17.01.2013 Views

Crystals, Liquid Crystals and Superfluid Helium on Curved Surfaces.

Crystals, Liquid Crystals and Superfluid Helium on Curved Surfaces.

Crystals, Liquid Crystals and Superfluid Helium on Curved Surfaces.

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<str<strong>on</strong>g>Crystals</str<strong>on</strong>g>, <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> <str<strong>on</strong>g>Crystals</str<strong>on</strong>g> <str<strong>on</strong>g>and</str<strong>on</strong>g> <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> <str<strong>on</strong>g>Helium</str<strong>on</strong>g> <strong>on</strong> <strong>Curved</strong><br />

<strong>Surfaces</strong>.<br />

A dissertati<strong>on</strong> presented<br />

by<br />

Vincenzo Vitelli<br />

to<br />

The Department of Physics<br />

in partial fulfillment of the requirements<br />

for the degree of<br />

Doctor of Philosophy<br />

in the subject of<br />

Physics<br />

Harvard University<br />

Cambridge, Massachusetts<br />

May 2006


c○2006 - Vincenzo Vitelli<br />

All rights reserved.


Thesis advisor Author<br />

David R. Nels<strong>on</strong> Vincenzo Vitelli<br />

<str<strong>on</strong>g>Crystals</str<strong>on</strong>g>, <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> <str<strong>on</strong>g>Crystals</str<strong>on</strong>g> <str<strong>on</strong>g>and</str<strong>on</strong>g> <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> <str<strong>on</strong>g>Helium</str<strong>on</strong>g> <strong>on</strong> <strong>Curved</strong> <strong>Surfaces</strong>.<br />

Abstract<br />

In this thesis we study the ground state of ordered phases grown as thin layers <strong>on</strong><br />

substrates with smooth spatially varying Gaussian curvature. The Gaussian curvature<br />

acts as a source for a <strong>on</strong>e body potential of purely geometrical origin that c<strong>on</strong>trols the<br />

equilibrium distributi<strong>on</strong> of the defects in liquid crystal layers, thin films of He 4 <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

two dimensi<strong>on</strong>al crystals <strong>on</strong> a frozen curved surface. For superfluids, all defects are<br />

repelled (attracted) by regi<strong>on</strong>s of positive (negative) Gaussian curvature. For liquid<br />

crystals, charges between 0 <str<strong>on</strong>g>and</str<strong>on</strong>g> 4π are attracted by regi<strong>on</strong>s of positive curvature while<br />

all other charges are repelled. As the thickness of the liquid crystal film increases,<br />

transiti<strong>on</strong>s between two <str<strong>on</strong>g>and</str<strong>on</strong>g> three dimensi<strong>on</strong>al defect structures are triggered in the<br />

ground state of the system. Thin spherical shells of nematic molecules with planar<br />

anchoring possess four short 1 disclinati<strong>on</strong> lines but, as the thickness increases, a three<br />

2<br />

dimensi<strong>on</strong>al escaped c<strong>on</strong>figurati<strong>on</strong> composed of two pairs of half-hedgehogs becomes<br />

energetically favorable. Finally, we examine the static <str<strong>on</strong>g>and</str<strong>on</strong>g> dynamical properties that<br />

distinguish two dimensi<strong>on</strong>al crystals c<strong>on</strong>strained to lie <strong>on</strong> a curved substrate from<br />

their flat space counterparts. A generic mechanism of dislocati<strong>on</strong> unbinding in the<br />

presence of varying Gaussian curvature is presented. We explore how the geometric<br />

potential affects the energetics <str<strong>on</strong>g>and</str<strong>on</strong>g> dynamics of dislocati<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> point defects such<br />

as vacancies <str<strong>on</strong>g>and</str<strong>on</strong>g> interstitials.<br />

iii


C<strong>on</strong>tents<br />

Title Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i<br />

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii<br />

Table of C<strong>on</strong>tents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv<br />

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi<br />

Dedicati<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii<br />

1 Introducti<strong>on</strong> 1<br />

1.1 Introducti<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.1.1 Experimental motivati<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . 2<br />

1.1.2 Mechanisms of coupling between defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature . . . . 5<br />

1.1.3 Cross-over between 2D <str<strong>on</strong>g>and</str<strong>on</strong>g> 3D physics . . . . . . . . . . . . . 13<br />

1.1.4 Outline of the thesis . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2 B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 16<br />

2.1 Introducti<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2 Hexatic order <strong>on</strong> a surface . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.1 Electrostatic analogy . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.2 Defect free texture . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.2.3 Energetics of defect pairs <strong>on</strong> a Gaussian bump . . . . . . . . . 30<br />

2.3 Curvature induced defect generati<strong>on</strong> . . . . . . . . . . . . . . . . . . 35<br />

2.3.1 Onset of the defect-dipole instability . . . . . . . . . . . . . . 35<br />

2.3.2 Numerical investigati<strong>on</strong> of defect-unbinding transiti<strong>on</strong>s . . . . 39<br />

2.3.3 Single vortex instability . . . . . . . . . . . . . . . . . . . . . 44<br />

2.3.4 Lattice of bumps, valleys <str<strong>on</strong>g>and</str<strong>on</strong>g> saddle points . . . . . . . . . . . 48<br />

2.4 Defect Dec<strong>on</strong>finement . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

2.5 C<strong>on</strong>clusi<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3 <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 60<br />

3.1 Introducti<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.2 Textures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.2.1 Tilted molecules <strong>on</strong> a sphere . . . . . . . . . . . . . . . . . . . 66<br />

3.2.2 Nematic Texture . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

3.3 Stability of liquid crystal textures to thermal fluctuati<strong>on</strong>s . . . . . . 80<br />

iv


C<strong>on</strong>tents v<br />

3.4 Valence transiti<strong>on</strong>s in thick nematic shells . . . . . . . . . . . . . . . 88<br />

3.4.1 Slab geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

3.4.2 Extrapolati<strong>on</strong> to thin spherical shells . . . . . . . . . . . . . . 94<br />

3.5 C<strong>on</strong>clusi<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97<br />

4 <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 99<br />

5 Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 112<br />

5.1 Introducti<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

5.2 Basic Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

5.3 Geometric Frustrati<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

5.4 Geometric potential for dislocati<strong>on</strong>s . . . . . . . . . . . . . . . . . . 119<br />

5.5 Dislocati<strong>on</strong> unbinding <str<strong>on</strong>g>and</str<strong>on</strong>g> Grain Boundaries . . . . . . . . . . . . . 123<br />

5.6 Glide suppressi<strong>on</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

5.7 Vacancies, Interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> Impurities . . . . . . . . . . . . . . . . . 127<br />

A Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Isothermal Coordinates 131<br />

