmetric surface such as the bump represented in Fig. 4.1,
we can explicitly obtain V (u), as a function of the radial distance from the top, by
employing a two dimensional analogue of Gauss law [74]. The resulting potential well
V (r) vanishes at innity and its width and depth are given respectively by the linear
size of the bump and its aspect ratio squared. Eq.(C.7) has an elegant geometrical
interpretation if a set of coordinates is chosen so that the metric tensor is cast in
2
We assume that a is much smaller than the radius of curvature and hence independent of the defect
position.
3
In both cases, the electrostatic interaction between the defects U(ui, uj) is given by the Greens function
of the covariant Laplacian that is not translationally invariant and depends on the shape of the surface [74].
Chapter 4: Superuid lms on a curved surface. 104
the form g

exp (2(u)) [40]. The conformal factor (u) is controlled by


the overall shape of the surface and it satises the same Poisson Eq.(C.7) as the
geometric potential [40]. We therefore proceed with the identication of V (u) with
(u)
4
. This observation will be the basis of our proof of Eq.(4.4) which results in
the novel prediction that a vortex in a superuid or superconducting lm is repelled
(attracted) by positive (negative) Gaussian curvature irrespective of its charge and
sign.
For liquid crystals, the geometric interaction E
v
(u
i
) contains an additional
term discussed in previous investigations of hexatic membranes [35], which arises
from the geometric frustration of the vector eld. This term, linear in q, happens to
contain the same function V (u) as the conformal anomaly of Eq.(4.4). When both
contributions are included, E
v
(u
i
) acquires an unexpected dependence on the charge
of the defect:
E
v
(u
i
) = Kq
i
_
1
q
i
4
_
V (u
i
) . (4.6)
The interpretation of V (u) as a geometric potential and the linear dependence on q in
the rst term of Eq.(4.6) are consistent with the general belief that a defect interacts
logarithmically with the Gaussian curvature, as an electrostatic particle would with
a background charge distribution. However, E
v
(u
i
) does not grow linearly with the
charge of the defect, as expected from the electrostatic analogy. Instead, the geometric
interaction peaks for a defect of charge 2 and eventually becomes repulsive for q
greater than the critical charge q
c
= 4
5
.
4
This identication is possible for corrugated planes at at innity. The details of the analysis valid for
deformed spheres and torii will be presented elsewhere.
5
The symmetry around q = 2 is reected in the pattern of eld lines around a defect. There are |
q2

|
Chapter 4: Superuid lms on a curved surface. 105
Figure 4.1: A corrugated substrate and its downward projection on a at plane. The
shaded strip surrounding P is more stretched than the one surrounding Q despite
their projections onto the plane having the same area. The energy stored in the eld
will be lower if the core of the defect is located at Q rather than P.
The quadratic coupling has an intuitive explanation in the case of azimuthally
symmetric surfaces. Consider a very thin superuid lm deposited on the surface il-
lustrated in Fig. 4.1 with a vortex of charge q placed on top of the bump. In order
to calculate the energy stored in the eld, we only need to know that the superuid
phase (u) changes uniformly by q along a circumference of length 2r centered on
the defect. Inspection of Eq.(4.2) reveals that the energy density of the eld in the
shaded strip at distance r is proportional to
_
q
r
_
2
, where r is the distance to the sin-
gularity measured in the plane of projection (see Fig. 4.1). By vertically stretching
the surface, the amount of area in the shaded strip is increased with respect to its
lobes.
Chapter 4: Superuid lms on a curved surface. 106
projection on the plane, while the energy density is unchanged. As a result, the total
energy stored in the eld is greater when a vortex sits on top of a bumpy surface than
when the same vortex is located at the center of a at disk of the same area. Hence, it
is energetically favorable for the vortex to migrate to the at portions of the surface.
In this case, the vortex is far away from the bump so that the total energy stored
in the eld does not dier much from the at plane result [75]. For less symmetric
surfaces, the resulting geometric interaction will depend on the shape of the entire
surface as embedded in the metric tensor.
The physical origin of the linear coupling between defects and curvature in
Eq.(4.6) is illustrated in Fig. 4.2 for a disclination of charge 2 centered on a bump.
As the curvature of the bump is increased, the bend or splay of the director of the
liquid crystal decreases and hence the energy stored in the vector eld is reduced.
As a result, this linear coupling causes positively (negatively) charged defects to be
attracted by positive (negative) Gaussian curvature [35]. However, this mechanism
competes with the repulsive geometric interaction illustrated in Fig. 4.1 that is at
work also in the case of liquid crystal order. We note that the linear coupling is absent
for superuids because, in Eq.(4.2),

(u) is not coupled to a curvature dependent


connection,

(u), as it is in Eq.(4.3). The critical value q


c
= 4, where the single
defect potential E
v
(u
i
) changes sign, can be determined from simple geometrical ar-
guments. Consider an isolated disclination of charge q on a hemispherical cup placed
on a at plane. On account of azimuthal symmetry, the force acting on the defect
depends only on the net Gaussian curvature enclosed by the circle on which it is
Chapter 4: Superuid lms on a curved surface. 107
Figure 4.2: Disclinations of charge 2 located on top of bumps with dierent aspect
ratios. The amount of splay in the liquid crystal director on the taller bump is reduced
and hence the energy density is lower.
placed, see Fig. 4.3a [74]. This interaction is unchanged if we deform the outer region
of the plane and eventually compactify it to form a sphere as illustrated in Fig. 4.3b.
In order to satisfy topological constraints
6
, we still need to place a shadow defect of
charge (4 q) at the south pole (the only position outside the circle that does not
destroy the azimuthal symmetry of the initial problem). The curvature-defect interac-
tion on the hemisphere is thus reduced to the well known defect-defect interaction on
the sphere [12]. The latter is proportional to q(4 q) and so is the curvature-defect
interaction on the deformed plane of Fig. 4.3a, in agreement with Eq.(4.6). This
provides evidence that a disclination of charge greater than 4 will be repelled from
6
The sum of the defects charges on a surface of genus g is equal to 4(1 g) for a vector eld and 0 for
a scalar [40].
Chapter 4: Superuid lms on a curved surface. 108
(a)
(b)
Figure 4.3: (a) An isolated disclination on a deformed plane feels a force that depends
only on the enclosed Gaussian curvature. (b) The deformed plane is compactied to
the sphere by placing a shadow defect at the south pole.
regions of positive curvature. We now present a derivation of the coupling between
curvature and defects in helium and superconducting lms that employs the method
of conformal mapping, often adopted in electromagnetism and uid mechanics to sim-
plify the boundary of complicated planar regions. In this context, we use conformal
mappings to relate the complex task of nding the eld energy on an arbitrarily de-
formed target surface to an equivalent problem on a homogeneous reference surface
(see Fig. 4.4). A conformal mapping has two equivalent dening properties: angles
map to equal angles, and very small gures map to gures of nearly the same shape.
One can always nd a conformal mapping from the target to the reference surface
[40] such that g
T
= e
2(u)
g
R
, where g
T
and g
R
are the metric tensors on the target
Chapter 4: Superuid lms on a curved surface. 109
and reference surfaces respectively. The scaling factor e
(u)
varies with the position u
on the target surface, so that larger gures are inhomogeneously distorted when they
are mapped from the target to the reference surface. We choose the reference surfaces
to be undeformed and of the same topology as the target spaces (eg., g
R
=

for
a corrugated plane). Defects on the target surface are mapped onto a set of image
defects on the reference surface.
The crucial property of the scalar free energy F
s
is its invariance under the
rescaling of the metric by the conformal factor
7
. However, the conformal symmetry of
F
s
is broken upon introducing a short distance cuto a that is necessary to prevent the
energy from diverging in the core of the defect. Because of the varying scaling factor,
the constant physical core radius a is stretched or contracted when projected on the
reference space by an amount dependent on the position of the defect (see Fig. 4.4).
The radius of the image of the i
th
core is a
i
= e
(u
i
)
a. It is this conformal anomaly that
is responsible for generating the geometric interaction in Eq.(4.4). In fact, the energy
of the defects in the target space E
T
is equal to the energy of a conguration of defects
(on the reference surface) whose core radii are position-dependent. This problem can
be further transformed into the simpler task of nding the energy E
R
for a set of
interacting defects with constant core radius a plus an eective geometric potential
that accounts for the variation of the core size with position. This geometric potential
can be derived with the aid of Fig. 4.4. If a
i
is smaller (larger) than a, the energies
stored in the annular regions indicated in Fig. 4.4 need to be added (subtracted)
7
When gT is substituted in Eq.(4.2) the conformal factor e
2(u)
cancels out and

gT g

T
=

gR g

R
[74].
Chapter 4: Superuid lms on a curved surface. 110
Figure 4.4: Conformal mapping of the target surface T onto the reference space
R. The continuous disks on both surfaces represent the physical cores of constant
radius a. The dashed lines represent the position dependent images on R of the defect
cores on T with variable radii a
i
. Note that the energy stored in the annuli comprised
by the dashed and continuous lines in R must be added or subtracted to E
R
to obtain
E
T
.
from E
R
to obtain E
T
. To calculate this extra energy, we introduce a set of polar
coordinates (r,) centered on the i
th
defect. Near the defect of charge q
i
, the phase
is given by
q
i
2
and the energy density is
Kq
2
i
8
2
r
2
. Upon integrating it over the
annulus comprised between a and a
i
= e
(u
i
)
a (see Fig. 4.2), we obtain
E
T
E
R
= K
N
d

i=1
q
2
i
4
(u
i
) , (4.7)
where N
d
is the number of defects. The energy E
R
accounts for defect-defect inter-
actions since any potential felt by a single defect would have to be constant because
all points are equivalent on the reference surface (undeformed sphere or plane). Re-
calling that (u) = V (u), we recover the result of Eq.(4.4) with no dependence on
Chapter 4: Superuid lms on a curved surface. 111
the microscopic physics because the core size a drops out in Eq.(4.7)
8
. In the case
of liquid crystal order, the contribution of the anomaly is simply added to the term
linear in q as indicated in Eq.(4.6) [74].
Experiments that test our predictions can be realized by coating a bump
with a thin layer of superuid helium and rotating it around its axis of symmetry so
that a single vortex forms [76]. The competition between the (repulsive) geometric
interaction and the conning parabolic potential (generated by the rotation) would
cause the equilibrium position of the vortex to shift from the center of the bump if
its height exceeds a critical value. The vortex line could be detected by trapping of
electrons on its core [77]. Other experiments may detect an inhomogeneous distribu-
tion of thermally induced defects resulting from the combined eect of the anomalous
coupling and the dependence of their Coulomb-like interaction on the varying curva-
ture.
We have demonstrated that the interaction between defects and curvature in
2D XY-like models depends crucially on the nature of the underlying order parameter
and we have shown how to explicitly derive the resulting geometric force from the
shape of the surface.
8
Additional couplings resulting from other terms in a Landau expansion introduce negligible corrections
proportional to the inverse of the area of the surface.
Chapter 5
Defects and crystalline order on
curved surfaces.
Based on V. Vitelli, J. B. Lucks and D. R. Nelson, cond-mat/0604203
112
Chapter 5: Defects and crystalline order on curved surfaces. 113
5.1 Introduction
The physics of two dimensional crystals on curved substrates is emerging as
an intriguing route to the engineering of self assembled systems such as the colloi-
dosome, a colloidal armor used for drug delivery [3], or devices based on ordered
arrays of block copolymers which are a promising tool for soft lithography [54, 78].
Curved crystalline order also aects the mechanical properties of biological structures
like clathrin-coated pits [5, 6] or HIV viral capsids [8, 9] whose irregular shapes appear
to induce a non-uniform distribution of disclinations in their shell [79].
In this chapter, we present a theoretical and numerical study of point-like
defects in asoft crystalline monolayer grown on a rigid substrate of varying Gaus-
sian curvature with lattice constant of order, say, 10 nm or more. The substrate can
then be assumed smooth on the scale of the monolayer lattice constant, as would be
the case for di-block copolymers [54, 78]. Disclinations and dislocations are important
topological defects that induce long range disruptions of orientational or translational
order respectively [2]. Disclinations are points of local 5- and 7-fold symmetry in a
triangular lattice (labeled by topological charges q =
2
6
respectively), while dislo-
cations are disclination dipoles characterized by a Burgers vector,

b, dened as the
amount by which a circuit drawn around the dislocation fails to close (see Figure
5.1 inset). Other point defects such as vacancies, interstitials or impurity atoms cre-
ate shorter range disturbances that introduce only local stretching or compression
in the lattice (see Figure 5.4). Such defects are important for particle diusion and
relaxation of concentration uctuations.
Chapter 5: Defects and crystalline order on curved surfaces. 114
These particle-like objects interact not only with each other, but also with
the curvature of the substrate via a one-body geometric potential that depends on the
particular type of defect [15, 4]. These geometric potentials are in general nonlocal
functions of the Gaussian curvature that we determine explicitly here for a model
surface shaped as a Gaussian bump. An isolated bump of this kind models long
wavelength undulations of a lithographic substrate, has regions of both positive and
negative curvature, and yet is simple enough to allow straightforward analytic and
numerical calculations. The presence of these geometric potentials triggers defect
unbinding instabilities in the ground state of the curved space crystal, even if no
topological constraints on the net number of defects exist. Geometric potentials also
control the dynamics of isolated dislocations whose motion in the glide direction
can be suppressed. Similar mechanisms inuence the equilibrium distribution and
dynamics of vacancies, interstitials and impurity atoms.
5.2 Basic Formalism
The in-plane elastic free energy of a crystal embedded in a gently curved
frozen substrate,given in the Monge form by r(x, y) = (x, y, h(x, y)), can be expressed
in terms of the Lame coecients and [14]
F =
1
2
_
dA
ij
(x)u
ij
(x) =
_
dA
_
u
2
ij
(x) +

2
u
2
kk
(x)
_
, (5.1)
where x = {x, y} represents a set of standard cartesian coordinates in the plane
and dA = dxdy

g, where

g =
_
1 + (
x
h)
2
+ (
y
h)
2
. In Eq.(5.1) the stress ten-
sor
ij
(x) = 2u
ij
(x) +
ij
u
kk
(x) was written in terms of the strain tensor u
ij
(x) =
Chapter 5: Defects and crystalline order on curved surfaces. 115
1
2
[
i
u
j
(x) +
j
u
i
(x) +A
ij
(x)]. The latter contains an additional termA
ij
(x) =
i
h(x)
j
h(x)
(compared to its at space counterpart) that couples the gradient of the displacement
eld u
i
(x) to the geometry of the substrate as embedded in the gradient of the sub-
strate height function h(x). We will often illustrate our results for a single Gaussian
bump whose height function is given by h(x) = x
0
e

r
2
2
where r
|x|
x
0
is the dimen-
sionless radial coordinate [58].
In Eq.(5.1) and the rest of this paper, we adopt the ordinary at space metric
(eg. setting dA dxdy) and absorb all the complications associated with the curved
substrate in the tensor eld A
ij
(x) that resembles the more familiar vector potential of
electromagnetism. Like its electromagnetic analog, the curl of the tensor eld A
ij
(x)
has a clear physical meaning and is equal to the Gaussian curvature of the surface
G(x) =
il

jk

k
A
ij
(x) where
ij
is the antisymmetric unit tensor (
xy
=
yx
= 1)
[15]. The consistency of our perturbative formalism to leading order in is assessed
in Appendix H.
Minimization of the free energy in Eq.(5.1) with respect to the displacements
u
i
(x), naturally leads to the force-balance equation,
i

ij
(x) = 0. If we write
ij
(x)
in terms of the Airy stress function (x)

ij
(x) =
il

jk

k
(x) (5.2)
then the force balance equation is automatically satised,

ij
=
jk

k
[
1
,
2
](x) = 0 , (5.3)
since the commutator of partial derivatives is zero. Any scalar function (x) will
generate a stress tensor eld
ij
(x) that satises the force balance equation. To lift
Chapter 5: Defects and crystalline order on curved surfaces. 116
this redundancy, one choses (x) and hence
ij
(x) so that the corresponding strain
tensor, u
ij
(x), matches boundary conditions for the problem under investigation.
The free energy in Eq.(5.1) can be recast (up to boundary terms) as a simple
functional of the scalar eld (x)
F =
1
2Y
_
dA ((x))
2
, (5.4)
where is the at space Laplacian and Y =
4(+)
2+
is the Young modulus [2]. Upon
minimizing F in the presence of defects, we obtain a bi-harmonic equation for (x)
whose source is controlled by the distribution of defects and by the varying Gaussian
curvature of the surface G(x) [2, 15]
1
Y

2
(x) = S(x) G(x) . (5.5)
The source S(x) for a distribution of N unbound disclinations with topological
charges {q

=
2
6
} and M dislocations with Burgers vectors {

} reads [2]
S(x) =
N

=1
q

(x, x

) +
M

=1

ij
b

i

j
(x, x

) , (5.6)
where |

b| is equal to the lattice constant a for dislocations with the smallest Burgers
vector. For N isotropic vacancies, interstitials or impurities, we have [2]
S(x) =
1
2
N

=1

(x, x

) , (5.7)
where

a
2
is the local area change caused by including the point defect at position
x

.
It is convenient to introduce an auxiliary function V (x) that satises the
Poisson equation V (x) = G(x) and vanishes at innity where the surface attens
Chapter 5: Defects and crystalline order on curved surfaces. 117
out. In order to determine the geometric potential, (x

), of a defect at

x

we
integrate by parts twice in Eq.(C.2) and use Eq.(5.5) to obtain
2
(x) and (x) in
terms of the Greens function of the biharmonic operator. The geometric potential
(on a deformed plane at at innity) follows from integrating by parts the cross tems
involving the source and the Gaussian curvature with the result
(x

) = Y
_
dA

S(x

)
_
dA
1

xx

V (x) . (5.8)
This formula needs to be supplemented with appropriate nite size corrections to
account for the presence of a boundary, as discussed in Appendix I.
5.3 Geometric Frustration
We start by calculating the energy of a relaxed defect-free two dimensional
crystal on a quenched topography. In analogy with the bending of thin plates we
expect some stretching to arise as an unavoidable consequence of the geometric con-
straints associated with the Gaussian curvature [14]. The resulting energy of geomet-
ric frustration, F
0
, can be estimated with the aid of Eq.(5.5) which, when S(x) = 0,
reduces to a Poisson equation whose source is given by V (x)
1
Y

G
(x) = V (x) +H
R
(x) . (5.9)
where H
R
(x) is an harmonic function of x parameterized by the radius of the circular
boundary R. As discussed in Appendix I, H
R
(x) vanishes in the limit R x
0
if free
boundary conditions are chosen. We denote the solution of Eq.(5.9) as
G
(x) where
the superscript G indicates that the Airy function and the corresponding stress tensor
Chapter 5: Defects and crystalline order on curved surfaces. 118

G
ij
(x) describe elastic deformations caused by the Gaussian curvature G(x) only,
without contribution from the defects. Upon using the general denition of the Airy
function (Eq. 5.2), we obtain

G
kk
(x) =
G
(x) = Y V (x) . (5.10)
For a surface with azimuthal symmetry, like the bump, Poissons equation
can be readily solved upon applying Gauss theorem with the Gaussian curvature G(r)
as a source [58]:
V (r) =
_

r
dr

r
_
r

0
dr

G(r

) =
1
4

2
e
r
2
.
Upon substituting Eq.(5.9) in Eq.(C.2), we have
F
0

Y
2
_
dA V (r)
2
=
Y x
2
0

4
64
.
The result in Eq.(5.3) is valid to leading order in , consistent with the assumptions of
our formalism after taking the limit R x
0
(see Appendix B for nite size corrections
in small systems). For a harmonic lattice, Y =
2

3
k, where k is an eective spring
constant that can be extracted from more realistic inter-particle potentials [80]. For
colloidal particles, k is typically of the order of a few hundred times k
B
T/a
2
, where
T is room temperature [10]. Our numerical calculations of F
0
in xed-connectivity
harmonic solids are in good agreement with the small expansion in Eq.(5.3) as long
as the aspect ratio is around 1/2 or lower (see Appendix H for a numerical plot
of F
0
versus ). An immediate implication of the geometric frustration embodied in
Eq. (5.3) and (5.3) is that nucleation of crystal domains on the bump will take place
preferentially away from the top in regions where the surface attens out.
Chapter 5: Defects and crystalline order on curved surfaces. 119
5.4 Geometric potential for dislocations
The energy of a two dimensional curved crystal with defects will include the
frustration energy, the inter-defect interactions (to leading order these are unchanged
from their at space form, see Appendix H), possible core energies and a character-
istic, one-body potential of purely geometrical origin that describes the coupling of
the defects to the curvature given by Eq.(5.8). The geometric potential of an isolated
dislocation, (x) D(x, ), is a function of its position as well as of the angle that
the Burger vector

b forms with respect to the radial direction (in the tangent plane
of the surface). Upon setting all q

= 0 in Eq.(5.6) and substituting it into Eq.(5.8),


we obtain, for an isolated dislocation, the resulting function D(r =
|x|
x
0
, )
D(r, ) = Y b
i

ij

j
_
dA

_
V (

) +

2
x
2
0
4R
2
_
Y b x
0

2
8
sin
__
e
r
2
1
r
_
+
_
x
0
R
_
2
r
_
. (5.11)
In view of the azimuthal symmetry of the surface, Gauss theorem as expressed in
Eq.(5.3), was employed in deriving the second equality in Eq.(5.11) which is function
only of the dimensionless radial coordinate r. The rst term in Eq.(5.11) corresponds
to the innite plane geometric potential obtained from Eq.(5.8) while the second term
is a nite size correction arising from a circular boundary of radius R (see Appendix
I for a detailed derivation). Eq.(5.11) is valid to leading order in perturbation theory,
consistent with the small approximation adopted in this work (see Appendix H).
In Fig. 5.1 we present a detailed comparison between the theoretical pre-
dictions for the geometric potential
D(r,)
Y bx
0
plotted versus r as continuous lines and
Chapter 5: Defects and crystalline order on curved surfaces. 120
numerical data from constrained minimization of an harmonic solid on a bump with
= 0.5, under conditions such that R x
0
a. (See Appendix J for a discussion
of our numerical approach). The lower and upper branches of the graph are obtained
from Eq. (5.11) by setting =

2
and letting
R
x
0
equal to 4 (blue curve) or 8 (red
curve). Indeed, the (scaled) data from simulations with dierent choices of
x
0
b
(see
caption) collapse on the two master-curves according to their ratio of
R
x
0
. Two dif-
ferent curves arise because the dislocation interacts with the curvature directly and
via its image. The image-mediated interaction is given by the R dependent term in
Eq.(5.11). The simple dependence of D(r, ) on the direction of the Burgers vector is
revealed by Fig. 5.1, since the upper branch of the graph corresponding to the unsta-
ble equilibrium =

2
is approximately symmetric to the lower one corresponding
to =

2
.
The analogy between the geometrical potential of the dislocation and the
more familiar interaction of an electric dipole in an external eld can be elucidated
by regarding the dislocation as a charge neutral pair of disclinations whose dipole
moment qd
i
=
ij
b
j
is a lattice vector perpendicular to

b that connects the two
points of 5 and 7-fold symmetry. The geometric potential, U(r), of a disclination
of topological charge q interacting with the Gaussian curvature satises the Poisson
equation U(r) = qV (r) as can be seen by substituting the source in Eq.(5.6) with
all

b

= 0 into Eq.(5.8) . For small r, positive (negative) disclinations are attracted


(repelled) from the center of the bump by the integrated background source V (r)
which increases for r 1 like
2
r
2
and is multiplied by the 2D electric eld
1
r
,
Chapter 5: Defects and crystalline order on curved surfaces. 121
0 0.5 1 1.5 2 2.5 3 3.5
0.02
0.015
0.01
0.005
0
0.005
0.01
0.015
0.02
r
D(r,)
Ybr
0
Figure 5.1: The dislocation potential D(x, =

2
) in Eq.(5.11) (including nite size
corrections) is plotted as a continuous line for a Gaussian bump parameterized by =
0.5 in the limit R >> x
0
>> a. Open symbols represent the numerical minimization
of a xed connectivity harmonic model for which the separation of the +/- disclination
pair representing the dislocation is xed while allowing the dislocation as a whole to
move radially with respect to the bump. Lower branch: = /2, R/x
0
=4 (blue),
8 (red); and x
0
/a = 10 (_), 20 (), 40 (2). Upper branch: = /2, x
0
/R = 8,
x
0
/a = 10. Inset: A schematic view of a dislocation with lled symbols representing
six- (circles), ve- (diamond) and seven- (square) coordinate particles. Also depicted
are the two rows of extra atoms emanating from the ve-coordinated particle. The
Burgers vector is shown as a red arrow, completing the circuit around the dislocation
(dashed line).
Chapter 5: Defects and crystalline order on curved surfaces. 122
resulting in a geometric force that increases linearly in r. If the positive disclination
within the dipole is closer to the top, the force it experiences will be opposite and
slightly less than the one experienced by the negative disclination that is further away
from the top. As a result a tidal force will push the dislocation (as a whole) down
hill as shown by the lower branch of the plot of Fig. 5.1. For large r, however, the
source V (r) saturates and the attractive force exerted on the positive disclination
wins and drags the dislocation towards the bump. The minimum of the geometric
potential occurs at |x
min
| 1.1 x
0
, close to the circle of zero Gaussian curvature
|x| = x
0
, as a result of the competing interactions of the two disclinations comprising
the dislocation. If the orientation of the disclination dipole is ipped, the geometric
potential ips sign. Similarly, if the sign of the Gaussian curvature is reversed, that
is to say the bump turns into a saddle, the sign of the geometric interaction ips. As
a result, a dislocation close to a saddle will have its Burger vector oriented so that
the closest disclination to the saddle is 7-fold coordinated.
The physical origin of the dislocation potential can be understood heuris-
tically without explicit recourse to our source formalism. According to standard
elasticity theory, a dislocation in an external stress eld
ij
(x) experiences a Peach
Kohler force,

f(x), given by f
k
(x) =
kj
b
i

ij
(x) [47]. Similarly, a dislocation intro-
duced into the curved 2D crystal will experience a Peach Kohler force as a result of
the pre-existing stress eld of geometric frustration
G
ij
(

) whose non-diagonal com-


ponents vanish. This interpretation is consistent with the geometric potential derived
in Eq.(5.11), provided we use Eq. (5.9) to write D(x) = b
i

ij

G
. With

b along its
Chapter 5: Defects and crystalline order on curved surfaces. 123
minimum orientation (azimuthal counter-clockwise), we obtain a radial Peach Kohler
force of magnitude f(r) = b
G

(r) that matches


D(r)
r
= b

2

G
(r)
r
2
.
5.5 Dislocation unbinding and Grain Boundaries
If the 2D crystal is grown on a substrate which is suciently deformed, the
resulting elastic strain can be partially relaxed by introducing unbound dislocations
into the ground state [15, 58]. Here we present a simple estimate of the threshold
aspect ratio,
c
, necessary to trigger this instability. Boundary eects will be ignored
in what follows by letting R in Eq.(5.11).
Consider two dislocations located at x
1
and x
2
a distance of approximately
x
0
from the center of the bump on opposite sides (see the inset of Fig. 5.2). Their
disclination-dipole moments are opposite to each other and aligned in the radial direc-
tion so that the two (antiparallel) Burger vectors are perpendicular to the separation
vector in the plane x
12
x
1
x
2
. In this case, the interaction between the disloca-
tions reduces to V
12

Y
4
b
2
ln
_
|x
12
|
a
_
[2]. The instability occurs when the energy gain
from placing each dislocation in the minima of the potential D(x) given by Eq.(5.11)
outweights the sum of the work needed to tear them apart plus the core energies 2E
c
.
The critical aspect ratio at the threshold which results is given by

2
c
c
b
x
0
ln
_
x
0
b

_
, (5.12)
where b


b
2
e

8Ec
Y b
2
is the magnitude of the rescaled Burger vector and c 1/2.
In Fig. 5.2a we present a comparison between Eq. (5.12) and numerical re-
sults. For each value of
x
0
b
, the corresponding
c
is obtained numerically by comparing
Chapter 5: Defects and crystalline order on curved surfaces. 124
10 12 14 16 18 20 22 24
0
0.1
0.2
0.3
0.4
0.5
0.6
r
0
/b

C
(a)
(b)
(c)
Figure 5.2: (a) Dislocation unbinding critical aspect ratio,
c
, as a function of
r
0
b
.
The theoretical estimate (5.12) is plotted vs. x
0
for core energies E
c
= 0 (dashed)
and E
c
= 0.1Y b
2
(solid). Numerical values (circles) were obtained as described in
Section 5.5 and Appendix J. Inset: particle conguration projected in the plane
for a dislocation pair straddling the bump. The extra atoms associated with the
dislocations are highlighted with black lines. (b) Log of the numerical strain energy
density on a bump for (b) the conguration shown in the inset. (c) The same quantity
for a defect-free bump. Both plots were constructed numerically with x
0
= 10, =
0.7 >
c
. Red represents high strain.
the energy of a lattice without defects to the conguration with the two dislocations
in their equilibrium positions. This interpretation for the origin of the instability is
corroborated by the (numerical) strain energy density plots of Fig. 5.2b and 5.2c ,
where it is shown that introducing the pair of dislocations reduces the strain energy
density on top of the bump at the price of creating some large, but localized strains
around the dislocation cores where u
ij
diverges. In the continuum limit b x
0
,
very small deformations are enough to trigger the instability. This is the regime in
which our perturbation treatment applies. As is increased even further a cascade of
dislocation unbinding transitions occurs involving larger numbers of dislocations and
more complicated equilibrium arrangements of zero net Burger vector. For suciently
Chapter 5: Defects and crystalline order on curved surfaces. 125
large aspect ratios, we expect that the dislocations tend to line up in grain boundary
scars similar to the ones observed in spherical crystals [3]. This scenario is consistent
with preliminary results from Monte Carlo simulations in which the xed-connectivity
constraint is lifted and more complicated surface morphologies are considered [18].
5.6 Glide suppression
The dynamics of dislocations proceeds by means of two distinct processes:
glide and climb. Glide describes motion along the direction dened by the Burgers
vector; in at space glide requires a very low activation energy and is the dominant
form of motion at low temperature, see Fig. 5.3a inset. Climb, or motion perpendicu-
lar to the Burger vector requires diusion of vacancies and interstitials and is usually
frozen out relative to glide which involves only local rearrangements of atoms. On a
curved surface, the geometric potential D(r, ) imposes constraints on the glide dy-
namics of isolated dislocations, in sharp contrast to at space where the only energy
barriers present are due to the periodic Peierls potential [47].
As the dislocation represented in the inset of Fig. 5.3a moves in the glide di-
rection, it experiences a restoring potential generated by the variation of the (scaled)
radial distance and the deviation from the radial alignment of the dislocation dipole.
For a small transverse displacement, y, the harmonic potential U(y) =
1
2
k
d
y
2
is con-
trolled by a radial, position-dependent eective spring constant, k
d
which depends on
the radial coordinate r. Upon expanding Eq.(5.11) to leading order in y and , we
Chapter 5: Defects and crystalline order on curved surfaces. 126
(b)
(a)
4 3 2 1 0 1 2 3 4
0
5
10
15
20
E(y)
Ybr
0
y
0 0.5 1 1.5 2 2.5
0
0.5
1
1.5
2
k
d
(r)r
0
Yb
r
Figure 5.3: Dislocation glide. (a) Filled circles represent numerical glide energies
vs. y (units of a) for x
0
/a = 10, R/x
0
= 8, r = 0.5 (at y = 0): = 0.1 (dark
blue), 0.3 (blue), 0.5 (orange) and 0.7 (red). Solid lines represent E(y) =
1
2
k
d
y
2
, with
k
d
determined from Eq. (5.13). The energy is scaled by 10
4
. Inset: schematic of
a dislocation (red) on a Gaussian bump. The glide path is highlighted in orange.
(b) The curvature-induced glide-suppression spring constant k
d
(r) (5.13) (solid line)
is plotted versus scaled in-plane distance from the top of the bump r = x/x
0
for
x
0
= 10a, = 0.5. Open symbols represent numerical results found by tting similar
data as in (a) to parabolic curves in the glide coordinate. The energy is scaled by
10
2
.
obtain
k
d
(r)