B Geometric Potential 137<br />

C Free energy <strong>on</strong> a corrugated plane 142<br />

D Bump with a boundary 150<br />

E Free energy of a vector field <strong>on</strong> a sphere 162<br />

F <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 170<br />

G Vibrati<strong>on</strong>al spectrum of colloidal molecules 179<br />

H Perturbati<strong>on</strong> theory of curved crystals 188<br />

I Curvature induced finite size effects 193<br />

J Numerical Methods 199<br />

Bibliography 203


Acknowledgments<br />

This work originates from the inspiring mentorship of Professor D. R. Nels<strong>on</strong><br />

to whom goes my lasting gratitude for patiently guiding my first steps in scientific<br />

research while always encouraging independence of thought.<br />

I had the privilege to have Professors B. I. Halperin <str<strong>on</strong>g>and</str<strong>on</strong>g> D. A. Weitz <strong>on</strong><br />

my thesis committee <str<strong>on</strong>g>and</str<strong>on</strong>g> I wish to thank them for enlightening interacti<strong>on</strong>s. As a<br />

beginning graduate student, I attended some stimulating courses taught by Professors<br />

S. Sachdev <str<strong>on</strong>g>and</str<strong>on</strong>g> D. S. Fisher that greatly fostered my interest in c<strong>on</strong>densed matter<br />

theory. One of the most rewarding aspects of my PhD experience at Harvard has been<br />

the unique chance to informally interact with a number of very talented graduate<br />

students <str<strong>on</strong>g>and</str<strong>on</strong>g> postdoctoral fellows from whom I learned a lot <str<strong>on</strong>g>and</str<strong>on</strong>g> shared moments of<br />

fun: M. Desai, A. Delmaestro, A. Imambekov, I. Finkler, W. P. W<strong>on</strong>g, R. Barnett, D.<br />

Podolsky, H. Chen, O. White, K. Papadodimas, E. Katifori, G. Rafael, A. Polkolnikov,<br />

R. da Silveira, Y. Kafri, J. Qiang, C. Kilic <str<strong>on</strong>g>and</str<strong>on</strong>g> P. J. Lu. A special thanks goes to A.<br />

Fern<str<strong>on</strong>g>and</str<strong>on</strong>g>ez-Nieves, A. M. Turner <str<strong>on</strong>g>and</str<strong>on</strong>g> J. B. Lucks with whom I collaborated <strong>on</strong> the<br />

work discussed in chapters I, IV <str<strong>on</strong>g>and</str<strong>on</strong>g> V respectively.<br />

I wish to thank the administrative staff of the Physics Department in par-<br />

ticular Dr. D. Norcross <str<strong>on</strong>g>and</str<strong>on</strong>g> S. Fergus<strong>on</strong> for providing c<strong>on</strong>stant help.<br />

As an undergraduate at Imperial College L<strong>on</strong>d<strong>on</strong>, I have been fortunate<br />

to meet many inspiring lecturers that I wish to thank for their crucial help <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

generosity particularly Professors M. P. Blencowe, J. P. C<strong>on</strong>nerade, P. L. Knight, A.<br />

McKinn<strong>on</strong>, M. B. Plenio <str<strong>on</strong>g>and</str<strong>on</strong>g> D. Vvedensky. An undergraduate bursary from the<br />

Nuffield foundati<strong>on</strong> allowed me to try research first h<str<strong>on</strong>g>and</str<strong>on</strong>g> <str<strong>on</strong>g>and</str<strong>on</strong>g> to spend a stimulating<br />

vi


Acknowledgments vii<br />

semester as a visiting student working under the supervisi<strong>on</strong> of Dr. T. S. Chang at<br />

MIT.<br />

I wish to thank my family <str<strong>on</strong>g>and</str<strong>on</strong>g> A. Alyahya for their support throughout.


To d<strong>on</strong> Rocco, to my gr<str<strong>on</strong>g>and</str<strong>on</strong>g>parents.<br />

viii


Chapter 1<br />

Introducti<strong>on</strong><br />

1


Chapter 1: Introducti<strong>on</strong> 2<br />

1.1 Introducti<strong>on</strong><br />

The physics of 2D c<strong>on</strong>densed matter systems is a rich <str<strong>on</strong>g>and</str<strong>on</strong>g> mature subject<br />

[1, 2]. In this thesis we study the effects induced in the ground state of a thin layer<br />

of material by the Gaussian curvature of the substrate <strong>on</strong> which it is deposited. Spe-<br />

cific examples include liquid crystalline order at the curved interface between two<br />

immiscible fluids, thin films of He 4 wetting a corrugated substrate <str<strong>on</strong>g>and</str<strong>on</strong>g> two dimen-<br />

si<strong>on</strong>al crystals <str<strong>on</strong>g>and</str<strong>on</strong>g> liquid crystals <strong>on</strong> a lithographically prepared curved substrate.<br />

While these systems are physically distinct, some of the c<strong>on</strong>sequences of the varying<br />

Gaussian curvature of the substrate or interface can be understood within a unifying<br />

perspective. It is the purpose of this introducti<strong>on</strong> to illustrate basic c<strong>on</strong>cepts <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

suggest possible experimental tests of the theoretical ideas presented in this thesis.<br />

1.1.1 Experimental motivati<strong>on</strong><br />

Recent experimental investigati<strong>on</strong>s of spherical crystallography have already<br />

revealed complex ground state structures riddled with topological defects forced in by<br />

the topology of the underlying curved space [3]. Fig. 1.1 shows a colloidosome engi-<br />

neered in the Weitz lab with 0.5 µm colloidal particles self assembled <strong>on</strong> an oil-water<br />

interface. Besides its intriguing technological applicati<strong>on</strong>s as a microcapsule for drug<br />

delivery, the colloidosome provides a natural laboratory to explore (under c<strong>on</strong>trolled<br />

experimental c<strong>on</strong>diti<strong>on</strong>s) basic scientific questi<strong>on</strong>s of curved space crystallography rel-<br />

evant to a larger class of materials including those of biological origin such as viruses<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> clathrin-coated pits [5, 6, 7, 8, 9]. C<strong>on</strong>focal microscopy of small colloidosomes


Chapter 1: Introducti<strong>on</strong> 3<br />

Figure 1.1: Image of a colloidosome by c<strong>on</strong>focal microscopy (courtesy of the Weitz<br />

lab). The right panel shows the arrangement of 5 <str<strong>on</strong>g>and</str<strong>on</strong>g> 7-fold disclinati<strong>on</strong>s, represented<br />

as red <str<strong>on</strong>g>and</str<strong>on</strong>g> yellow dots respectively, in the ground state of a spherical crystal<br />

(reproduced from Ref.[4]).<br />

reveals 12 five-fold coordinated lattice sites symmetrically placed approximately at<br />

the vertices of an icosahedr<strong>on</strong> inscribed in the sphere. When the droplet radius R<br />

is larger than 5-6 times the colloidal radius, the nucleati<strong>on</strong> of grain boundary scars<br />

emanating from each disclinati<strong>on</strong> becomes energetically favorable [4, 10].<br />