2
Y b
4 x
0
_
1 (1 +r
2
)e
r
2
r
3
_
. (5.13)
The harmonic potential
1
2
k
d
y
2
is shown in Fig. 5.3a where the data obtained from
numerical minimization of the harmonic lattice is explicitly compared to the predic-
tion of Eq.(5.13) for dierent values of the aspect ratio. Note that the eective spring
constant plotted in Fig. 5.3b vanishes in the limit
b
x
0
0 but can still be important
for small systems since Y b
2
is of the order of hundreds of k
B
T
1
(see section
5.3). The conning potential plotted in Fig. 5.3a is similar to the one experienced
by a dislocation bound to a disclination [81, 10].
Chapter 5: Defects and crystalline order on curved surfaces. 127
The resulting thermal motion in the glide direction of dislocations in this
binding harmonic potential is modeled by an over-damped Langevin equation for the
glide coordinate, y. We have y
2
) =
1e
2k
d
t
k
d
[81, 10]. In the case of a bump
geometry, the eective spring constant k
d
can be evaluated using Eq.(5.13). We
emphasize, however, that the glide suppression mechanism considered here is not
caused by the interaction of the dislocation with other defects but purely by the
geometric interaction with the curvature of the substrate.
5.7 Vacancies, Interstitials and Impurities
We now turn to a derivation of the geometric potential, I(x

), for inter-
stitials, isotropic vacancies and impurities. Inspection of Fig. 5.4 reveals that an
interstitial (vacancy) can be viewed either as the product of locally adding (remov-
ing) an atom to the lattice or as a composite object made up of three disclination
dipoles. In order to derive I(x

), we substitute the source term of Eq.(5.7) into


Eq.(5.8) and integrate by parts twice. The result reads
I(x

)
Y
2

V (x

) , (5.14)
where boundary terms and a position independent nucleation energy have been dropped.
The constant

represents the area excess or decit associated with the defect. In


Fig. 5.4 a comparison of Eq.(5.14) with the results obtained from mapping out I(r

)
numerically is presented. The area changes
i
and
v
for interstitials and vacancies
respectively were t to the numerical data. We nd that
v
is negative and greater in
magnitude than
i
. The large r behavior of I(r) indicates that the core energy of a
Chapter 5: Defects and crystalline order on curved surfaces. 128
0 0.5 1 1.5 2 2.5
4.15
4.2
4.25
4.3
4.35
4.4
4.45
4.5
I(r)
Ya
2
r
Figure 5.4: The scaled potential energy I(x)/Ya
2
of interstitials (bottom) and va-
cancies (top) is plotted versus the scaled distance r for the same bump as Fig. 5.1.
Solid lines are obtained from Eq. (5.14), where the area changes of interstitials and
vacancies

i
and

v
were t to the data. Sample congurations are plotted in the
insets, where diamonds, circles and squares represent ve, six and seven fold coordi-
nated particles respectively. Filled and unlled circles in the interstitial plot represent
two distinct but energetically equivalent orientations of the disclination dipoles com-
prising the interstitial. They dier by swapping the 5 and 7-fold disclinations, which
rotates the orientation of the lled, triangular plaquette by

3
.
vacancy in at space is greater than the one of an interstitial for the harmonic lattice.
Interstitials tend to seek the top of the bump whereas vacancies are pushed in the at
space regions. From the denition of V (r) in Eq.(5.3), we deduce that an interstitial
(vacancy) is attracted (repelled) by regions of positive (negative) Gaussian curva-
ture similar to the behavior of an electrostatic charge interacting with a background
charge distribution given by G(r). The function V (r) controls the curvature-defect
Chapter 5: Defects and crystalline order on curved surfaces. 129
interaction for other types of defects that can be modeled as a Coulomb gas in curved
space such as disclinations in liquid crystals and vortices in
4
He lms [58, 43]. The
expression for V (r) in Eq.(5.3) reveals that I(r) is indeed a non local function of
the Gaussian curvature determined as in electrostatics by the application of Gauss
law. Thus, the vacancy potential on a bump does not reach a minimum at the point
where the Gaussian curvature is maximally negative but rather at innity where the
integrated Gaussian curvature vanishes.
We now argue heuristically how a localized point defect can couple non-
locally to the curvature and in the process we provide an alternative justication of
Eq. (5.14) analogous to the informal derivation of the dislocation potential. The
energy cost, I(x

), of a local compression or stretching in the presence of an ar-


bitrary elastic stress tensor
ij
(x) is given by I(x

) = p(x

)V where V is the
local volume change and p(

) is the local pressure related to the stress tensor via

ij
(x

) = p(x

)
ij
[14]. In two dimensions, we have I(x

) =
kk
(x

2
. We re-
cover the result in Eq.(5.14), by assuming that the local deformation

(induced by
the nucleation of a point defect) couples to the preexisting stress of geometric frus-
tration
G
kk
= Y V (x) (see Eq.(5.10)), which is a non-local function of the Gaussian
curvature. Elastic deformations created by the geometric constraint throughout the
curved two dimensional solid are propagated to the position of the point defect by
force chains spanning the entire system. The point defect can then be viewed as a
local probe of the stress eld that does not measure the additional stresses induced
by its own presence.
Chapter 5: Defects and crystalline order on curved surfaces. 130
Note that the geometrical potential of an isotropic point defect is unchanged
if we swap the 5 and 7-fold disclinations comprising it (corresponding to a rotation of
the point defect by

3
around its center), as demonstrated for interstitials in the lower
branch of Fig. 5.4. Contrast this situation with the clear dipolar character of the
dislocation potential plotted in Fig. 5.1. The more complicated case of non-isotropic
point defects appears to still be captured qualitatively by Eq.(5.14); this was explicitly
checked by plotting the geometric potential of crushed vacancies [82], which have
both a lower symmetry and a lower energy than their isotropic counterparts.
An arbitrary conguration of weakly interacting point defects will relax to
its equilibrium distribution by diusive motion in a force eld f(r)= I(r). This
geometric force leads to a biased diusion dynamics with drift velocity v |f|a
D
a

DY
x
0
, where =
1
k
B
T
and D is the defect diusivity [47]. Eventually, a dilute gas of
point defects equilibrates to a non-unform spatial density proportional to e
I(r)
.
Appendix A
Greens function and Isothermal
Coordinates
131
Appendix A: Greens function and Isothermal Coordinates 132
N
Conformal plane
S
Figure A.1: Graphic construction of the stereographic projection. Regions close to
the north pole have larger images in the conformal plane than regions of equal areas
close to the south pole. The stereographic projection preserves the topology of the
surface provided all points at innity are identied with the north pole.
The analysis of ordered phases on curved substrates can be simplied by
rewriting the original metric of the surface g
ab
(u) in terms of a convenient set of
coordinates x(u) = (x(u), y(u)) such that the new metric g
ab
(x) reads
g
ab
(x) = e
(x,y)

ab
. (A.1)
The metric g
ab
diers from the at space one
ab
only by a conformal factor e
(x,y)
that embodies information on the curvature of the surface [44]. These isothermal
coordinates can be used to map arbitrary corrugated surfaces onto the plane [83]. The
mapping is conformal so angles are left unchanged but areas are stretched according to
the position-dependent conformal factor e
(x,y)
. A familiar example is provided by the
stereographic projection that maps a sphere onto the conformal plane as illustrated
in Fig. A.1. The Greens function assumes a very simple form after a conformal
transformation, because the Laplace operator reduces to the familiar at space result
when expressed in terms of isothermal coordinates. In what follows, we demonstrate
that this transformation provides the basis for an ecient strategy to determine the
Appendix A: Greens function and Isothermal Coordinates 133
Greens function on a bumpy substrate. We start by deriving the radial change of
coordinates '(r) that transforms the original metric of the Gaussian bump, i.e.,
ds
2
=
_
1 +

2
r
2
r
2
o
e

r
2
r
2
o
_
dr
2
+r
2
d
2
, (A.2)
into the locally at metric (in polar coordinates),
ds
2
= e
(r)
(d'
2
+'
2
d
2
) , (A.3)
where (r) and '(r) are independent of the azimuthal coordinate because of cylin-
drical symmetry. This metric is equivalent to g
ab
(x) upon switching from cartesian
(x,y) to polar coordinates ('(r), ). To simplify the notation we introduce the -
dependent function l(r) dened by
l(r) 1 +

2
r
2
r
2
0
exp
_

r
2
r
2
0
_
, (A.4)
and plotted in Fig. 2.2 for dierent choices of .
The equivalence of the metrics in Equations (A.2) and (A.3) requires that
'(r) satises the dierential equation
d'
'
=
_
l(r)
r
dr . (A.5)
The conformal factor is thus given by
e
(r)
= (
r
'
)
2
. (A.6)
The solution of Eq.(A.5) is
'(r) = Ar e

R
c
r
dr

l(r

) 1)
, (A.7)
Appendix A: Greens function and Isothermal Coordinates 134
where it is convenient to set the arbitrary constants A and c to unity and innity
respectively. This non-linear stretch of the radial coordinate leaves the origin and the
point at innity invariant and can be concisely written as
'(r) = r e
V (r)
, (A.8)
where the function V (r) dened by
V (r)
_

r
dr

_
l(r

) 1
r

, (A.9)
plays an important role in our formalism and its interpretation as a sort of geometric
potential is explored in detail in Appendix B.
The Poisson equation for the Greens function (u, u

) on a surface with
metric tensor g

and point source (u, u

) reads [40]:
D

(u, u

) =
(u, u

g
, (A.10)
where the covariant Laplacian is given for general coordinates by:
D

(1/

g)

gg

] . (A.11)
The conformal change of coordinates transforms
_
g(r, ) into e
(r)
_
g(', ) and
g

(r, ) into e
(r)
g

(', ). The factors of e


(r)
inside the square brackets in
Eq.(A.11) then cancel and we are left with the at space Laplacian in the polar
coordinates ('(r), ). We conclude that (u, u

) is simply the Greens function of an


undeformed plane expressed in terms of the polar coordinates ('(r), ):
(u, u

) =
1
4
ln['(r)
2
+'(r

)
2
2'(r)'(r

) cos(

)] , (A.12)
Appendix A: Greens function and Isothermal Coordinates 135
where an arbitrary additive constant C (which can be used to satisfy boundary con-
ditions at innity) has been dropped. We note that (u, u

) diers from the at


space Greens function by a non linear stretch of the radial coordinate. In Appendix
C, we will use (u, u

) to solve Poissons equation on an innite bumpy domain and


calculate the energy stored in the eld. As in at space, the Greens function (u, u

)
will be suitably modied in a nite system in a way that depends on the boundary
conditions chosen at the edge of the sample (see Appendix D).
We conclude this appendix by evaluating (u, u

) when the two points u and


u

are assumed to be separated by a xed distance a small enough so that the surface
can be approximated by the local tangent plane to the Gaussian bump. This short
distance behavior will be useful when evaluating the eect of a constant core radius
on the energetics of a disclination at an arbitrary position. The xed microscopic
length a on the bump is stretched when projected in the conformal plane (see, e.g.,
Fig. A.1) and assumes the position dependent value (x, y) given by
(x, y) = ae

(x,y)
2
. (A.13)
For a Gaussian bump cylindrical symmetry requires that (r, ) is dependent only on
r and can be explicitly written upon using Equations (A.6) and (A.8) as
(r, ) = a
'(r)
r
= ae
V (r)
. (A.14)
We can now evaluate (u, u

) in the limit u

u +a, where this concise notation


means that the two points on the surface with coordinates u and u

are separated
by an innitesimal distance a measured on the bump. It does not matter in what
Appendix A: Greens function and Isothermal Coordinates 136
direction the two points approach each other as long as a is small compared to the
local radii of curvature. Upon using Equations (A.12) and (A.14), we obtain
(u, u +a) =
1
4
ln(a
2
)
V (r)
2
. (A.15)
We note that (u, u + a) for xed a is not a constant like in at space but varies
with position as the function V (r), reecting the lack of translational invariance on
an inhomogeneous surface, where properties such as the Gaussian curvature also vary
with position.
Appendix B
Geometric Potential
137
Appendix B: Geometric Potential 138
In this appendix we present two equivalent ways of determining the explicit
form of the geometrical potential V (u) valid for azimuthally symmetric surfaces like
the Gaussian bump. The starting point is the general denition introduced in Section
2.2.3:
V (u)
_
dA

G(u

) (u, u

) , (B.1)
with the Greens function as dened in Eq.(A.10). In the electrostatic analogy, V (u)
is thus the potential induced by a continuous distribution of charge represented by
the Gaussian curvature (with sign reversed). We shall derive an analogue of Gauss
law for corrugated surfaces where the curvature of the surface (with sign reversed)
plays the role of a continuous density of electrostatic charge.
The rst derivation makes use of the fact that V (u) is a scalar under confor-
mal transformations. This symmetry can be checked explicitly by applying the same
reasoning adopted for the equation satised by the Greens function in Appendix A
to Eq.(B.1). In fact, upon operating on both sides of Eq.(B.1) with the covariant
Laplacian and using Eq.(A.10), the dening equation for the geometric potential can
be cast into the dierential form:
D

V (u) = G(u) . (B.2)


The Gaussian curvature in Eq.(B.2) can be written in conformal coordinates [44] as
G(x, y) = e
(x,y)
(
2
x
+
2
y
)
(x, y)
2
, (B.3)
where (x, y) is the conformal factor introduced in Appendix A. Similarly the left
Appendix B: Geometric Potential 139
hand side of Eq.(B.2) can be expressed in conformal coordinates [44] as
D

V (x) = +e
(x,y)
(
2
x
+
2
y
)V (x, y) . (B.4)
Upon substituting Equations (B.3) and (B.4) in Eq.(B.2), we conclude immediately
that the geometric potential in conformal coordinates V (x) is given by
V (x, y) =
(x, y)
2
. (B.5)
Upon using Equations (A.6), (A.8) and (A.9) to substitute in Eq.(B.5), one obtains
the explicit form of the geometric potential for the bump parameterized by the coor-
dinates (r, ):
V (r) =
_

r
dr

_
l(r

) 1
r

, (B.6)
where the -dependent function l(r) was dened in Eq.(A.4) and plotted in Fig. 2.2.
The result of the integration in Eq.(B.6) is independent of because of azimuthal
symmetry. The upper limit of integration is chosen consistently with Eq.(B.2) as
usually done in electrostatics.
A second derivation of this result is obtained by making explicit use of the
azimuthal symmetry of the bump and deriving a covariant form of Gauss law which
allows an intuitive understanding of the interaction between defects and curvature.
This curved space version of Gauss law illuminates the electrostatic analogy used
throughout the text. The gradient of the geometric potential denes an electric
eld E

given by
E

V (u) =
_
dA

G(u

) D

(u, u

) , (B.7)
Appendix B: Geometric Potential 140
where we have used Eq.(B.1). One might expect that the ux of the vector E

through a closed loop is proportional to the enclosed Gaussian curvature in analogy


with Gauss law that relates the ux of the electric eld to the electrostatic charge
enclosed. To prove this assertion, we invoke the generalized Stokes formula [40] that
relates the surface integral over A of the gradient of a eld to its ux through the
contour loop C:
_
A
dA D

=
_
C
du

. (B.8)
The covariant antisymmetric tensor

is given by

, (B.9)
where

is the anti-symmetric tensor with


r
=
r
= 1. Similarly

equals

g
and the following identity holds [35]

. (B.10)
The tensor

performs anti-clockwise rotations of



2
when acting on an
arbitrary tangent vector V

, as can be checked by evaluating V

=
0, where we have used the antisymmetry of

. Thus, the vector du

in Eq.(B.8)
represents an innitesimal contour length times the inward unit vector perpendicular
to it. The dot product with the eld E

then generates the ux. To calculate the ux


piercing a circular circuit centered on a Gaussian bump, we will need to explicitly
evaluate

r
,

r
=
r
_
l(r)
. (B.11)
Appendix B: Geometric Potential 141
Upon using Eq.(B.7) we can rewrite the left hand side of Eq.(B.8) as
_
dA D

=
_
dA
_
dA

G(u

) D

(u, u

) . (B.12)
If we now recall the dening Equation (A.10) of the Greens function (u, u

) and
keep in mind that the Laplacian in Eq.(B.12) operates on the variables labelled by u

(not u), we obtain using Eq.(B.8) a general result for the ux piercing a closed loop
on the surface, namely
_
C
du

=
_
A
d
2
u

g G(u) . (B.13)
We can explicitely evaluate the right hand side of Eq.(B.13) with the aid of Eq.(2.12).
By appealing to the cylindrical symmetry, in the special case of the Gaussian bump
one can apply this covariant form of Gauss theorem to nd the radial eld E
r
(r) in
terms of the integrated Gaussian curvature divided by the length of a boundary circle
of radius r, with the result
E
r
=
1
_
l(r)
r l(r)
, (B.14)
where we used Equations (A.2) and (B.11). The angular component E

is zero
everywhere. Note that E
r
(r) vanishes linearly with r for small r and decays like
re

r
2
r
2
0
for r r
0
. From Eq.(B.14) one can obtain the geometric potential V (r) by
performing a line integral,
V (r) =
_

r
dr

g
rr
E
r
=
_

r
dr

_
l(r

) 1
r

, (B.15)
which matches the result previously obtained in Eq.(B.6).
Appendix C
Free energy on a corrugated plane
142
Appendix C: Free energy on a corrugated plane 143
In this Appendix, we derive the eective free energy for a charge neutral
conguration of defects conned on an innite surface of varying Gaussian curvature
with the topology of the plane. A general method was introduced in Ref.[43] that
allows treatments of the more complicated case of deformed spheres. A detailed
treatment of boundary-eects is developed in Appendix D. Here we simply assume
that the size of the system is much larger than the size of the bump and that the
boundary does not impose any topological constraint to the director of the liquid
crystal. The results presented here match those obtained in Appendix D for free
boundary conditions, as long as the defects are far from the boundary. Suppose that
all the N
d
defects have the same circular core radius a which does not depend on where
they are located on the surface. This assumptions is justied if the radius of curvature
r
0

is much greater than a. In this limit, the microscopic physics that determines a
is insensitive to the presence of the curvature and since the bump is locally at a
is approximately constant everywhere. The starting point of our analysis is the free
energy expressed in terms of the singular part of the bond angle
s
(u)
F =
1
2
K
A
_
S
dAg

s
A

)(

s
A

) . (C.1)
The cores of the defects are excluded from the area integral in Eq.(C.1), hence S
is a disconnected domain corresponding to the corrugated surface punctured at the
positions u
i
= (r
i
,
i
) of the defects. In the conformal plane parameterized by the
coordinate '(r) dened in Eq.(A.8), the defect cores are circles whose position de-
pendent radius is given by ae
V (r
i
)
. The boundaries of the core-disks are labelled by
C
i
while the circular edge of the sample of radius R is denoted by B, see Fig. C.1.
Appendix C: Free energy on a corrugated plane 144


C3
C1
C2
B
(a)
(b)
Figure C.1: (a) Defects with xed core size a on a Gaussian bump encircled by a
circular boundary of radius R denoted by B. (b) The size of the vortex cores varies
with position when plotted in the conformal plane. One can avoid the singularities
associated with the defects cores by puncturing the conformal plane. This introduces
circular boundaries C
i
of varying radius at the position of each defect in the conformal
plane, reecting corresponding constant core radii on the Gaussian bump.
Upon introducing a Cauchy conjugate function (u) dened by

s
A

, (C.2)
the free energy in Eq.(C.1) can be cast in the form:
F =
1
2
K
A
_
S
dAg

)(

) . (C.3)
In deriving Eq.(C.3) we used the identity
g

= g

, (C.4)
which can be proved with the aid of Eq.(B.10) and the discussion following it.
Eq.(C.4) implies that the (covariant) dot product between two vectors after rotating
each of them by

2
is equivalent to taking the dot product between the two initial
vectors. The integral in Eq.(C.3) can be rewritten as
F
K
A
=
1
2
_
S
dAD

(D

)
1
2
_
S
dAD

, (C.5)
Appendix C: Free energy on a corrugated plane 145
where
D

(D

) (1/

g)

gg

) , (C.6)
and D

is dened in Eq.(A.11). Upon taking an additional covariant derivative,


we can recast Eq.(C.2) in the form of a Poisson equation for the electrostatic-like
potential (u),
D

(u) = (u) , (C.7)


where the analogue of the electrostatic charge density (u) is given by
(u)
N
d

i=1
q
i
(u, u
i
)

g
G(u) . (C.8)
It is useful to compare Eq.(C.7) to Eq.(B.2) used to dene the geometric potential
in Appendix B. Both expressions are Poisson equations, the only dierence being
that the source term of Eq.(C.8) includes both the point-like charges of the defects
and the Gaussian curvature with its sign reversed. Hence the Gauss law discussed in
Appendix B for the geometric eld E

applies also to

, provided that Eq.(B.13)


is suitably modied to include the contribution from the topological charges of the
defects:
_
C
du

=
_
A
d
2
u

g
_
G(u)
N
d

i=1
q
i
(u, u
i
)

g
_
, (C.9)
where C is the contour enclosing an arbitrary surface A. This relation will be useful
later.
We can formally solve for (u) in Eq.(C.7) in terms of the Greens function
(u, u

) found in Appendix A:
(u) =
_
A
dA

(u

) (u, u

) , (C.10)
Appendix C: Free energy on a corrugated plane 146
where boundary terms have been dropped using charge neutrality and the fact that
the edge of the sample is assumed to be far away from the defects. The integral in
Eq.(C.10) can be evaluated, with the result
(u) =
N
d

i=1
q
i
(u, u
i
)
_
A
dA

G(u

) (u, u

) . (C.11)
Upon using Eq.(B.1), we obtain:
(u) =
N
d

i=1
q
i
(u, u
i
) +V (u) . (C.12)
We note that the rst term is singular at positions u
i
but when is substituted in
Eq.(C.3) the resulting energy is nite because the core of the defects are excluded
from the domain of integration S. Upon substituting Eq.(C.7) in the second term of
Eq.(C.5) we obtain:
_
S
dAD

=
_
S
dA(u)
=
_
S
dAG(u) , (C.13)
where we dropped terms involving the delta functions in Eq.(C.8) because they vanish
everywhere except at the coordinates of the defects which are excluded from the
domain of integration S. Upon substituting Eq.(C.12) in Eq.(C.13) we obtain:

_
S
dAD

=
N
d

i=1
q
i
V (u
i
)
+
_
dA
_
dA

G(u) (u, u

) G(u

) , (C.14)
where we used Eq.(B.1).
To evaluate the rst term in Eq.(C.5), we apply the generalized Stokes for-
mula of Eq.(B.8) and convert the surface integrals into line integrals over the bound-
Appendix C: Free energy on a corrugated plane 147
aries:
_
S
dA D

(D

) =
N
d

i=1
_
C
i
du

_
B
du

, (C.15)
where the dierence in sign between the two boundary integrals in Eq.(C.15) is due
to the fact that the outward normals for the paths C
i
are oriented opposite to the
normal for B, the outermost boundary of the system.
To evaluate the last term in Eq.(C.15), we note that the ux through the
distant boundary B due to a charge neutral distribution of defects is approximately
zero, provided that the integrated Gaussian curvature enclosed by the boundary is
vanishingly small (see Eq.(C.9)). Hence

_
B
du

0 , (C.16)
where we used the fact that , dened in Eq.(C.2), is approximately constant on B
since

0. By contrast, the ux of
r
piercing the boundary C
i
in Eq.(C.15)
is approximately equal to the charge q
i
of the inclosed defect
1
. In evaluating the
integrals around the innitesimal boundaries C
i
, we used the fact that the function
(u
i
+ a) evaluated on the rim of the defect core centered at u
i
and of radius
ae
V (u
i
)
is dominated by a logarithmically diverging contribution due to the i
th
defect.
This leading contribution is approximately constant on C
i
. On the other hand, the
non diverging part of (u
i
+ a) is multiplied by the perimeter of C
i
and hence its
contribution is of the order of a. The result of the integration will be insensitive to
1
The integrated Gaussian curvature in the microscopic disk is vanishingly small.
Appendix C: Free energy on a corrugated plane 148
the orientation of the vectors along the boundary C
i
, provided the defect core is small
2
. In this way, we nd
_
C
i
du

(u
i
+a)
_
C
i
du

q
i
(u
i
+a) . (C.17)
Upon substituting Equations (C.17) and (C.16) in Eq.(C.15), we obtain
_
S
dAD

(D

) =
N
d

i=1
q
i
(u
i
+a) =
N
d

i=1
q
i
V (u
i
) +
N
d

i=1
N
d

j=i
q
i
q
j
(u
i
, u
j
)
+
N
d

i=1
q
2
i
(u
i
, u
i
+a) , (C.18)
where we used Eq.(C.12) to substitute for . Substitution of Eq.(C.18) and Eq.(C.14)
in Eq.(C.5) then yields:
F
K
A
= F
0
+
N
d

i=1
q
i
(1
q
i
4
)V (r
i
)
N
d

i=1
q
i
2
8
ln(a
2
)
+
1
2
N
d

i=1
N
d

j=i
q
i
q
j
(u
i
, u
j
) , (C.19)
where we have used Eq.(A.15) to evaluate (u
i
, u
i
+a). The rst term in Eq.(C.19)
is the free energy of the smooth defect-free texture (see Eq.(2.17) and preceding dis-
cussion). The free energy dierence
F
K
A
between a charge neutral defect conguration
2
If the defect core is very large, it may be necessary to place images within the defect core itself to impose
a desired boundary condition on its rim. This is not the regime considered in the present work (see Ref.
[84] for similar calculations performed in at space).
Appendix C: Free energy on a corrugated plane 149
and a smooth texture thus reads
F()
K
A
=
1
2
N
d

i=1
N
d

j=i
q
i
q
j

a
(r
i
,
i
, r
j
,
j
) +
E
c
K
A
N
d

i=1
q
2
i
+
N
d

i=1
q
i
_
1
q
i
4
_
V (r
i
) , (C.20)
where the subscript a in the Greens function indicates

a
(r
i
,
i
, r
j
,
j
) =
1
4
ln[
'(r)
2
a
2
+
'(r

)
2
a
2
2
'(r)
a
'(r

)
a
cos(

)] . (C.21)
In order to absorb the core radius a in the inter-defect interaction, we used the
elementary algebraic identity
N
d

i=1
q
2
i
=
N
d

i=1
N
d

j=i
q
i
q
j
, (C.22)
valid for charge neutral congurations. The core energy E
c
in Eq.(C.20) was added
by hand and represents short distance physics on scales less than or equal to the core
radius. Although the second part of the last term in Eq.(C.20) arises as a position
dependent self energy, it has the same functional form as the geometrical potential
discussed in Appendix B, and hence depends on the global shape of the surface.
Appendix D
Bump with a boundary
150
Appendix D: Bump with a boundary 151
(a)
(b)
R
R
Figure D.1: (a) Schematic illustration of the boundary director texture corresponding
to free boundary conditions. The vector order parameter orientation close to the
edge of the sample does not vary appreciably as one moves along the radial direction
and it is parallel to itself at every point on the boundary (b) Tangential boundary
conditions. The vector order parameter is locally aligned to a wall located at the edge
of the sample.
The aim of this appendix is to study the energetics of a singular vector eld
on a bumpy surface of circular shape and nite size R. To evaluate the ground state
energy in Eq.(C.1), we rst need to solve the covariant Laplace equation for the bond
angle eld (u) in the presence of N
d
defects at positions u
i
. This is more easily
done by switching from (u) to the conjugate eld (u), as shown in Eq.(C.3), and
solving the Poisson Equation (C.7) satised by for both free and xed boundary
conditions (see Fig. D.1). The bond angle eld satises free boundary conditions if
the following relation holds on the circular edge B:

r
|
r=R
= 0 , (D.1)
while for xed tangential boundary conditions we have:

|
r=R
= 0 . (D.2)
To understand Eq.(D.2), recall that we measure the bond angle with respect to a
rotating basis vector E
r
in the radial direction. With this convention, is equal to a
Appendix D: Bump with a boundary 152
constant when the vector order parameter is aligned with the circular boundary B.
We can convert Equations (D.1) and (D.2) into boundary conditions to be satised
by the conjugate eld on B. Upon substituting Eq.(D.1) in Eq.(C.2) and using the
fact that A
r
is equal to zero, we obtain the constraint that (u) satises on B in the
case of free boundary conditions:

|
r=R
= 0 . (D.3)
This corresponds to a Dirichlet problem where
D
(u) evaluated on the boundary B
assumes an arbitrary constant value, c:

D
(B) = c . (D.4)
Upon substituting Eq.(D.2) in Eq.(C.2), we obtain

r
=
A

r
, (D.5)
since

= 0. Upon substituting Equations (B.11) and (2.8) in Eq.(D.5) we obtain


the boundary condition on that corresponds to Eq.(D.2)

N
|
r=R
=
1
R
, (D.6)
where the superscript indicates that this is a Neumann boundary problem with the
normal derivative assuming a constant value.
To solve the Poisson Eq.(C.7) with Neumann or Dirichlets boundary condi-
tions in terms of suitable Greens functions we exploit a covariant version of Greens
Appendix D: Bump with a boundary 153
theorem expressed in terms of two invariant functions of position (u) and (u) [85]:
_
S
dA ((u) D

(u) (u) D

(u)) =

_
B
du

((u)

(u) (u)

(u)) . (D.7)
By applying Eq.(D.7) to (u) = (u) and (u) = (u, u

) and using Equations


(A.10) and (C.7) we obtain
(u) =
_
S
dA

(u

, u) (u

)
+
_
B
du

(u

(u

, u)

_
B
du

(u

, u)

(u

) , (D.8)
where u and u

have been exchanged


1
. The boundary conditions for the Greens
function can be conveniently chosen to eliminate unknown quantities in Eq.(D.7) as
in at space [86].
For the Dirichlets problem, we choose the Greens function
D
so that it
vanishes when u

is on the boundary B

D
(B, u) = 0 . (D.9)
Upon substituting Eq.(D.9) in Eq.(D.8) and noting that (u

) is constant on the
boundary B (see Eq.(D.4)), we obtain:

D
(u) =
_
S
dA


D
(u

, u) (u

)
+
D
(B)
_
B
du

D
(u

, u) , (D.10)
1
These manipulations are common in electromagnetism, see for example Ref. [86].
Appendix D: Bump with a boundary 154
The contour integral in the second term of Eq.(D.10) corresponds to the ux piercing
B which is, in turn, equal to the (unit) charge of the singularity located at u
2
. The
nal result reads

D
(u) =
_
S
dA

(u

)
D
(u

, u) +
D
(B) , (D.11)
where
D
(B) can be set to zero, since the energy in Eq.(C.3) is only dened in terms
of derivatives of . One can check that
D
(u) in Eq.(D.11) satises both the Poissons
Equation (C.7) and the required boundary condition in Eq.(D.3). This can be more
easily proved by noting that
D
(u

, u) is symmetric under exchange of its arguments


u

and u
3
.
A similar reasoning applies to
N
(u). However, we cannot choose the
Greens function so that the second term of Eq.(D.8) (that contains the unknown
quantity (u

)) vanishes. In fact, by invoking Stokes theorem (see Eq.(B.8)), we note


that
_
B
du

(u

, u) = 1, (D.12)
where

(u

= B) is constant on the circular boundary B and can be brought out


of the integral. An appropriate choice of boundary condition on
N
that satises the
constraint in Eq.(D.12) is

r

N
(r

; r, )|
r

=R
=
1
2

(R)
=
_
l(R)
2R
, (D.13)
2
This assertion can be proved by applying Stokes Theorem (as stated in Eq.(B.8) with E

replaced by

(u

, u)) and using Eq.(A.10)) to evaluate the surface integral.