<str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal (LC) analogs of these systems can be realized by trapping a<br />

thin layer of thermotropic liquid crystals in a double emulsi<strong>on</strong> [11]. Fig. 1.2 shows<br />

four s = 1<br />

2<br />

disclinati<strong>on</strong> lines visible under crossed polarizers as double-brushes, which<br />

c<strong>on</strong>fer to the colloidal particle an effective valence Z = 4 [12, 13]. The disclinati<strong>on</strong>s in<br />

the droplet represented in Fig. 1.2 do not lie exactly at the vertices of a tetrahedr<strong>on</strong><br />

because of the n<strong>on</strong> uniform thickness of the layer. An alternative route to colloids<br />

with a valence is provided by m<strong>on</strong>o-layers of lyotropic nematogens [11]. The main<br />

technological thrust behind these experiments is the desire to functi<strong>on</strong>alize droplets<br />

or colloids using chemical linkers. Such a tetravalent functi<strong>on</strong>alizati<strong>on</strong> could be a<br />

preliminary step towards the self assembly of a diam<strong>on</strong>d lattice of colloids which has


Chapter 1: Introducti<strong>on</strong> 4<br />

Figure 1.2: The left panel shows an image of a double emulsi<strong>on</strong> engineered (via<br />

microfluidics) by trapping a ∼ 8 µm nematic layer between two spherical water interfaces<br />

(picture courtesy of Alberto Fern<str<strong>on</strong>g>and</str<strong>on</strong>g>ez-Nieves). The radius of the drop is<br />

approximately 100 µm. Under crossed polarizers each of the four disclinati<strong>on</strong>s with<br />

topological charge s = 1<br />

2<br />

appears as a double brush. The right panel presents a<br />

schematic illustrati<strong>on</strong> of the nematic texture in the m<strong>on</strong>o-layer limit with the disclinati<strong>on</strong>s<br />

at the vertices of a tetrahedr<strong>on</strong>. The core of the disclinati<strong>on</strong>s are indicated<br />

as black dots to which chemical linkers (drawn as wiggly lines emanating from the<br />

defects) can be attached by self assembly.<br />

c<strong>on</strong>siderable potential for phot<strong>on</strong>ic b<str<strong>on</strong>g>and</str<strong>on</strong>g> gap materials [13].<br />

To characterize ground states in such systems, it is often c<strong>on</strong>venient to switch<br />

from a c<strong>on</strong>tinuum descripti<strong>on</strong> of the basic degrees of freedom (e.g. the angle that<br />

indicates the local orientati<strong>on</strong> of the LC molecules or the displacements of individual<br />

atoms in the crystal) to an effective theory parameterized by the positi<strong>on</strong>s of a few<br />

topological defects. As an example, the LC texture in Fig. 1.2 can be understood as<br />

a result of the electrostatic-like repulsi<strong>on</strong> between the four disclinati<strong>on</strong>s (see chapter<br />

4 for more details).


Chapter 1: Introducti<strong>on</strong> 5<br />

FIG. 2: A surface with negative intrinsic curvature. The two principal curvatures are denoted by κ1 <str<strong>on</strong>g>and</str<strong>on</strong>g> κ2 <str<strong>on</strong>g>and</str<strong>on</strong>g> their product,<br />

Figure 1.3: A surface with negative Gaussian curvature. The two principal curvatures<br />

(with dimensi<strong>on</strong>s of an inverse length) are denoted by κ1 <str<strong>on</strong>g>and</str<strong>on</strong>g> κ2; their product is the<br />

Gaussian curvature G(x) = κ1κ2, which is negative in this example.<br />

the intrinsic curvature, is negative (reproduced from Ref. [1]).<br />

defect of charge q <str<strong>on</strong>g>and</str<strong>on</strong>g> the background curvature distributi<strong>on</strong> is quadratic in q, hence both vortices <str<strong>on</strong>g>and</str<strong>on</strong>g> anti-vortices<br />

are repelled from the top of the bump.<br />

It is our hope to test the existence of this interacti<strong>on</strong> in He films by studying how it competes against the familiar<br />

1.1.2 Mechanisms of coupling between defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature<br />

c<strong>on</strong>fining potential generated by the rotati<strong>on</strong> of the bump. This geometric force exists in a variety of ordered phases<br />

whose defects interact like a Coulomb gas <str<strong>on</strong>g>and</str<strong>on</strong>g> provides an unexpected coupling between matter <str<strong>on</strong>g>and</str<strong>on</strong>g> geometry.<br />

Defects in two dimensi<strong>on</strong>al m<strong>on</strong>olayers can be modeled as point-like particles<br />

II. BASIC THEORETICAL IDEAS<br />

that live in the local tangent plane to the surface <str<strong>on</strong>g>and</str<strong>on</strong>g> cannot escape in the third<br />

A defect at positi<strong>on</strong> r will feel a total potential energy, E(r), given by<br />

dimensi<strong>on</strong>. The defects are sensitive to the Gaussian curvature of the substrate (see<br />

E(r)<br />

= V (r) + A(r)<br />

+ c , (3)<br />

λ2 � 2 ρ<br />

m 2<br />

Fig. 1.3), the latter being an intrinsic property of space that can be probed without<br />

where m is the mass of 4He, ρ is the superfluid density <str<strong>on</strong>g>and</str<strong>on</strong>g> c is an arbitrary c<strong>on</strong>stant. The length λ is defined as<br />

�<br />

ever leaving the surface. Experiments �<br />

λsuch ≡ as those . initiated in the Weitz lab can be (4)<br />

mω<br />

described fairly well by approximating the droplets as spheres of c<strong>on</strong>stant positive<br />

The first c<strong>on</strong>tributi<strong>on</strong> to E(r), �2ρ m2 V (r), accounts for the interacti<strong>on</strong> of the defect with the curvature (see Fig. 3). The<br />

sec<strong>on</strong>d c<strong>on</strong>tributi<strong>on</strong> is �ρω<br />

m A(r), where A(r) is the area within a cup of radius r <str<strong>on</strong>g>and</str<strong>on</strong>g> it is multiplied by the number<br />

density Gaussian <str<strong>on</strong>g>and</str<strong>on</strong>g> thecurvature. energy quantum A �ω. significant This last term porti<strong>on</strong> generalizes of this (tothesis curved space) addresses the familiar the less parabolic restrictive potential in<br />

flat space that tries to c<strong>on</strong>fine the defect at the center of the bump as a result of the rotati<strong>on</strong> (see Fig. 4). As <strong>on</strong>e varies<br />

α a sec<strong>on</strong>d order transiti<strong>on</strong> occurs. In fact, Fig. 5 reveals that for α greater than a critical value αc the total energy<br />

E(r) case assumes of c<strong>on</strong>densed a mexican hat matter shape whose order minima <strong>on</strong> aissubstrate offset with respect of varying to the top Gaussian of the bump. curvature. Taking a derivative See<br />

of Eq.(4) with respect to r leads to an implicit equati<strong>on</strong> for the positi<strong>on</strong> of the minimum, rm, (or maximum):<br />