3
This assertion can be proved by applying Greens theorem in Eq.D.7 to (u) = (u, u

) and (u) =
(u, u

) and noticing that the right hand side vanishes if the boundary condition in Eq.(D.9) is assumed.
We can then conclude that
D
(u

, u

) =
D
(u

, u

) in analogy with familiar results in 3D electrostatics


[86].
Appendix D: Bump with a boundary 155
where we used Eq.(B.11). The -dependent function l(r

) was dened in Eq.(A.4).


Note that l(R) 1, for R r
0
(see Fig. 2.2). By substituting Equations (D.13) and
(D.6) in Eq.(D.8) we obtain

N
(u) =
_
S
dA


N
(u

, u) (u

)
+
1
2
_
2
0
d

(R,

1
_
l(R)
_
2
0
d

N
(R,

; r, ) , (D.14)
The last two integral are constant and hence can be dropped. We can check explicitly
that
N
(u) satises Eq.(D.6) by evaluating the radial derivative of
N
(u) in Eq.(D.14)

N
(r, )|
r=R
=
_
S
dA

(u

)
r

N
(R,

; r, )|
r=R
, (D.15)
where the radial derivative of the Greens function assumes the constant value derived
in Eq.(D.13), provided that
N
(u

, u) is constructed so that it is symmetric under


exchange of its arguments u

and u (see Eq.(D.19)). With the aid of Equations (C.8)


and (2.12), we obtain
_
S
dA

(u

) =
N
d

i
q
i
+ 2
_
1
_
l(R)
1
_
. (D.16)
Upon substituting Equations (D.16) and (D.13) in Eq.(D.15), we conclude that the
the Neumann boundary condition in Eq.(D.6) is satised provided that
N
d

i
q
i
= 2 . (D.17)
It is reassuring that the topological constraint on the vorticity of the eld imposed
by the presence of the wall arises as a natural requirement within this formalism.
Similarly, the Poissons equation for
N
(u) is automatically satised.
Appendix D: Bump with a boundary 156
R
B
(a)
R
B
(b)
- - + +
+ +
Figure D.2: Schematic illustration of the method of images. The image defect is
of the same sign for free boundary conditions (a) and opposite for xed boundary
conditions (b). Defects closer to the center of the circle have images further away
from it.
We are now left with the task of guessing the Greens functions for the
Dirichlets and Neumann problems satisfying the boundary conditions in Equations
(D.9) and (D.13) respectively. In both cases the Greens function can be determined
by the method of images applied in the conformal plane. For every defect with
radial coordinate r
i
we need an image defect of opposite (equal) topological charge
at position r

i
to ensure that Dirichlets (Neumann) boundary conditions are enforced
(see Fig. D.2). The radial coordinate of the image defect r

i
is determined by the
relation:
'(r

i
) =
'
2
(R)
'(r
i
)
. (D.18)
Except for the coordinate change r '(r), a similar relation arises in elementary
electrostatic problems in at space [48]. A geometric argument that justies this
choice of images in at space is illustrated in Fig. D.3.
Once the position of the source is chosen according to Eq.(D.18), we can
Appendix D: Bump with a boundary 157
express the two Greens functions with the concise notation
D/N
as follows:

D/N
(u, u

) =
1
4
ln['(r)
2
+'(r

)
2
2'(r)'(r

) cos(

)]

1
4
ln['(r)
2
+
'(R)
4
'(r

)
2
2'(r)
'(R)
2
'(r

)
cos(

)] f(r

) , (D.19)
where we have introduced a function f(r

) to make
D/N
(u, u

) symmetric under
exchange of its arguments and to remove a singularity at r

= 0. Note that we
can add f(r

) since the dening equation of the Greens function does not contain
derivatives of r

, only of r.
f(r

) =
1
2
ln
_
'(r

)
'(R)
_
. (D.20)
The plus and minus signs in Eq.(D.19) insure that the Dirichlet and Neumann bound-
ary conditions respectively are obeyed. In what follows the sign placed above in the
symbols or always indicates the choice suitable for the Dirichlets problem while
the one below refers to Neumanns boundary conditions. One can explicitly check by
substitution that the symmetrized Greens functions
D/N
(u, u

) satisfy the correct


boundary conditions, as long as the plus sign is chosen when
D
is substituted in
Eq.(D.9) and the minus sign when
N
is substituted in Eq.(D.13). Note that, with-
out the extra term f(r

) in the expressions for both Greens functions,


D
would not
be equal to zero on the boundary B and the last term in Eq.(D.14) would not be
constant when
N
is substituted in.
Once the Greens function is obtained, one can readily write down
D/N
(u)
Appendix D: Bump with a boundary 158
by dropping the constant terms in Equations (D.11) and (D.14)

D/N
(u) =
N
d

i=1
q
i

D/N
(u, u
i
)

_
S
dA

G(u

)
D/N
(u, u

) . (D.21)
The Gaussian curvature is given by the covariant Laplacian of the geometric potential
introduced in Eq.(B.2). Upon integrating by parts twice the second term in Eq.(D.21)
and applying Stokes theorem repeatedly, we nd

D/N
(u) =
N
d

i=1
q
i

D/N
(u, u
i
) +V (u) , (D.22)
where we assume that R r
0
so that we can neglect boundary terms. The geometric
potential V (u) has the same functional form previously discussed in Appendix B,
despite the change in the Greens function.
The evaluation of the energy stored in the eld now proceeds along the lines
sketched in Appendix C with the only caveat that one needs to choose the appropriate
Greens function in Eq.(D.19). In the case of Dirichlets boundary conditions one can
prove that the boundary integral in Eq.(C.16) still vanishes by virtue of the fact that
is constant on the boundary B and the defects conguration is charge neutral. For
the Neumanns problem we have:

N
_
B
du

N
4 ln ('(R)) , (D.23)
where we assumed that R r
0
. In this limit '(R) is approximately equal to R, as
can be checked with the aid of Equations (A.8) and (A.9).
All the remaining intermediate steps to derive the free energy follow as in
Appendix C without further assumptions. We can readily generalize Eq.(C.20) to
Appendix D: Bump with a boundary 159
Figure D.3: A topological defect located at position P in a circular domain of radius
OQ in at space. Fixed boundary conditions are obtained by placing an image defect
of the same sign at a distance OP

from the center such that OP OP

= OQ
2
. The
two triangles OQP and OQP

are similar and OQP = OP

Q =

. By the
theorem of the external angle, we conclude that

+ = + as long as Q lies on the


circumference B. This is equivalent to the boundary condition in Eq.(D.2) if a non
rotating vector basis is used. Similarly, for free boundary conditions the symmetric
Green function evaluated is constant on the boundary if the image defect is negative.
Since
PQ
P

Q
=
OP
OQ
, the potential ln(PQ) ln(P

Q) (generated by the defect at distance


OP and its image) is constant on the circumference of radius OQ.
Appendix D: Bump with a boundary 160
evaluate the energy stored in the singular eld in the presence of a boundary in the
case of both free and xed boundary conditions. We assume R r
0
, but the defects
do not need to be far away from the boundary. The result is
F
D/N
K
A
=
1
2
N
d

j=i
q
i
q
j

D/N
(x
i
; x
j
) +F
0
+
N
d

i=1
q
i
(1
q
i
4
)V (r
i
) +
N
d

i=1
q
i
2
4
ln
_
'(R)
a
_

N
d

i=1
q
i
2
4
ln
_
1 x
2
i

, (D.24)
where F
0
is dened in Eq.(2.17). The Greens function expressed in scaled coordinates
reads

D/N
(x
i
; x
j
) =
1
4
ln
_
x
2
i
+x
2
j
2x
i
x
j
cos (
i

j
)

1
4
ln
_
x
2
i
x
2
j
+ 1 2x
i
x
j
cos (
i

j
)

.
(D.25)
In the case of Neumanns boundary conditions, we have suppressed a term diverging
like ln('(R)) associated with the boundary contribution in Eq.(D.23). Eq.(D.25) is
expressed in terms of a dimensionless defect position x
i
x
i

'(r
i
)
'(R)
. (D.26)
The plus sign in Equations (D.25) and (D.24) is to be chosen for Dirichlets boundary
conditions and the minus sign for Neumanns. The last term in Eq.(D.24) represents
the interaction , U
D/N
b
(x
i
), between a single defect located at x
i
and the boundary
U
D/N
b
(x
i
) = K
A
N
d

i=1
q
i
2
4
ln
_
1 x
2
i

. (D.27)
Appendix D: Bump with a boundary 161
Note that the q-dependent prefactors of U
D/N
b
(x
i
) and the quadratic correction to the
curvature interaction (III term in Eq.(D.24) have the same magnitude. This is not a
coincidence but a clue to their common origin. As the geometry of a plane is modied,
either by creating a varying curvature or imposing boundaries, the defects feel an
additional interaction caused by the conformal transformation of the underlying space.
This line of reasoning is powerful and it has been pursued in Ref. [43] to explain
some basic features of the interaction between defects and curvature without explicit
recourse to the Greens function techniques adopted in this work.
Appendix E
Free energy of a vector eld on a
sphere
162
Appendix E: Free energy of a vector eld on a sphere 163
We start our analysis with the Frank free energy with splay and bend terms
proportional to K
1
and K
3
and expressed in terms of the covariant derivative D
i
n
j
F =
1
2
_
d
2
x

g [K
1
_
D
i
n
i
_
2
+ K
3
(D
i
n
j
D
j
n
i
)
_
D
i
n
j
D
j
n
i
_
] , (E.1)
where
D
i
n
j
=
i
n
j
+
j
i k
n
k
. (E.2)
The Christoel connection
j
i k
[60] is unchanged if the lower indices i and k are
interchanged. As a result, the covariant derivative D
i
can be replaced by
i
in the
second term of Eq.(E.1) because the covariant curl is antisymmetric. It follows that

D n = n. (E.3)
The covariant form of the divergence is given by [60]
D
i
n
i

g

i
_

g n
i
_
. (E.4)
For a rigid sphere of radius R with polar coordinates {, }, we have

g = R
2
sin ,


= sin cos ,


= cot , and all other
j
i k
= 0.
Upon adding and subtracting the expression for the curl of n multiplied by
K
1
from the rst and second term in Eq.(E.1) we obtain
F =
1
2
_
d
2
x

g [K
1
_
D
i
n
j
) ( D
i
n
j
_
+(K
3
K
1
)(n)
2
] . (E.5)
Similarly, upon adding and subtracting the covariant divergence of n multi-
Appendix E: Free energy of a vector eld on a sphere 164
plied by K
3
from the second and rst term in Eq.(E.1) we obtain
F =
1
2
_
d
2
x

g [K
3
_
D
i
n
j
) ( D
i
n
j
_
+(K
1
K
3
) ( n)
2
] . (E.6)
Upon adding the two equivalent expressions for F in Equations (E.5) and (E.6) and
dividing by two we can express the free energy in terms of the constants

K
3
K
1
K
3
+K
1
, (E.7)
K
K
3
+K
1
2
, (E.8)
namely
F =
K
2
_
d
2
x

g [
_
D
i
n
j
) ( D
i
n
j
_
+
_
(n)
2
( n)
2
_
] . (E.9)
We now parameterize the orientation of the unit vector n in terms of the
bond angle eld (, ) that it forms with respect to the longitudinal direction e

which in polar coordinates (, ) is given by


e

= R(cos sin , cos cos , sin ) , (E.10)


while the orthogonal unit vector e

is given by
e

= R(sin sin , sin cos , 0) . (E.11)


The components of the vector n with respect to e

and e

are given by
n

=
cos
R
, (E.12)
n

=
sin
Rsin
. (E.13)
Appendix E: Free energy of a vector eld on a sphere 165
Upon substituting the relevant expressions for the non vanishing components of the
connection in the covariant derivative (see Eq.(E.2)) and using Eq.(E.13), the free
energy density f(, ) is given by
4R
2
K
1
+K
3
f =
_

_
2
+

2
(, )
+ cos 2
_
_

_
2

2
(, )
_
+ 2 sin 2
_

(, ) , (E.14)
where the -dependent function

(, ) is

(, )
1
sin
_

+ cos
_
. (E.15)
The energies of the latitudinal ( = /2) and longitudinal ( = 0) tilted-molecules
textures favored for less or greater than zero respectively (see section 3.2.1) are easily
determined by substituting the appropriate and in Eq.(E.14). After integrating
the free energy density f in Eq.(E.14) we obtain
F = 2K(1 ||)
_
2
a
R
cos
2
()
sin()
d
= 2K(1 ||)
_
ln
_
2R
a
_
1
_
+ 2E
c
, (E.16)
where we have introduced a core radius, a, and corresponding core energy E
c
for each
defect.
In the zero anisotropy limit ( = 0), only the rst line in Eq.(E.15) survives.
The resulting free energy F =
_
dSf(, ) then matches the one obtained using the
spin connection in the one Frank constant approximation upon setting K
1
= K
3
= K,
F =
1
2
K
_
dS g
ij
(
i
A
i
)(
j
A
j
) , (E.17)
Appendix E: Free energy of a vector eld on a sphere 166
where dS = ddR
2
sin and the metric tensor g

= diag
_
1
R
2
,
1
R
2
sin
2

_
. The curl of
the spin-connection A
i
is the Gaussian curvature
1
R
2
[40, 42] and its only non-vanishing
component is A

= cos .
We now adopt a Coulomb gas representation of the liquid crystal free energy
(in the one Frank constant approximation) obtained by exploiting in Eq.(E.17) the
relation

ij

i
(
j
A
j
) = s(u) G(u) n(u) , (E.18)
where
ij
is the covariant antisymmetric tensor [40], G(u) is the Gaussian curvature
and s(u)
1

N
d
i=1
q
i
(u u
i
) is the disclination density with N
d
defects of charge
q
i
at positions u
i
. The nal result is an eective free energy whose basic degrees of
freedom are the defect positions themselves [35, 4]:
F =
K
2
_
dA
_
dA

n(u) (u, u

) n(u

) , (E.19)
where n(u), the defect density relative to the Gaussian curvature, was dened in
Eq.(E.18). The equilibrium positions of the defects are determined only by defect-
defect interactions because the Gaussian curvature is constant G =
1
R
2
on an un-
deformed sphere. To calculate the Greens function (u, u

) we need to invert the


covariant Laplacian dened on the sphere
(u, u

)
_
1

_
uu

, (E.20)
As shown below, this inversion can be accomplished by performing a weighted sum
over eigenmodes of the covariant Laplacian, [4].
Appendix E: Free energy of a vector eld on a sphere 167
We rst recall that the (generalized) Green function (u, u

) is dened by

u
(u, u

) =
(u, u

g

1
S
, (E.21)
where S = 4R
2
denotes the area of the surface and =
1

i
(

gg
ij

j
). The
presence of the second term on the left hand side of Eq.(E.21) can be understood
as follows. The Green function of the Laplacian (according to the usual denition
without the area correction in Eq.(E.21)) can be interpreted physically as the steady
temperature response of the system to a point-like unit source of heat. However, for
a closed system such as the surface of the sphere heat cannot escape. Hence, it is
impossible to impose a point source, that would inject heat at a constant rate and
have the system respond with a time-independent distribution. To prevent energy
from building up in such a system, we put the spherical surface of area S in contact
with a reservoir that uniformly absorbs heat at the same rate it is pumped in. The
need for subtracting the neutralizing background heat
1
S
in Eq.(E.21) will become
transparent mathematically once we proceed to determine (u, u

) explicitly.
The rst step consists in writing the delta function as a sum over spherical
harmonics Y
m
l
(, ),
(u, u

g

(

)(

)
R
2
sin

=
1
R
2

l=0
m=+l

m=l
Y
m
l
(, )Y
m
l

(, ) , (E.22)
and recall the eigenvalue equation
Y
m
l
(, ) =
l(l + 1)
R
2
Y
m
l
(, ) . (E.23)
Appendix E: Free energy of a vector eld on a sphere 168
Upon substituting Eq.(E.22) in Eq.(E.21) and using the eigenvalue Equation (E.23),
we can write down the Green function as
(u, u

) = R
2

l=1
m=+l

m=l
Y
m
l
(, )Y
m
l

)
l(l + 1)
. (E.24)
We have used the fact that Y
0
0
=
_
1
4
, and used the neutralizing background charge
1
S
in Eq.(E.21) to cancel out the l = 0 diverging mode.
To simplify the series in Eq.(E.24), we exploit the familiar identity [86]
m=+l

m=l
Y
m
l
(, )Y
m
l

) =
2l + 1
4
P
l
(cos ) , (E.25)
where is the angle (relative to the center of the sphere) between the directions {, }
and {

} (see also Eq.(3.36)). Upon substituting Eq.(E.25) in Eq.(E.24), we nd


(u, u

) =
m=+l

m=l
_
1
l
+
1
l + 1
_
P
l
(cos )
4
. (E.26)
The rst term of the sum in Eq.(E.24) can be simplied using the following identity
[87]
l=

l=l
P
l
(cos )
l
= ln
_
1 + sin
_

2
__
ln
_
sin
_

2
__
, (E.27)
while for the second term we substitute
l=

l=l
P
l
(cos )
l + 1
= ln
_
1 +
_
sin

2
__
ln
_
sin
_

2
__
1, (E.28)
with the result
(u, u

) =
1
4
ln
_
1 cos
2
_
+
1
4
. (E.29)
Appendix E: Free energy of a vector eld on a sphere 169
Upon dropping additive constants that do not contribute to the energy and substi-
tuting in Eq.(E.19) we obtain
F =
K
2p
2

i=j
q
i
q
j
ln [1 cos
ij
] +
N
d

i=1
q
2
i
E
c
, (E.30)
where the phenomenologically determined core energy E
c
has been added by hand
and reect the microscopic physics not captured by our long-wavelength theory.
Appendix F
Liquid crystal textures and
conformal mappings
170
Appendix F: Liquid crystal textures and conformal mappings 171
This Appendix collects a number of results from the theory of complex
variables relevant to the study of liquid crystal textures
1
. The perspective adopted
is to link the liquid crystal elasticity to the intrinsic geometry of the texture by the use
of conformal transformations. The same method provides an elegant route to nding
the ow lines of simple incompressible uids in 2D [88] and to the exact solution of
analogous problems in electromagnetism and elasticity [45].
Nematic textures in the plane in the one Frank constant approximation
can be obtained by solving Laplace equation, which is conformally invariant, and in
complex coordinates {z = x +iy, z = x iy} reads

z

z
= 0 . (F.1)
Here (x, y) (z, z) is the bond angle that the director n = (cos , sin ) forms
with respect to a xed direction, say the real axis x. The at space laplacian equation
(F.1) is also obtained by minimizing the free energy of a vector eld on a sphere (see
Appendix A) provided that a stereographic projection gauge is chosen to carry out
the calculation [12, 89]. The stereographic projection maps an arbitrary point on the
sphere R(, ) to the corresponding point z = 2Rtan
_

2
_
e
i
in the complex plane.
The metric reads [12]
g
ij
=
1
2
_
1 +
z z
4R
2
_
_
_
_
0 1
1 0
_
_
_
, (F.2)
and the components of the gauge eld in Eq.(4.3) are given by
A
z
=

A
z
=
1
2iz
_
1
z z
4R
2
1 +
z z
4R
2
_
. (F.3)
1
An inspiring coverage of complex analysis that emphasizes the geometric viewpoint is given in Ref. [63].
Appendix F: Liquid crystal textures and conformal mappings 172
In this representation of liquid crystal order on a sphere, the Frank free energy is
F =
K
4
_
d
2
z |
z
A
z
|
2
, (F.4)
and the corresponding Euler Lagrange equation is indeed Eq.(F.1), since the diver-
gence of the gauge eld is zero (
z
A
z
+
z
A
z
= 0) [12]. The stereographic projection
provides an example of how conformal transformations can be used to map physics
on an arbitrary curved surface onto simpler planar problems. This technique can
also be employed to analyze two dimensional order on surfaces of varying Gaussian
curvature (see Ref.[58, 59]).
A second use of conformal mappings is as generators of 2D liquid crystal
textures in bounded geometries or in the presence of defects. Consider an analytic
function w = f(z) that maps a grid of horizontal and vertical lines in the complex
plane z = x + iy onto a family of orthogonal curves in the w = u + iv plane that
are respectively streamlines and equipotential lines of the corresponding ow (see
Fig. F.1). Similarly, we can dene the inverse function (z) = f
1
(z) and note
that (z) = (x, y) + i(x, y) maps equipotential lines and streamlines, given by
the contour lines of (x, y) and (x, y), into a grid of vertical and horizontal lines
respectively.
The connection between liquid crystal textures and conformal mappings
rests on the following observation: if the director n(z) forms a constant angle with
respect to the streamlines (or the equipotential lines) of (z), then (z) automatically
satises Eq.(F.1) [63]. In the one Frank constant approximation, the complex nematic
Appendix F: Liquid crystal textures and conformal mappings 173
director n(z) = n
x
(x, y) +in
y
(x, y) is given (up to an arbitrary global rotation) by
2
n(z) =

(z)
|

(z)|
, (F.5)
The bond angle is readily expressed in terms of (z) via the relation
(z) = mlog [

(z)] , (F.6)
where m denotes the imaginary part of a complex number.
As an illustration consider the simple case of two disclinations on the real
axis at positions x = 1 respectively. The nematic director rotates by on a path
encircling only one defect and by 2 on a path enclosing both (see Fig. F.1). These
requirements are met by choosing the complex function (z) = arccos(z) that is
analytic everywhere except for a branch cut on the real axis from x = 1 to x =
1. The streamlines and equipotential lines are a family of hyperbolas and ellipses
with coinciding foci at x = 1; they correspond to two distinct nematic textures
dominated by splay and bend respectively. The bond angle of the director oriented
along the streamlines is easily extracted from the argument of the complex vector
eld in Eq.(F.5), with the result
(x, y) = arctan
_
y
2
+ 1 x
2

_
(y
2
+x
2
+ 1)
2
4x
2
2xy
_
. (F.7)
In this simple example, one can explicitly check that the result in Eq.(F.7) is recovered
through the more familiar route of superposing the solutions corresponding to the two
isolated defects
(x, y) =
1
2
arctan
_
y
x 1
_
+
1
2
arctan
_
y
x + 1
_
. (F.8)
2
The complex function

(z) denotes the derivative with respect to z of the complex function =


(x, y) + i(x, y) and we dene = (x, y) i(x, y).
Appendix F: Liquid crystal textures and conformal mappings 174
-2 -1 0 1 2
-2
-1
0
1
2
Figure F.1: (Color online) Splay rich (hyperbolic) and bend-rich (ellipsoidal) families
of nematic ow lines generated by two s =
1
2
disclinations. The two families of ow
lines are the equipotential and eld lines of a complex function (z) whose branch
cut is a horizontal line connecting the two defects.
The applicability of the method of conformal mappings to nding liquid
crystal textures can be justied by means of simple geometric reasoning. We start by
noting that the curl and divergence of a two dimensional vector eld v, whose stream-
lines and orthogonal trajectories are labelled by the subscripts s and p respectively,
can be expressed geometrically via the relations [63]

x
v
x
+
y
v
y
v =
s
|v| +
p
|v| . (F.9)

x
v
y

y
v
x
v =
p
|v| +
s
|v| , (F.10)
where
s
and
p
are the respective curvatures while
s
and
p
are the directional
Appendix F: Liquid crystal textures and conformal mappings 175
derivatives along the two orthogonal families of level curves
3
. For example the black
lines in Fig. F.1 trace the electric eld v generated by a uniformly charged plate or
the ow lines of an ideal uid exiting a slit of width given by the branch cut. Unlike
the liquid crystal director in Eq.(F.5), the magnitude of v is allowed to vary with
position. By construction, such a vector eld is divergence free and curl free, hence

s
=
p
log |v| , (F.11)

p
=
s
log |v| . (F.12)
By combining Equations (F.11) and (F.12), we obtain the geometric condition that a
family of equipotential lines (or streamlines) needs to satisfy in order to be identied
as level curves of an harmonic potential, namely

p
p
+

s
s
= 0 . (F.13)
This condition is entirely cast in terms of the curvatures of the equipotential lines
and streamlines without explicit reference to either the potential to be assigned or to
the magnitude of the vector eld v(z) [90, 91]. This is a natural language to discuss
orientational order in liquid crystals since the director n(z) is a vector eld of unit
magnitude.
If we take the liquid crystal director to form a constant angle with respect
to v(z), the curvatures
s
and
p
in Eq.(F.13) can be simply cast as the directional
derivatives of along the streamlines and equipotential lines respectively. In fact,
the curvature of these contour lines is the rate of change of their directions which is
3
The direction of increasing p is chosen to make a counterclockwise

2
angle with v.
Appendix F: Liquid crystal textures and conformal mappings 176
naturally parameterized by . Hence Eq.(F.13) reduces to

p
2
+

2

2
s
= 0 . (F.14)
The left hand side of Eq.(F.17) is the Laplacian of expressed in terms of orthogonal
coordinates along s and p. Since the Laplacian is coordinate independent, Eq.(F.14) is
equivalent to Eq.(F.1) and (x, y) represents (apart from an arbitrary global rotation)
the desired texture that minimizes the Frank free energy when K
1
= K
3
.
As a byproduct, Equations (F.11) and (F.12) can help to visualize how the
elastic energy stored in every portion of the texture of Fig. F.1 is distributed between
bend and splay deformations. For most liquid crystals K
3
> K
1
, so the texture with
director tangent to the streamlines will be energetically favored (black lines in Fig.
F.1). In this case, the full Frank energy density can be rewritten in terms of the local
curvatures of streamlines and equipotential lines via the simple relation
K
3
(n)
2
+K
1
( n)
2
= K
3

s
2
+K
1

p
2
. (F.15)
The energetically costly deformation involving bend takes place only around the de-
fects in a region of radius of the order of their separation. Elsewhere
s
is vanishingly
small. In contrast,
p
drops o more slowly at large distances like the inverse of
the radius of a circle centered on the midpoint between the two disclinations. Splay
deformations are present throughout the system but they have a smaller energy cost
K
1
. The converse situation occurs if K
3
< K
1
so that the texture represented by the
red equipotential lines in Fig. F.1 becomes energetically favored.
The curvatures
s
and
p
are respectively the real and imaginary parts of
the complex curvature, K(z) =
s
+i
p
, of the mapping. This quantity can be readily
Appendix F: Liquid crystal textures and conformal mappings 177
derived from the complex potential (z)
4
K(z)
s
+i
p
= i
|

2
. (F.16)
The reader is referred to the mathematical literature [90, 63] for a proof of Eq.(F.16).
The intuitive signicance of the complex curvature can be grasped by considering
how a conformal mapping f =
1
acts on a curve with local curvature at a
point in the z = x + iy plane. The curve is mapped onto an image curve in the
w = u + iv plane whose curvature

at the corresponding point diers from .


The curvature

of the image curve is determined by two mechanisms. Firstly,


the mapping f(z) locally stretches distances by a factor |f

(z)|, hence the radius of


curvature of the image curve will be naturally multiplied by this amplifying factor.
The second mechanism arises because a conformal mapping can introduce curvature
even if none was originally present (in the isothermal net) simply by locally twisting
the direction of the isothermal net by an angle equal to arg [f

(z)]
5
. The non-analytic
function K(z) controls the amount of curvature generated ex-novo by the mapping.
For example, the mapping f(z) = Cos(z) transforms a grid of horizontal lines (think
of them as a possible direction for the nematic molecules in the defect-free ground
state) into the family of hyperbolae in Fig. F.1 corresponding to a defected texture
with two +1/2 disclinations. It is not surprising that the free energy density stored in
the defected texture is simply proportional (in the one Frank constant approximation)
4
For calculational purposes, it is more convenient to recast Eq.(F.16) in the form p + is =
|

2
.
5
For this reason, the complex derivative f

(z) is sometimes called an amplitwist and encodes information


on the local eect of the mapping.
Appendix F: Liquid crystal textures and conformal mappings 178
to
_

p
_
2
+
_

s
_
2
=
2
s
+
2
p
= |K(z)|
2
, (F.17)
where the two elastic constants were set to be equal in Eq.(F.15) and the director
n(z) was parameterized in terms of the bond angle (z). The Frank free energy
is thus proportional to the complex curvature modulus-squared in analogy with the
Helfrich free energy of a membrane whose derivation rests on an higher dimensional
generalization of Eq.(F.15).
Appendix G
Vibrational spectrum of colloidal
molecules
179
Appendix G: Vibrational spectrum of colloidal molecules 180
a
b
Figure G.1: (a) The ground state of a tetratic phase exhibits eight short disclination
lines located at the vertices of a square antiprism inscribed in the sphere. (b) The
twelve disclination that characterize hexatic order s sphere lie at the vertices of an
icosahedron.
In this appendix we provide an introduction to the group theoretical treat-
ment of the vibrational spectrum of colloidal molecules. The more complicated
cases of hexatic (p=6) and tetratic (p=4) order are analyzed in some detail (see Fig.
G.1).
The starting point of the group theoretical treatment of the vibrational
spectrum of colloidal molecules is the observation that the defects displacements
from equilibrium, q (see Eq.(3.39), form the basis of a reducible representation of the
point group of the molecule. If a molecule is acted upon by a symmetry operation,
a new conguration will result in which the displacements of each defect will be
permuted and transformed
1
, but inter-defect distances and angles will be preserved.
The liquid crystal free energy (in the one Frank constant approximation) is therefore
invariant under all the operations of the point group of the colloidal defect array.
The action of each operation of the group is naturally represented by a dis-
1
Here we take the point of view that the defects themselves are not permuted only their displacements,
for example defect i may exchange its displacement coordinates qi with defect j.
Appendix G: Vibrational spectrum of colloidal molecules 181
Table G.1: Character for the irreducible representations of the icosahedral point
group together with the character of the twenty-four-dimensional representation
generated by the defect displacements.
Y E 12C
5
12C
2
5
20C
3
15C
2
A 1 1 1 1 1
F
1
3
2

1
0 1
F
2
3
1
0 1
G 4 1 1 1 0
H 5 0 0 1 1
Y 24 2
1
2 0 0
tinct 2N 2N matrix (N is the number of defects) that relates the new and old
defect positions. This representation can be completely reduced by choosing a set
of symmetry-related normal coordinates that are obtained from the original ones by
means of a linear transformation. When normal coordinates are used, the matrixes
representing the action of the symmetry group can be brought in block diagonal form
simultaneously. Energetically degenerate linear combination of the original coordi-
nates form the smallest sets invariant under application of any symmetry operation of
the group. The members of any one set generate an irreducible representation of the
group. For each point group there is only a small number of inequivalent irreducible
representations generally classied by the characters of their transformations. The
characters of the transformations are simply dened as the traces of the matrices cor-
responding to each symmetry operation and they are conveniently tabulated in most
Appendix G: Vibrational spectrum of colloidal molecules 182
texts of group theory [64, 65] (see Tables 3.1-G.2 for the character tables relevant
to the tetrahedral, icosahedral and twisted-cube shaped distributions of defects on a
sphere).
The task of nding the number of degenerate eigenmodes, n
()
, with a given
symmetry (labelled by ) reduces to counting how many times the corresponding irre-
ducible representation appears in a reducible representation. Note that the characters
of the original representation are the same as the ones of the completely reduced one
since the two dier only by a change of coordinates which preserves the trace. Thus,
the character
R
of the completely reduced representation will be the sum of the
characters of the various irreducible representations that it contains

R
=

n
()

()
R
, (G.1)
where
()
R
labels the character of the symmetry operation R in the irreducible rep-
resentation . By appealing to the orthogonality of the characters one can write an
expression for n
()
in analogy with the familiar expression for the component of a
vector along a given basis axis [64, 65]
n
()
=
1
g

()
R

R
, (G.2)
where g is the number of the symmetry operations in the group and
R
is the character
of the completely reduced representation. Eq.(G.2) is equivalent to Eq.(3.43) quoted
in the main text, as long as the sum over the group elements R is replaced by a
weighted sum over the classes in the group since the characters of group elements in
the same class are equal.
Appendix G: Vibrational spectrum of colloidal molecules 183
P
P