Figures 1.3 <str<strong>on</strong>g>and</str<strong>on</strong>g> 1.4 for examples.<br />

rm<br />

λ = sin(θ[rm]) . (5)<br />

The varying Gaussian curvature of the substrate acts as a source for a <strong>on</strong>e<br />

where θ(r) is the angle that the tangent at r to the bump forms with the xy plane in Fig. 1a. A simple graphical<br />

c<strong>on</strong>struct allows to solve Eq.(5) by finding the intercept(s) of the sine curve <strong>on</strong> the RHS with the straight line of slope<br />

2


Chapter 1: Introducti<strong>on</strong> 6<br />

body potential V (�x) of purely geometrical origin that c<strong>on</strong>trols the equilibrium dis-<br />

tributi<strong>on</strong> of the defects in the ground state <str<strong>on</strong>g>and</str<strong>on</strong>g> poses c<strong>on</strong>straints <strong>on</strong> their dynamics.<br />

These geometric interacti<strong>on</strong>s depend crucially <strong>on</strong> the scalar or vector topological<br />

charge of the defect <str<strong>on</strong>g>and</str<strong>on</strong>g> <strong>on</strong> the symmetry of the order parameter. The defect poten-<br />

tial is a n<strong>on</strong> local functi<strong>on</strong> of the Gaussian curvature that can be explicitly determined<br />

using the ideas <str<strong>on</strong>g>and</str<strong>on</strong>g> methods discussed in this thesis. To make the ideas more c<strong>on</strong>-<br />

crete, I shall compare the geometric potential felt by vortices in superfluid 4 He <strong>on</strong><br />

curved surfaces to the <strong>on</strong>e experienced by vacancies in crystalline solids grown <strong>on</strong><br />

a curved substrate. The former are topological defects that introduce l<strong>on</strong>g distance<br />

disturbances in the superfluid flow lines. The latter are point defects that arise from<br />

locally subtracting <strong>on</strong>ly <strong>on</strong>e atom from the curved space solid. Despite their obvious<br />

differences these two objects experience the same geometric potential V (x) that is<br />

given (apart from multiplicative c<strong>on</strong>stants) by the soluti<strong>on</strong> of a Poiss<strong>on</strong>-like equati<strong>on</strong><br />

∆V (x) = −G(x) (1.1)<br />

where ∆ is the Laplacian <str<strong>on</strong>g>and</str<strong>on</strong>g> the Gaussian curvature G(x) plays the role of an elec-<br />

trostatic like background charge. For the bump-like deformati<strong>on</strong> shown in Fig. 1.4,<br />

the qualitative form of V (x) in Eq.(1.1) can be guessed by appropriately generalizing<br />

Gauss’s law of electrostatics. (An indented, dimple-like surface is equivalent within<br />

our model to the <strong>on</strong>e of Fig. 1.4 because the Gaussian curvature is the same).<br />

Provided there is circular symmetry, the field −∇V (x) (proporti<strong>on</strong>al to the<br />

force <strong>on</strong> a defect) can be determined as in electrostatics from integrating G(x) over<br />

a circular patch centered around the top of the bump <str<strong>on</strong>g>and</str<strong>on</strong>g> with radius given by the


Chapter 1: Introducti<strong>on</strong> 7<br />

(a)<br />

x<br />

r<br />

Figure 1.4: (a) A surface with a bump-like deformati<strong>on</strong>. (b) Top view of (a) showing<br />

a schematic representati<strong>on</strong> of the positive <str<strong>on</strong>g>and</str<strong>on</strong>g> negative Gaussian curvature as a n<strong>on</strong>uniform<br />

background ”charge” distributi<strong>on</strong> that switches sign at a characteristic radius<br />

r = r0 proporti<strong>on</strong>al to the size of the bump. The varying density of + <str<strong>on</strong>g>and</str<strong>on</strong>g> - signs is<br />

intended to mimic the changing Gaussian curvature in the vicinity of the bump.<br />

positi<strong>on</strong> of the defect. This integral is always positive <str<strong>on</strong>g>and</str<strong>on</strong>g> vanishes <strong>on</strong>ly at infinity<br />

because the Gaussian bump is topologically equivalent to the plane. Hence the defect<br />

potential V (x) is a m<strong>on</strong>ot<strong>on</strong>ically decreasing functi<strong>on</strong> of the radial distance <str<strong>on</strong>g>and</str<strong>on</strong>g> both<br />

vacancies <str<strong>on</strong>g>and</str<strong>on</strong>g> vortices are repelled from the top of the bump where the curvature is<br />

positive. These defects will not remain trapped in the regi<strong>on</strong>s where the curvature<br />

is most negative, because the net force they experience results from the l<strong>on</strong>g range<br />

(logarithmic) interacti<strong>on</strong> with the Gaussian curvature everywhere <strong>on</strong> the surface. A<br />

detailed derivati<strong>on</strong> of these results al<strong>on</strong>g with heuristic arguments that make them<br />

plausible is presented in the following chapters. Here we simply show how these<br />

results fit in our general c<strong>on</strong>ceptual scheme.<br />

Although they share a comm<strong>on</strong> geometrical potential V (x), vacancies <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

x<br />

(b)


Chapter 1: Introducti<strong>on</strong> 8<br />

vortices are coupled to the Gaussian curvature by two distinct mechanisms that we<br />

broadly classify as gauge <str<strong>on</strong>g>and</str<strong>on</strong>g> anomalous interacti<strong>on</strong>s respectively. The former results<br />

from a direct coupling between the gradient of the displacement field ui(�x) <str<strong>on</strong>g>and</str<strong>on</strong>g> the<br />

geometry of the substrate as described by gradients of its height functi<strong>on</strong> h(�x). The<br />

in-plane elastic free energy for a crystal embedded in a gently curved frozen substrate<br />

can be expressed in terms of the flat space Lamé coefficients µ <str<strong>on</strong>g>and</str<strong>on</strong>g> λ [14] as<br />

�<br />

Fc = dS<br />

�<br />

µ u 2 ij + λ<br />

2 u2 �<br />

kk<br />

, (1.2)<br />

where dS is the infinitesimal area element. The strain tensor uij given by<br />

uij(�x) = 1<br />

2 [∂iuj(�x) + ∂jui(�x) + Aij(�x)] , (1.3)<br />

c<strong>on</strong>tains an additi<strong>on</strong>al term Aij(�x) = ∂ih(�x)∂jh(�x) (compared to its flat space coun-<br />

terpart) that encodes informati<strong>on</strong>s <strong>on</strong> the geometry of the substrate. This rank two<br />

tensor resembles the vector potential of electromagnetic theory; an appropriately de-<br />

fined curl is equal to the Gaussian curvature [15]. A free energy like Eq.(2) has the<br />

typical structure of the gauge theories often employed to describe c<strong>on</strong>densed matter<br />

order such as the L<strong>on</strong>d<strong>on</strong> theory of a superc<strong>on</strong>ductor well below TC. In the super-<br />

c<strong>on</strong>ductor analogy, the Gaussian curvature plays the role of a frozen <str<strong>on</strong>g>and</str<strong>on</strong>g> spatially<br />

varying external magnetic field.<br />

The gauge coupling generates a force <strong>on</strong> each defect which is given by the<br />

product of the field −∇V (x) <str<strong>on</strong>g>and</str<strong>on</strong>g> the “charge” of the α th point defect Ωα. The corre-<br />

sp<strong>on</strong>ding defect ”strength” Ωα for an isotropic vacancy (interstitial) is given by the<br />

negative (positive) area change caused by removing (adding) an atom at positi<strong>on</strong> xα.


Chapter 1: Introducti<strong>on</strong> 9<br />

Figure 1.5: An impurity in a crystalline matrix. The large shaded atom causes a local<br />

dilati<strong>on</strong> of the lattice. The square lattice has been chosen for simplicity, although the<br />

ground state of a flat two dimensi<strong>on</strong>al solid is typically an hexag<strong>on</strong>al lattice. By<br />

c<strong>on</strong>trast, a vacancy would corresp<strong>on</strong>d to removing the shaded atom from the lattice<br />

leaving a local compressi<strong>on</strong> (reproduced from Ref.[16]).<br />

The case of substituti<strong>on</strong>al impurities can be treated phenomenologically by choosing<br />

Ωα according to the characteristic ”size” of the foreign atom introduced in the original<br />

lattice (see Fig. 1.5); Ωα will hence be allowed to vary c<strong>on</strong>tinuously in our treatment<br />

[16]. The coupling c<strong>on</strong>stant for this elastic interacti<strong>on</strong> is c<strong>on</strong>trolled by the Young’s<br />

modulus Y = 4µ(µ+λ)<br />

. As shown in Chapter 5, the resulting force <strong>on</strong> a vacancy (in-<br />

2µ+λ<br />

terstitial) in the outward (inward) radial directi<strong>on</strong>, � fc, has the familiar form of the<br />

electrostatic interacti<strong>on</strong> between a charge <str<strong>on</strong>g>and</str<strong>on</strong>g> an external field,<br />

�fc = − Y<br />

2 Ωα � ∇V (x) . (1.4)<br />

The tensor Aij(�x) accounts for another key feature of two dimensi<strong>on</strong>al crys-<br />

tals <strong>on</strong> a curved substrate which is intimately related to the existence of a gauge<br />

coupling: geometric frustrati<strong>on</strong>. In the presence of Gaussian curvature, the elastic


Chapter 1: Introducti<strong>on</strong> 10<br />

ground state will always be characterized by some strain u G ij(x) <str<strong>on</strong>g>and</str<strong>on</strong>g> stress σ G ij(x) which<br />

result in a n<strong>on</strong>-vanishing stretching energy, even in the absence of defects. The origin<br />

of the l<strong>on</strong>g range geometric potential V (�x) experienced by vacancies <str<strong>on</strong>g>and</str<strong>on</strong>g> interstitials<br />

lies in the fact that the local compressi<strong>on</strong> or dilati<strong>on</strong> (measured by Ωα) that they<br />

induce couples to the preexisting isostatic pressure of the stress σ G kk (xα), which is a<br />

n<strong>on</strong>-local functi<strong>on</strong> of the Gaussian curvature. In fact, elastic deformati<strong>on</strong>s created<br />

by the geometric c<strong>on</strong>straint throughout the curved 2D solid are propagated to the<br />

positi<strong>on</strong> of the point defect, xα, by force chains spanning the entire system. Vacan-<br />

cies, interstitials or impurities atoms can all be viewed as local probes of the stress<br />

field that, as a first approximati<strong>on</strong>, are unaffected the additi<strong>on</strong>al stresses induced by<br />

their own presence (see Chapter 5). The additi<strong>on</strong>al self-energies which are not taken<br />

into account by this approach are of order Y Ω 2 αG(�x) <str<strong>on</strong>g>and</str<strong>on</strong>g> can be neglected as l<strong>on</strong>g as<br />

the ”size” of these point defects is much smaller than the radii of curvature of the<br />

substrate.<br />

By c<strong>on</strong>trast, no geometric frustrati<strong>on</strong> exists for a 4 He film <strong>on</strong> a corrugated<br />

substrate because its wave functi<strong>on</strong> Ψ(x) = A e iθ(x) is defined in a different space from<br />

the <strong>on</strong>e in which the superfluid is c<strong>on</strong>fined. In the ground state of a 4 He film, the<br />

phase θ(x) is c<strong>on</strong>stant throughout the surface <str<strong>on</strong>g>and</str<strong>on</strong>g> the corresp<strong>on</strong>ding energy vanishes.<br />

The free energy, Fs, of a n<strong>on</strong>rotating superfluid film does not include a geometric<br />

gauge field. As discussed in Chapter 3, this free energy reads<br />

Fs = K<br />

2<br />

�<br />

dS g αβ ∂αθ(x) ∂βθ(x) , (1.5)<br />

where gαβ is the metric tensor of the substrate <str<strong>on</strong>g>and</str<strong>on</strong>g> the coupling c<strong>on</strong>stant K = ρs�2<br />

m4 2 is


Chapter 1: Introducti<strong>on</strong> 11<br />

expressed in terms of the mass of an 4 He molecule m4 <str<strong>on</strong>g>and</str<strong>on</strong>g> the superfluid mass den-<br />

sity per unit area ρs. When vortices are introduced into Eq.(4.2), these ”topological<br />

charges” behave somewhat like electrostatic charges when c<strong>on</strong>fined in a bounded ge-<br />

ometry. Even in the absence of externally imposed supercurrents, positi<strong>on</strong>-dependent<br />

self-energies originate from the broken translati<strong>on</strong>al symmetry implied by the presence<br />

of the boundary or the varying Gaussian curvature. In an electrostatic analogy, each<br />

charged ”particle” induces polarizati<strong>on</strong> charges <strong>on</strong> the c<strong>on</strong>ducting walls (distributed<br />

according to the shape of the boundary) with which it interacts [17]. On a curved<br />

substrate, the induced topological charge depends both <strong>on</strong> the “charge” (quantum of<br />

circulati<strong>on</strong>) of the vortex <str<strong>on</strong>g>and</str<strong>on</strong>g> <strong>on</strong> the Gaussian curvature of the surface in which it is<br />

imbedded. The resulting force, fs is given in terms of the soluti<strong>on</strong> of Eq.(1.1) by<br />