P
P

Figure G.2: Schematic illustration of the inversion through an equatorial plane of


symmetry (shaded circle bisecting a sphere) for a reection . The longitudinal
displacement of the defect is unchanged while the direction of the latitudinal one is
inverted.
We now adopt the analysis of Ref.[64, 65] to provide a set of rules that
produce the characters
R
of the reducible representation generated by the coordi-
nates without working out the full form of the transformation matrices. There are
two key points to notice. First only the defects located on a symmetry axis or plane
contribute to
R
the trace of the transformation matrix; defects whose displacements
are instead interchanged or permuted by the symmetry operation contribute only to
the non-diagonal terms of the matrix and hence can be ignored in determining the
character
R
. Second the directions along which the displacements from equilibrium
are measured can be chosen freely since the trace is invariant upon coordinate trans-
formations. It is generally convenient to choose them so that only one of the two
displacements components is aected by the symmetry operation.
Appendix G: Vibrational spectrum of colloidal molecules 184
As a simple example, consider a defect lying on a reection plane (see Fig.
G.2). The displacement vectors before and after the symmetry operation is applied are
{, } and {, } respectively. The resulting contribution to the character from
a single defect is thus 1 1 = 0. Inversions through a center of symmetry (coinciding
with the center of the sphere) have a vanishing contribution to
R
because there are
no defects there that can contribute to
R
.
A rotation C
k
n
by an angle
2k
n
through an n-fold axis of symmetry on which
the defect lies leads to the following transformation laws for the longitudinal and
latitudinal displacements
_
_
_
_
_
_
_

_
_
_
_
_
_
_
=
_
_
_
_
_
_
_
cos
_
2k
n
_
sin
_
2k
n
_
sin
_
2k
n
_
cos
_
2k
n
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_

i
_
_
_
_
_
_
_
. (G.3)
We measure displacements using polar coordinates with respect to the symmetry
axis; the prime denotes the orthogonal displacements after the symmetry operation
C
k
n
is applied. Inspection of Eq.(G.3) shows that the contribution from C
k
n
to the
character
R
is equal to 2 cos
_
2k
n
_
times the number of defects lying on the axis
of rotation. On the other hand the contribution to
R
from the improper rotation
S
k
n
is zero. To see this note that the symmetry operation S
k
n
is a rotary reection
achieved by performing a successive rotation through an (alternating) axis followed
by a reection in the plane perpendicular to the axis k times. An example of a
molecule possessing the symmetry operation S
4
is methane, CH
4
, with the carbon
atom lying at the intersection between an alternating axis and the reection plane
Appendix G: Vibrational spectrum of colloidal molecules 185
3
. The tetrahedral defect conguration considered in this paper does not posses any
defect at the position occupied by the carbon atom of the methane molecule. More
generally, the possibility of having a defect whose equilibrium position is unchanged
by the rotary reection is ruled out because such defect would have to lie o the
spherical surface at the intersection between the alternating axis and the plane of
reection.
To sum up, each of the characters,
R
, of the completely reduced repre-
sentation formed by the displacement coordinates is given by the number of atoms
whose equilibrium positions are not changed by the symmetry operation R times its
fundamental character as derived in the previous paragraphs
4
. The resulting charac-
ters for the tetrahedral, icosahedral and tilted cube defects congurations are listed
in Tables 3.1-G.2.
Upon using Eq.(3.43) and the character table G.1 we can decompose the
24 dimensional representation, Y , formed by the displacements from an icosahedral
equilibrium conguration into irreducible representations. The result reads
Y = 2H + 2G+ 2F
1
. (G.4)
The three rigid body rotations correspond to one of the two triplets in F
1
while the
remaining 21 independent normal coordinates form energetically degenerate multi-
plets with the following degeneracy factors: 2 quintets, 2 quartets and 1 triplet. This
analysis is conrmed upon direct diagonalization of the representation Y which leads
3
See Fig. 5-2 in Ref. [64]
4
Similar results that apply to unconstrained molecules whose atoms have three dimensional displacements
are listed in Table 6-1 of Ref. [64].
Appendix G: Vibrational spectrum of colloidal molecules 186
Table G.2: Character for the irreducible representations of the tilted cube point group
together with the character of the sixteen-dimensional representation generated by
the defect displacements.
D
4d
E 2S
8
2C
4
2S
3
8
C
2
4C
1
2
4
2
A
1
1 1 1 1 1 1 1
A
2
1 1 1 1 1 1 1
B
1
1 1 1 1 1 1 1
B
2
1 1 1 1 1 1 1
E
1
2

2 0

2 2 0 0
E
2
2 0 2 0 2 0 0
E
3
2

2 0

2 2 0 0
16 0 0 0 0 0 0
Appendix G: Vibrational spectrum of colloidal molecules 187
the 21 non-vanishing eigenvalues
i
, with the multiplicities shown bold in parenthesis

i
= {0.87(5), 0.09(5),
0.74(4), 0.22(4), 0.96(3)} . (G.5)
Note that the normal modes in the second quintet are much softer than the rest.
A similar analysis applied to the twisted cube conguration of defects leads
to the following decomposition of the (defects displacements) representation, ,
= 2E
1
+ 2E
2
+ 2E
3
+A
1
+A
2
+B
1
+B
2
. (G.6)
where the three rigid body rotations are contained in one of the two doublets E
3
and
in the singlet A
2
, which leaves ve doublets and three singlets for the eigenvalues

i
. Direct diagonalization of leads four doublets, one triplet and two singlets of
non-vanishing eigenvalues,

i
= {1.37(3), 1.31(2), 0.89(2),
0.47(2), 0.06(2), 0.11, 1.26} (G.7)
The discrepancy between the degeneracies found by direct diagonalization on one
hand and group theory on the other is caused by an accidental symmetry of the
potential energy of the tilted-cube arrangement of defects. Hence the rst triplet is
to be interpreted as the missing doublet and singlet that happen to have the same
energy even if there is no symmetry reasons to expect so. The modes in the last
doublet of Eq.(G.7) are the softest.
Appendix H
Perturbation theory of curved
crystals
188
Appendix H: Perturbation theory of curved crystals 189
The starting point of our perturbative analysis of curved crystalline order
is Eq.(5.5) which incorporates Gaussian curvature by adding an extra source to the
defect term but it is nonetheless written using a at space metric. To underscore the
subtleties involved we write a covariant generalization of the force balance equation,

ij
(x) = 0, [92]
1

x
i
(

g
i
j
)
k
ji

i
k
= 0. (H.1)
where g is the determinant of the metric tensor g
ij
and
k
ji
is the Christoel symbol.
Eq.(H.1) can be concisely written as
i
;j;i
= 0, where the semicolons indicate the
covariant derivatives D
i
[92].
Strictly speaking, this equation cannot be solved by using the at space
trick of writing the stress tensor in terms of the Airy function because of a distinctive
property of curved space: torsion [93, 40]. An arbitrary vector

T parallel transported
around a closed loop on a surface of non vanishing Gaussian curvature is rotated from
its original orientation
1
. In dierential form, this constraint reads
[D
i
, D
j
] T
k
= G(x)
ij

m
k
T
m
, (H.2)
where
ij
=

ij

g
denotes the antisymmetric tensor [40]. Upon substituting T
m
=

m
n

n
(x) into Eq.(H.2), we see that the covariant analogue of Eq.(5.3) does not
hold because the commutator of covariant derivatives (known as the torsion tensor)
does not vanish.
We can nonetheless make progress by noting rst that the Gaussian cur-
vature of a bump in Eq.(H.2) is proportional to

2
x
2
0
, so that Eq. (5.3) is in fact
1
In fact, the net eect of parallel transport is equivalent to monitoring the change of

T rst in one
direction then in the other followed by subtracting changes in the reverse order.
Appendix H: Perturbation theory of curved crystals 190
0 0.1 0.2 0.3 0.4 0.5 0.6
0
0.002
0.004
0.006
0.008
0.01
F
0
()
Yr
0
2

Figure H.1: The geometric frustration energy F


0
() (Eq. (5.3)) is plotted versus for
x
0
/R = 10/80 (_) and 10/40 (2). Solid (x
0
/R = 10/80) and dashed (x
0
/R = 10/40)
lines indicate plots of Eq. (5.3) using Y =
2

3
k [80], including the nite size corrections
discussed in Appedix I for these two conditions, respectively.
approximately correct for small . We rst consider the case S(x) = 0, for which
Eq.(5.5) reduces to one of the two celebrated Foppl-von Karman equations describing
slightly deformed thin plates [14]. For a frozen substrate, the second Foppl-von Kar-
man equation (arising from the variation of the normal coordinate h(x,y)) determines
the adhesion pressure (normal to the surface) which is needed to constrain the 2D
solid to the curved substrate [14]. The Foppl-von Karman analysis rests on a con-
sistent perturbation theory in and predicts that (x) is O(
2
), as can be seen by
applying dimensional analysis to Eq.(5.5) with S(x) = 0. To this order, the commu-
Appendix H: Perturbation theory of curved crystals 191
tator in Eq.(H.2) indeed vanishes (the leading order corrections are O(
4
)) and one
is justied in expressing
ij
(x) in terms of an Airy function (x).
The situation is more delicate in the presence of defects because (x) no
longer vanishes as 0 but tends instead to the at space form (see Ref.[2] for
explicit results). Dimensional analysis reveals that the commutator in Eq.(H.2) is
now proportional to

2
Y
x
0
b
x
0
for dislocations and to

2
Y
x
0

x
2
0
for vacancies, interstitials
and impurities where corrections of order ln
_
R
a
_
are ignored. Hence, the commutator
in Eq.(H.2), set to zero in the derivation of Eq.(5.5), appears to be of the same order
in as the curvature corrections to (x) that we wish to calculate. The commutator
is still small, however, in the continuum limit x
0
a. The use of Eq.(5.5) to study
isolated disclinations to leading order in is more dicult to justify because in this
case the commutator can be as large as

2
Y
x
0
R
x
0
. However, such an investigation on
a surface with the topology of the plane is of limited interest, in view of the large
energy cost of isolated disclinations (quadratic in R). In this work, we concentrate
primarily on the physics of dislocations, vacancies, interstitials and impurities, which
(as conrmed by our simulations) is adequately described by the formalism embodied
in Equations (C.2) and (5.5). Our analytical results are compared whenever possible
to numerical minimizations of an harmonic lattice model; these computations cor-
roborate our qualitative conclusions even beyond the limits 1 and a x
0
for
which Equations (C.2) and (5.5) are strictly valid. For example, Fig. H.1 illustrates
how our theoretical predictions of the frustration energy F
0
, based on Eq.(5.3), t
numerical data for increasingly larger values of . Similarly, rather good agreement
Appendix H: Perturbation theory of curved crystals 192
0 0.5 1 1.5 2 2.5 3 3.5
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01
0
r
D(r,/2)
Ybr
0
Figure H.2: The dislocation potential D(r, /2) in Eq.(5.11) is plotted as a continu-
ous line for a Gaussian bump parameterized by = 1. Open symbols represent the
numerical minimization of a xed connectivity harmonic model for which the sepa-
ration of the two 5-7 disclinations is xed while sliding the dislocation as a whole
radially with respect to the bump: = /2, R/x
0
=4 (blue), 8 (red); x
0
/a = 10 (_),
15 (2), 20 (), 30 (). The R-dependent nite size corrections of Appendix I are
included.
between theoretical predictions and numerics is obtained for the dislocation potential
D(r, ) even if = 1, as illustrated by Fig. H.2.
Appendix I
Curvature induced nite size
eects
193
Appendix I: Curvature induced nite size eects 194
In this appendix, we discuss how the geometric potentials derived above for
an innite system are modied by the presence of a circular boundary of radius R on
which free boundary conditions apply
n
i

ij
(R) =
1
R
_

r
_
r=R
= 0 . (I.1)
where n
i
is the unit normal to the circumference of the system. We rst determine
the Airy function
G
(r) that describes elastic deformations caused by the Gaussian
curvature G(r) only, without any contributions from the defects. As explained in
Section 5.3, once
G
(r) is known the energy of geometric frustration can be easily
calculated. We start by xing the harmonic function H
R
(x) introduced in Eq.(5.9)
upon using the boundary condition of Eq.(I.1). Note that by azimuthal symmetry,
the Gaussian curvature G(r) and hence H
R
(r) are constant on the circular boundary.
This allows to conclude that the harmonic function is constant everywhere; we denote
this constant by H
R
. The subscript indicates that the constant H
R
is a function of
the system size R that we can explicitly determine by integrating
1
Y

G
(r) = V (r) +H
R
, (I.2)
over the area of the circular disk, with the result
_
r

r
_
R
0
= Y
_
R
0
[H
R
V (r)] rdr . (I.3)
Upon substituting Equations (I.1) and (5.3) into Eq.(I.3), we obtain by integration
H
R
=

2
x
2
0
4R
2
_
e

R
x
0

2
1
_
, (I.4)
Appendix I: Curvature induced nite size eects 195
which insures that the forces on the boundary vanish. Note that
G
r
=
1
r
2

2
=
0 as a consequence of azimuthal symmetry, so the second boundary condition is
automatically satised. For large system sizes R x
0
, H
R

2
x
2
0
4R
2
.
Upon using the general denition of the Airy function, we obtain

G
kk
(r) =
G
(r) = Y V (r)

2
Y x
2
0
4R
2
. (I.5)
Substitution of Eq.(I.5) in Eq.(C.2) leads an estimate of the stretching energy, F
0
, of
the defect-free crystal that accounts for nite size eects, namely
F
0

Y
2
_
dA
_
V (r) +

2
x
2
0
4R
2
_
2
. (I.6)
The graph in Fig. H.1 is a numerical plot of F
0
versus for
R
x
0
equal to 8 and 4 that
corroborates the results of Eq.(I.6). This result is obtained from the one quoted in
the main text by simply performing the substitution V (r) V (r) H
R
in Eq.(5.3).
Note that the form of H
R
was determined by solving for
G
(r) in Eq.(I.2). Strictly
speaking, in the presence of defects one should solve the full Eq.(5.5) that accounts for
the presence of defects by means of an extra source term. The solution (x) would
not be azimuthally symmetric especially when defects are located well away from
the center of the bump. However, as detailed in Sections 5.4 and 5.7, the geometric
forces experienced by the defects at dierent locations on the bump can be easily
calculated once
G
ij
(r) is known without solving the full biharmonic equation provided
that x
0
R. The leading nite size correction to
G
ij
(r) was indeed calculated in
Eq.(I.5).
We can thus make progress in the calculation of the defect potentials in the
presence of the boundary by simply letting V (r) V (r) H
R
in Eq.(5.8) that now
Appendix I: Curvature induced nite size eects 196
reads
(x

) = Y
_
dA

S(x

)
_
dA
1

xx

[V (x) H
R
] . (I.7)
The result for the geometric potential of dislocations D(r, ) with nite size eects
(see Section 5.4) reads
D(r, ) = Y b
i

ij

j
_
dA

_
V (r

) +

2
x
2
0
4R
2
_
Y b x
0

2
8
sin
__
e
r
2
1
r
_
+
_
x
0
R
_
2
r
_
. (I.8)
The extra boundary term in D(r, ) can be viewed as the geometric potential in
an innite system (let R in Eq.(5.11)) evaluated at the position of an image
dislocation located at r

=
R
2
r
and with Burger vector

b

=

b. Thus, the disloca-
tion interacts with the curvature directly and via its image. This choice of image
insures that the stress normal to the boundary,
rr
=
1
r

r
, vanishes, similar to the
normal component of the magnetic eld generated by a wire parallel to the axis of
a surrounding cylindrical wall of innite magnetic permeability [94]. The curvature
induced boundary term included in Eq.(I.8) has no analogue in at space but its
eect is noticed by examining the results of our curved space simulations presented
in Fig. 5.1 for two dierent choices of
R
x
0
equal to 8 and 4 and aspect ratio = 0.5.
The comparison between numerics and the analytical expression derived in Eq.(I.8)
appears to conrm the validity of our approximation scheme.
We should stress, the electromagnetic analogy does not guarantee that the
second boundary condition
r
= 0 is satised when the presence of a dislocation
breaks the azimuthal symmetry. Moreover, as the dislocation approaches the bound-
Appendix I: Curvature induced nite size eects 197
0 0.5 1 1.5 2 2.5 3 3.5 4
4
3
2
1
0
1
r
D(r,/2)
Ybr
0
Figure I.1: The dislocation potential D(r, /2) is plotted versus the scaled coordinate
r = x/x
0
, x
0
= 10a, R = 40a for = 0 (_), 0.05 (), 0.1 () and 0.2 (2). The
numerical energy is scaled by 10
3
x
0
, and shifted so that D(0) 0. Notice that the
minimum is gradually lost for >
S
0.1.
ary it will be strongly attracted to it, an eect that is not captured by our formalism
and that requires solving the full boundary value problem. A systematic treatment
of boundary eects is rather convoluted even in at space and it is outside the scope
of this work. We refer the interested reader to Ref.[95]. Here we will only provide
some estimates of how the attraction to the boundary when R is nite wins out over
the geometric force in the at space limit 0. Below a critical aspect ratio
s
the
minimum of the geometric potential D(r, ) is lost, as illustrated numerically in Fig.
I.1. The aspect ratio corresponding to this spinodal point can be obtained by simple
Appendix I: Curvature induced nite size eects 198
dimensional analysis by a force or equivalently an energy balance. The geometric
potential scales like Y bx
0
whereas the attraction to the boundary is a function of the
dimensionless variable
_
x
R
_
2
and is multiplied by Y b
2
. This choice insures that the
boundary interaction is unchanged if the position of the defects is ipped x x
and suggests that the energy of the boundary interaction evaluated at the minimum
of D(r, ) is given by
Y b
2
x
2
0
R
2
(to leading order in a perturbation expansion in the small
parameter
x
0
R
). This leads to an estimate of the spinoidal aspect ratio
s

bx
0
R
which agrees with our numerical results.
Appendix J
Numerical Methods
199
Appendix J: Numerical Methods 200
We have complemented our analytical studies with numerical minimizations
of the in-plane stretching energy of a triangular lattice of points connected by springs
draped over the Gaussian bump described in the text. The discrete stretching energy
for a set of points at positions r

is dened by
F
discrete
=
1
4

,
k
,
(r

a)
2
, (J.1)
where r
,
= |r

|, r

= (x

, y

, h(x

, y

)), and a is the lattice spacing. The


height function h(x, y) denes the xed topography of the system and is given by
h(x, y) = x
0
e
(x
2
+y
2
)
2
/2x
2
0
. The spring constant matrix is k
,
= kn
,
, where the
connectivity matrix n
,
species that the underlying lattice is triangular, with a fully
coordinated particle having 6 nearest neighbors (n.n.)
n
,
=
_

_
1 if , n.n.
0 otherwise
. (J.2)
Defects are introduced into the lattice by changing the number of nearest neighbors
from 6 to 5/7 for +/- disclinations, respectively. Dislocations, interstitials and vacan-
cies are composed of specic congurations of +/- disclinations. By coarse-graining
Eq.(J.1) [80], we can make an explicit connection to the macroscopic energy formula
(5.1) with Youngs modulus, Y =
2

3
k and Poisson ratio, = 1/3, from which the
Lame coecients and can be determined [14]. In this work, we use units of energy
and length such that a = 1 and k = 1.
Numerical minimizations of Eq.(J.1) are performed for circular patches of
triangular lattice draped over a bump, with frozen-in defect congurations. Thus
every point in Figure 5.1 represents a separate energy minimization. To perform the
Appendix J: Numerical Methods 201
Table J.1: Particles constrained in energy minimizations.
Conguration Number of Particles Fixed Particles Fixed
Defect Free 1 Particle at (x,y) = (0,0)
Isolated Dislocation 2 +/- Disclinations
Two Opposing Dislocations 4 +/- Disclinations
Interstitials/Vacancies (I/V) 2 Particle at (x,y) = (0,0) and
Particle at Center of I/V
minimization, we use the standard Fletcher-Reeves (FR) conjugate-gradient method
[96, 97] in which particles are moved along successive directions determined by the
gradient of the energy in Eq.(J.1) with respect to particle coordinates. After taking
into account the fact that the z-coordinate is determined by h(x, y), the gradient with
respect to the x-coordinate of particle reads
F
x

,
k
,
2
(r

a)
r

(
,

,
)
_
(x

) + [h(x

, y

) h(x

, y

)]

x

h(x

, y

)
_
,
(J.3)
with
,
the Kronecker delta. To obtain the gradient with respect to the y-coordinate,
interchange x y.
1
Convergence is achieved when the magnitude of the gradient
of the energy drops below some dened tolerance, which was set to 10
5
(k = 1).
In this work, convergence was accepted only if the |F
discrete
| between the last two
iterations of the algorithm was less than 10
8
(k = 1).
Since there is a non-zero frustration energy for a defect-free lattice on a
1
The FR algorithm requires the use of a one-dimensional minimization algorithm, for which we used
the Brent algorithm [96] as implemented by the Gnu Scientic Library [98]. The Brent algorithm requires
bounds to be placed on the dimensionless parameter , which we typically take to be 5 < < 10.
Appendix J: Numerical Methods 202
bump, the particles will collectively slide o of the bump into atland if allowed. To
prevent this, some particles were always xed during minimization as indicated in
Table J.1. These constraints were implemented so that a particle was always located
at the center of the bump.
Bibliography
[1] P. M. Chaikin and T. C. Lubensky. Principles of Condensed Matter Physics.
Cambridge University Press, Cambridge, 1995.
[2] D. R. Nelson. Defects and Geometry in Condensed Matter Physics. Cambridge
University Press, Cambridge, 2002.
[3] A. R. Bausch, M. J. Bowick, A. Cacciuto, A. D. Dinsmore, M. F. Hsu, D. R.
Nelson, M. G. Nikolaides, A. Travesset, and D. A. Weitz. Science, 299:1716,
2003.
[4] A. Travesset M. Bowick, D. R. Nelson. Phys. Rev. B, 62:8738, 2000.
[5] J. Heuser. J. Cell. Biol., 108:401, 1989.
[6] D. M. Kroll T. Kohyama and G. Gompper. Europhys. Lett., 68:061905, 2003.
[7] S. Schneider and G. Gompper. Europhys. Lett., 70:136, 2005.
[8] B. K. Ganser, S. Li, V. Y. Klishko, J. T. Finch, and W. I. Sundquist. Nature,
407:409, 2000.
[9] S. Li, C. P. Hill, Sundquist W. I., and J. T. Finch. Nature, 407:409, 2000.
203
Bibliography 204
[10] P. Lipowsky and al. Nat. Mat., 4:407, 2005.
[11] A. Fernandez-Nieves. private communication.
[12] T. C. Lubensky and J. Prost. J. Phys. II (France), 2:371, 1992.
[13] D. R. Nelson. Nano Lett., 2:11252, 2002.
[14] L. D. Landau and E. M. Lifshitz. Theory of Elasticity. Butterworth, Oxford,
1999.
[15] S. Sachdev and D. R. Nelson. J. Phys. C., 17:5473, 1984.
[16] D. R. Nelson. Phys. Rev. B, 27:2902, 1983.
[17] D. R. Nelson V. Ambegaokar, B. I. Halperin and E. D. Siggia. Phys. Rev. B,
21:1806, 1980.
[18] V. Vitelli A. Hexemer, E. Kremer and D. R. Nelson. American Physical Society
March Meeting, poster session, Baltimore 2006.
[19] D. R. Nelson and B. I. Halperin. Phys. Rev. B, 19:2457, 1979.
[20] J. M. Kosterlitz and D. J. Thouless. J. Phys. C., 6:1181, 1973.
[21] A. P. Young. Phys. Rev. B, 19:1855, 1979.
[22] C. F. Chou, A. J. Jin, S. W. Hui, C. C. Huang, and J. T. Ho. Science, 280:1424,
1998.
[23] C. Knobler and R. Desai. Ann. Rev. Phys. Chem., 43:207, 1992.
Bibliography 205
[24] R. Seshadri and R. M. Westervelt. Phys. Rev. Lett., 66:2774, 1991.
[25] C. C. Grimes and G. Adams. Phys. Rev. Lett., 42:795, 1979.
[26] E. Y. Andrei D. C. Glattli and F. I. B. Williams. Phys. Rev. Lett., 60:420, 1998.
[27] G. Deville, A. Valdes, E. Y. Andrei, and F. I. B. Williams. Phys. Rev. Lett.,
53:588, 1984.
[28] C. M. Murray. in Bond Orientational Order in Condensed Matter Systems, edited
by K. J. Strandburg, (Springer, Berlin, 1992).
[29] R. Lenke K. Zahn and G. Maret. Phys. Rev. Lett., 82:2721, 1999.
[30] R. A. Segalman, A. Hexemer, R. C. Hayward, and E. J. Kramer. Macromolecules,
36:3272, 2003.
[31] D. R. Nelson. Phys. Rev. B, 28:5515, 1983.
[32] F. David, E. Guitter, and L. Peliti. J. Phys., 48:2059, 1987.
[33] D. R. Nelson and L. Peliti. J. Phys., 48:1085, 1987.
[34] E. Guitter and M. Kardar. Europhys. Lett., 13:441, 1990.
[35] J. M. Park and T. C. Lubensky. Phys. Rev. E, 53:2648, 1996.
[36] P. Lenz and D. R. Nelson. Phys. Rev. E, 67:031502, 2003.
[37] E. Kramer. private communication.
[38] P. G. de Gennes and J. Prost. The Physics of Liquid Crystals. Clarendon Press,
Oxford, 1993.
Bibliography 206
[39] M. Bowick, D. R. Nelson, and A. Travesset. Phys. Rev. E (in press).
[40] F. David. in Statistical Mechanics of Membranes and Surfaces, edited by D. R.
Nelson et al., (World Scientic, Singapore, 2004).
[41] D. J. Struik. Lectures on Classical Dierential Geometry. Dover, New York,
1961.
[42] R. Kamien. Rev. Mod. Phys., 74:953, 2002.
[43] V. Vitelli and A. M. Turner. Phys. Rev. Lett., 93:215301, 2005.
[44] B. A. Dubrovin, A. T. Fomenko, and S. P. Novikov. Modern Geometry- Methods
and Applications, volume 1. Springer-Verlag, New York, 1992.
[45] V. I. Smirnov. A Course of Higher Mathematics, volume 3 part 2. Pergamon
Press, Oxford, 1964.
[46] A. M. Polyakov. 103:307, 1981.
[47] M. Kleman and O. D. Lavrentovich. Soft Matter Physics. Springer-Verlag, New
York, 2003.
[48] W. K. H. Panofsky and M. Phillips. Classical Electricity and Magnetism.
Addison-Wesley, Reading, 1962.
[49] W. F. Vinen. in Superconductivity, vol. 2, edited by R. D. Parks, (Dekker, New
York, 1969).
[50] G. B. Hess. Phys. Rev., 161:189, 1967.
Bibliography 207
[51] M. Kleman. Points, Lines and walls. Wiley, New York, 1983.
[52] R. D. Kamien. Science, 299:1671, 2003.
[53] D. R. Nelson. Nano Lett., 2:1125, 2002.
[54] Park and al. Science, 276:1401, 1997.
[55] O. D. Lavrentovich. Liq. Cryst., 24:117, 1998.
[56] G. P. Crawford and S. Zumer. Liquid Crystals in Complex Geometries. Francis
and Taylor, London, 1996.
[57] F. C. MacKintosh and T. C. Lubensky. Phys. Rev. Lett., 67:1169, 1991.
[58] V. Vitelli and D. R. Nelson. Phys. Rev. E, 70:051105, 2004.
[59] V. Vitelli and A. M. Turner. Phys. Rev. Lett., 93:215301, 2005.
[60] S. Weinberg. Gravitation and Cosmology. Wiley, New York, 1972.
[61] S. Chandrasekhar. Liquid Crystals. Cambridge University Press, Cambridge,
1994.
[62] F. Dyson. private communication.
[63] T. Needham. Visual Complex Analysis. Oxford University Press, Oxford, 1997.
[64] E. B. Wilson, J. C. Decius, and P. C. Cross. Molecular Vibrations. Dover, New
York, 1955.
[65] E. M Lifshitz and L. D. Landau. Quantum Mechanics: Non-Relativistic Theory.
Butterworth-Heinemann, Oxford, 1977.
Bibliography 208
[66] C. Chiccoli, I. Feruli, O. D. Lavrentovich, P. Pasini, S. V. Shyianovskii, and
C. Zannoni. Phys. Rev. E, 66:030701, 2002.
[67] H. Stark. Phys. Rep., 351:387, 2001.
[68] T. Lubensky, D. Pettey, N. Currier, and H. Stark. Phys. Rev. E, 57:610, 1998.
[69] P. Poulin, H. Stark, T. C. Lubensky, and D. A. Weitz. Science, 275:1770, 1997.
[70] D. Caspar and A. Klug. Principles in the Construction of Regular Viruses, Cold
Spring Harbor Symposium on Quantitative Biology, 27, 1, (1962).
[71] D. R. Nelson and R. A. Pelcovits. Phys. Rev. B, 16:2191, 1977.
[72] D. R. Nelson and J. M. Kosterlitz. Phys. Rev. Lett., 39:1201, 1977.
[73] M. Bowick, D. R. Nelson, and A. Travesset. Phys. Rev. B, 62, 8378, (2000) and
Phys. Rev. E, 69, 041102, (2004).
[74] V. Vitelli and D. R. Nelson. cond-mat/0406328 .
[75] B. I. Halperin. private communication.
[76] D. R. Nelson. private communication.
[77] E. J. Yarmchuk and R. E. Packard. J. Low Temp. Phys., 46:479, 1982.
[78] E. J. Kramer. Nature, 437:824, 2005.
[79] N. Toan, Bruinsma R. F., and W. M. Gelbart.
http://arxiv.org/abs/physics/0506127 .
Bibliography 209
[80] S. Seung and D. R. Nelson. 38:1005, 1988.
[81] R. Bruinsma, B. I. Halperin, and A. Zippelius. Phys. Rev. B, 25:579, 1982.
[82] S. Jain and D. R. Nelson. Phys. Rev. E, 61:1599, 2000.
[83] M. Spivak. A Comprehensive Introduction to Dierential Geometry. Publish or
Perish,Inc., Berkeley, 1979.
[84] A. Budzin and D. Feinberg. Phys. C, 235:2755, 1994.
[85] A. J. McConnell. Applications of Tensor Analysis. Dover Publications, New
York, 1957. pp. 184-189.
[86] H. W. Wyld. Mahtematical Methods for Physics. Perseus Books, Reading, 1999.
pp. 272-276.
[87] I. S. Gradstein and I. M. Ryshik. Tables of series, products and integrals. Verlag
Harri Deutsch, Frankfurt, 1981.
[88] R. P. Feynman, R. B. Leighton, and M. Sands. The Feynman Lectures on Physics,
volume 2. Addison-Wesley, Reading, 1963.
[89] B. A. Ovrut and S. Thomas. Phys. Rev. D, 43:1314, 1991.
[90] I. C. Bivens. Math. Mag., 65:226, 1992.
[91] T. Needham. Math. Mag., 67:92, 1994.
[92] J. H. Heinbockel. Introduction to Tensor Calculus and Continuum Mechanics.
Traord, 2001.
Bibliography 210
[93] J. L. van Hemmen and M. A. Peterson. cond-mat/0503158 .
[94] J. Vanderlinde. Classical Electromagnetic Theory. Butterworth, Oxford, 1993.
[95] J. S.Koehler. Phys. Rev., 60:397, 1941.
[96] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery. Numerical
Recipes in C++ (Second Edition). Cambridge University Press, Cambridge,
2002.
[97] E. Polak. Computational Methods in Optimization. Academic Press, New York,
1971.
[98] M. Galassi, J. Davies, J. Theiler, B. Gough, G. Jungman, M. Buth, and F. Rossi.
Gnu Scientic Library Reference Manual. Network Theory Ltd., 2003.

You might also like