�fs = πK s 2 � ∇V (x) , (1.6)<br />

where the integer s measures the vorticity of the defect. Unlike the charge Ωα that<br />

describes the local area deformati<strong>on</strong>s in crystals, this topological charge of a vortex<br />

or anti-vortex must be quantized. A comparis<strong>on</strong> of Eq.(1.4) <str<strong>on</strong>g>and</str<strong>on</strong>g> Eq.(1.6) reveals that<br />

the geometric forces experienced by vacancies (for which Ωα < 0) <str<strong>on</strong>g>and</str<strong>on</strong>g> vortices or<br />

anti-vortices in 4 He are indeed the same (apart from multiplicative c<strong>on</strong>stants). Note<br />

however that the dependence <strong>on</strong> the ”charge” of the point defect, Ωα, is linear in<br />

Eq.(1.4) whereas the force in Eq.(1.6) depends quadratically <strong>on</strong> the vortex charge<br />

s. The mechanism of coupling between curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> vortices in liquid helium is<br />

qualitatively different from the gauge coupling discussed in the c<strong>on</strong>text of curved<br />

space crystallography <str<strong>on</strong>g>and</str<strong>on</strong>g> in what follows we will refer to it as anomalous coupling.


Chapter 1: Introducti<strong>on</strong> 12<br />

Figure 1.6: M<strong>on</strong>te Carlo simulati<strong>on</strong> of curvature induced unbinding of dislocati<strong>on</strong>s for<br />

particles interacting with a Yukawa pair potential <strong>on</strong> two similar curved substrates<br />

[18]. The dark <str<strong>on</strong>g>and</str<strong>on</strong>g> light particles in the dislocati<strong>on</strong> core represent a disclinati<strong>on</strong> pair.<br />

The two panels represent c<str<strong>on</strong>g>and</str<strong>on</strong>g>idate ground states that have lower energies than the<br />

respective c<strong>on</strong>figurati<strong>on</strong>s without defects. In the left panel, unbound dislocati<strong>on</strong>s (i.e.<br />

disclinati<strong>on</strong>s pairs) are oppositely aligned in the radial directi<strong>on</strong> towards the bump.<br />

In the right panel, dislocati<strong>on</strong> dipoles neutralize their Burger’s vectors by having<br />

their (opposite) dipole orientati<strong>on</strong> alternate between being pointed towards bumps<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> away from saddles. The first scenario is favored for large bump separati<strong>on</strong>s.<br />

Other defects such as disclinati<strong>on</strong>s in liquid crystals fit in this classificati<strong>on</strong> with both<br />

types of couplings playing an important role.<br />

An interesting c<strong>on</strong>sequence of geometric frustrati<strong>on</strong> in curved crystals <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

liquid crystals is the possibility of structural transiti<strong>on</strong>s between a defect-free ground<br />

state to energetically favored c<strong>on</strong>figurati<strong>on</strong>s where defects nucleate to partially screen<br />

the Gaussian curvature. Fig. 1.6 shows the result of minimizing the energy of point<br />

particles interacting via a Yukawa potential <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>strained to lie <strong>on</strong> a corrugated<br />

geometry with two different values of the bump spacing relative to the bump size<br />

[18]. The two c<str<strong>on</strong>g>and</str<strong>on</strong>g>idate ground states characterized by a ”charge-neutral” set of<br />

dislocati<strong>on</strong>s have lower energies than a frustrated but defect-free hexag<strong>on</strong>al lattice.<br />

Both the actual distributi<strong>on</strong> of dislocati<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> the critical aspect ratio of the bumps


Chapter 1: Introducti<strong>on</strong> 13<br />

necessary to trigger the instability can be calculated from the geometric potential of<br />

an isolated dislocati<strong>on</strong>, which is a functi<strong>on</strong> of both its positi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> the orientati<strong>on</strong> of<br />

its Burger vector (see Chapter 5).<br />

1.1.3 Cross-over between 2D <str<strong>on</strong>g>and</str<strong>on</strong>g> 3D physics<br />

The theoretical framework described in the previous pages applies to very<br />

thin films of approximately c<strong>on</strong>stant thickness which can be treated as two dimen-<br />

si<strong>on</strong>al systems. It is interesting to study the crossover from the two to three di-<br />

mensi<strong>on</strong>al regime that occurs as the thickness of the layer of material, h, increases.<br />

C<strong>on</strong>sider for c<strong>on</strong>creteness the case of a nematic shell coating a solid colloidal parti-<br />

cle <strong>on</strong> which the nematic molecules are aligned tangentially or a ”double emulsi<strong>on</strong>”<br />

composed of a nematic liquid coating PVA enriched water droplets in a soluti<strong>on</strong> of<br />

glycerol, PVA <str<strong>on</strong>g>and</str<strong>on</strong>g> water. Experimental investigati<strong>on</strong>s of this later system are cur-<br />

rently underway in the Weitz lab [11].<br />

In the m<strong>on</strong>o-layer limit, the baseball like texture reproduced in Fig. 1.2 is<br />

characterized by four s = 1<br />

2<br />

disclinati<strong>on</strong>s. However, for thicker shells three dimen-<br />

si<strong>on</strong>al defect c<strong>on</strong>figurati<strong>on</strong>s characterized by two pairs of half-hedgehogs at the inner<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> outer surface compete with the planar texture discussed previously leading to<br />

structural transiti<strong>on</strong>s above a critical value of hc. The escaped 3D texture is illus-<br />

trated schematically in the right most panel of Fig. 1.7 where the pair of surface<br />

half-hedgehogs at the south pole is highlighted by the small circle <str<strong>on</strong>g>and</str<strong>on</strong>g> the center of<br />

each defect indicated by a dot. For samples between crossed polarizers, the optical<br />

signature of this transiti<strong>on</strong> is a change from the quadrupolar to the bipolar defect


Chapter 1: Introducti<strong>on</strong> 14<br />

Figure 1.7: Quadrupolar <str<strong>on</strong>g>and</str<strong>on</strong>g> bipolar double-emulsi<strong>on</strong>s observed through crossed polarizers.<br />

In the left panel, the 4 disclinati<strong>on</strong>s are visible as 4 two-fold brushes as in<br />

Fig. 1.2. Each of the two pairs of half-hegehogs is visible in the middle panel as<br />

a 4-fold brush. Experimental c<strong>on</strong>diti<strong>on</strong>s are similar to Fig. 1.2. The right panel<br />

shows a schematic illustrati<strong>on</strong> of a c<str<strong>on</strong>g>and</str<strong>on</strong>g>idate nematic texture escaped in the third<br />

dimensi<strong>on</strong> that is c<strong>on</strong>sistent with the image of the bipolar droplet presented in the<br />

middle panel. Thickness inhomogeneities in the nematic shell are believed to bring<br />

these defect patterns into the same hemisphere, which allows easy visualizati<strong>on</strong>.<br />

patterns illustrated in Fig. 1.7.<br />

1.1.4 Outline of the thesis<br />

This thesis is organized around the themes sketched in the previous sec-<br />

ti<strong>on</strong>s. In Chapter 2 we present a detailed study of liquid crystal order <strong>on</strong> surfaces<br />

of varying Gaussian curvature that relies <strong>on</strong> the geometric potential of an individual<br />

disclinati<strong>on</strong>. The (exact) n<strong>on</strong>-perturbative soluti<strong>on</strong> of the Poiss<strong>on</strong> equati<strong>on</strong> by means<br />

of c<strong>on</strong>formal mappings introduced in this c<strong>on</strong>text is an important mathematical in-<br />

gredient of our approach. In Chapter 3, the crucial distincti<strong>on</strong> between the gauge <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

anomalous couplings to the Gaussian curvature is discussed by comparing the physics<br />

of liquid crystal <str<strong>on</strong>g>and</str<strong>on</strong>g> superfluid films <strong>on</strong> a corrugated substrate. The mathematical<br />

descripti<strong>on</strong> of these two systems is very similar in the plane [1, 2], but this chapter<br />

reveals remarkable differences <strong>on</strong> a curved substrate. In Chapter 4 the ground state


Chapter 1: Introducti<strong>on</strong> 15<br />

of liquid crystal shells is explored including an analysis of the crossover between two<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> three dimensi<strong>on</strong>al regimes. In Chapter 5, we c<strong>on</strong>clude with a study of the curved<br />

space crystallography of dislocati<strong>on</strong>s, vacancies <str<strong>on</strong>g>and</str<strong>on</strong>g> interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> impurity atoms.<br />

Our approach starts from the derivati<strong>on</strong>s of the geometric potentials which act <strong>on</strong><br />

these defects <str<strong>on</strong>g>and</str<strong>on</strong>g> proceeds with a discussi<strong>on</strong> of stress relaxati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> defect unbind-<br />

ing instabilities in two dimensi<strong>on</strong>al crystals <strong>on</strong> curved substrates. The dynamics of<br />

dislocati<strong>on</strong> moti<strong>on</strong> <strong>on</strong> a curved surface is discussed as well.


Chapter 2<br />

B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a<br />

corrugated substrate.<br />

Based <strong>on</strong> V. Vitelli <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong> PRE 70, 051105 (2004).<br />

16


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 17<br />

2.1 Introducti<strong>on</strong><br />

The melting of a two dimensi<strong>on</strong>al crystal can occur c<strong>on</strong>tinuously via two<br />

sec<strong>on</strong>d order topological phase transiti<strong>on</strong>s characterized by the successive unbinding<br />

of dislocati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> disclinati<strong>on</strong> pairs. At low temperatures, dislocati<strong>on</strong>s are suppressed<br />

due to their large energy cost, but as the temperature is increased, the entropy gained<br />

by creating defects overcomes their cost in elastic energy <str<strong>on</strong>g>and</str<strong>on</strong>g> dislocati<strong>on</strong> unbinding<br />

occurs to reduces the overall free energy of the system [19, 20, 21]. The quasi-l<strong>on</strong>g<br />

range order of the crystal is thus destroyed leading to an hexatic phase that still<br />

preserves quasi-l<strong>on</strong>g range orientati<strong>on</strong>al order. This phase can be characterized by<br />

a complex order parameter with six-fold symmetry. As the temperature is increased<br />

still further, an additi<strong>on</strong>al disclinati<strong>on</strong>-unbinding transiti<strong>on</strong> occurs <str<strong>on</strong>g>and</str<strong>on</strong>g> the hexatic<br />

order is finally lost in an isotropic liquid phase [19].<br />

Experimental evidence for hexatic order <str<strong>on</strong>g>and</str<strong>on</strong>g> defect-mediated melting has<br />

been obtained in systems as diverse as free st<str<strong>on</strong>g>and</str<strong>on</strong>g>ing liquid crystal films [22], Langmuir-<br />

Blodgett surfactant m<strong>on</strong>olayers [23], two-dimensi<strong>on</strong>al magnetic bubble arrays [24],<br />

electr<strong>on</strong>s trapped <strong>on</strong> the surface of liquid helium [25, 26, 27], two-dimensi<strong>on</strong>al colloidal<br />

crystals [28, 29] <str<strong>on</strong>g>and</str<strong>on</strong>g> self-assembled block copolymers [30].<br />

The unbinding of defects in the plane is entropically driven <str<strong>on</strong>g>and</str<strong>on</strong>g> at low tem-<br />

perature defects are tightly bound. By c<strong>on</strong>trast, <strong>on</strong> surfaces with n<strong>on</strong> zero (integrated)<br />

Gaussian curvature, excess defects must be present even at very low temperatures.<br />

The theory of topological defects in ordered phases c<strong>on</strong>fined to frozen topographies<br />

with positive or negative Gaussian curvature has been investigated previously; see,


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 18<br />

e.g., [15, 31, 2]. As a general rule, regi<strong>on</strong>s of positive or negative curvature (valleys,<br />

hills or saddles) lead to unpaired disclinati<strong>on</strong>s in the ground state, possibly screened<br />

by clouds of dislocati<strong>on</strong>s. These clouds can in turn c<strong>on</strong>dense into grain boundaries<br />

at low temperature. The predicti<strong>on</strong>s of recent studies of crystalline order <strong>on</strong> a sphere<br />

[4] have been c<strong>on</strong>firmed in elegant studies of colloidal particles packed <strong>on</strong> the surface<br />

of water droplets in oil [3]. Investigati<strong>on</strong>s of the physics of defects in curved spaces<br />

have also been carried out for fluctuating geometries [32, 33, 34, 35]. The dynamics of<br />

hexatic order <strong>on</strong> fluctuating spherical interfaces was studied in Ref. [36]. Quenched<br />

r<str<strong>on</strong>g>and</str<strong>on</strong>g>om topographies in the limit of small deviati<strong>on</strong>s from flatness were investigated<br />

in Ref. [15] .<br />

In the present work, we investigate topography-driven generati<strong>on</strong> of defects<br />

<strong>on</strong> simple frozen surfaces with spatially varying Gaussian curvature whose topology<br />

does not automatically enforce their presence in the ground state. We study in<br />

particular a two-dimensi<strong>on</strong>al ”bump” with a Gaussian shape <str<strong>on</strong>g>and</str<strong>on</strong>g> dimensi<strong>on</strong> large<br />

compared to the particle spacing. For such a hilly l<str<strong>on</strong>g>and</str<strong>on</strong>g>scape, flat at infinity, the<br />

geometric c<strong>on</strong>trol parameter is an aspect ratio given by the bump height divided by<br />

its spatial extent. C<strong>on</strong>sider a hexatic phase draped over such a bump. For small<br />

bumps, the ideal hexatic texture is distorted, but there are no defects in the ground<br />

state. As the aspect ratio is increased, we find that disclinati<strong>on</strong> pairs progressively<br />

unbind at T = 0 in a sequence of transiti<strong>on</strong>s occurring at critical values of the aspect<br />

ratio. The defects subsequently positi<strong>on</strong> themselves to partially screen the Gaussian<br />

curvature. For bumps embedded in surfaces of sufficiently small spatial extent, a


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 19<br />

sec<strong>on</strong>d instability of the smooth ground state needs to be c<strong>on</strong>sidered. In this case,<br />

the energy stored in the field can be lowered by generating a single positive defect at<br />

the center of the bump. Novel effects also arise when the hilly surface is encircled by<br />

an aligning circular wall that insures a 2π rotati<strong>on</strong> of the orientati<strong>on</strong>al order in the<br />

ground state. In this case, some of the positive defects required to match the curvature<br />

of the boundary are c<strong>on</strong>fined to a hemispherical cup centered <strong>on</strong> the bump, provided<br />

the aspect ratio α is larger than a critical value αD. When α is lowered below αD,<br />

the positive defects originally ”trapped” in the hemispherical cup start undergoing<br />

a series of sharp ”dec<strong>on</strong>finement transiti<strong>on</strong>s”, as they progressively migrate to new<br />

equilibrium positi<strong>on</strong>s dictated by boundary c<strong>on</strong>diti<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> the finite system size. We<br />

also suggest possible ground states for periodic arrays of bumps, like those <strong>on</strong> the<br />

bottom of an egg cart<strong>on</strong>.<br />

A natural arena to experimentally study the interplay between geometry<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> defects is provided by thin copolymer films <strong>on</strong> SiO2 patterned substrates [37].<br />

Flat space experiments by Segalman et al. have already dem<strong>on</strong>strated that spherical<br />

domains in block copolymer films form hexatic phases [30].<br />

Our results for hexatics <strong>on</strong> frozen topographies also apply to other XY-<br />

like models, as might be appropriate for tilted surface-active molecules <strong>on</strong> curved<br />

substrates with interacti<strong>on</strong>s which favor alignment. The results are relevant as well to<br />

two-fold nematic order <strong>on</strong> frozen topographies. In both cases, we expect qualitatively<br />

similar defect unbinding transiti<strong>on</strong>s, although the equivalence becomes more exact<br />

in the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong> [38]. Related results have been obtained


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 20<br />

recently for order <strong>on</strong> a torus [39]. Even though the integrated Gaussian curvature<br />

vanishes, defects appear in the ground state in the limit of fat torii, unless the number<br />

of degrees of freedom is very large.<br />

The outline of this paper is as follows. In Secti<strong>on</strong> II, the relevant mathemat-<br />

ical formalism is introduced <str<strong>on</strong>g>and</str<strong>on</strong>g> used to highlight similarities <str<strong>on</strong>g>and</str<strong>on</strong>g> differences between<br />

defects <strong>on</strong> surfaces of varying curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> electrostatic charges in a n<strong>on</strong>-uniform<br />

background charge distributi<strong>on</strong> in flat space. As an example, we calculate the dis-<br />

torted, but defect-free, ground state texture of a hexatic c<strong>on</strong>fined to a surface shaped<br />

as a ”Gaussian bump” for aspect ratios below the first disclinati<strong>on</strong>-unbinding instabil-<br />

ity. In Secti<strong>on</strong> III, we investigate curvature-induced defect formati<strong>on</strong> for an isolated<br />

bump <str<strong>on</strong>g>and</str<strong>on</strong>g> a periodic array of bumps. In secti<strong>on</strong> IV, defect dec<strong>on</strong>finement is discussed<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> in secti<strong>on</strong> V various experimental issues related to our analysis are highlighted<br />

al<strong>on</strong>g with some directi<strong>on</strong>s for future work. The development of the mathematical<br />

formalism is largely relegated to Appendices. In Appendix A the Green’s functi<strong>on</strong><br />

for the covariant Laplacian is derived by means of c<strong>on</strong>formal transformati<strong>on</strong>s. In Ap-<br />

pendix B, we introduce a geometric potential whose source is the Gaussian curvature.<br />

In Appendix C, we present the general formula for the energy of textures with defects<br />

in terms of the two functi<strong>on</strong>s derived in Appendix A <str<strong>on</strong>g>and</str<strong>on</strong>g> B. We thus explore the<br />

existence of positi<strong>on</strong>-dependent defect self-interacti<strong>on</strong>s that arise from the varying<br />

Gaussian curvature. Finally, boundary effects are discussed in Appendix D.


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 21<br />

2.2 Hexatic order <strong>on</strong> a surface<br />

2.2.1 Electrostatic analogy<br />

The free energy for hexatic degrees of freedom embedded in an arbitrary<br />

frozen surface can be written as<br />

F = KA<br />

2<br />

�<br />

dA Dαn β (u)D α nβ(u) , (2.1)<br />

where u = {u1, u2} is a set of internal coordinates, n(u) is a unit vector in the tangent<br />

plane, Dα is the covariant derivative with respect to the metric of the surface <str<strong>on</strong>g>and</str<strong>on</strong>g> dA is<br />

the infinitesimal surface area [33, 32, 35, 40]. The generalizati<strong>on</strong> to systems with a p-<br />

fold symmetry is straightforward provided that the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong><br />

is used <str<strong>on</strong>g>and</str<strong>on</strong>g> the c<strong>on</strong>sequences of the uniaxial coupling neglected [38]. This choice of<br />

free energy implies that the minimal energy c<strong>on</strong>figurati<strong>on</strong> will be given locally by<br />

neighboring n(u) vectors which differ <strong>on</strong>ly by parallel transport. The curvature of the<br />

surface induces ”frustrati<strong>on</strong>” in the texture. In fact, by Gauss’ ”Theorema egregium”<br />

[41, 42], tangent vectors parallel transported al<strong>on</strong>g a closed loop are rotated by an<br />

amount equal to the Gaussian curvature integrated over the enclosed area. On a<br />

sphere, for example, the hexatic ground state always has twelve excess disclinati<strong>on</strong>s<br />

as a result of this frustrati<strong>on</strong> [31, 12]. More generally, the sum of the topological<br />

charges <strong>on</strong> any closed surface is equal to the integrated Gaussian curvature.<br />

By introducing a local b<strong>on</strong>d-angle field θ(u), corresp<strong>on</strong>ding to the angle<br />

between n(u) <str<strong>on</strong>g>and</str<strong>on</strong>g> an arbitrary local reference frame, we can rewrite the hexatic free


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 22<br />

energy introduced in Eq.(3.1) as:<br />

F = 1<br />

2 KA<br />

�<br />

dA g αβ (∂αθ − Aα)(∂βθ − Aβ) , (2.2)<br />

where dA = d 2 u √ g, g is the determinant of the metric tensor gαβ <str<strong>on</strong>g>and</str<strong>on</strong>g> Aβ is the spin-<br />

c<strong>on</strong>necti<strong>on</strong> whose curl is the Gaussian curvature G(u) [40, 42]. The spin c<strong>on</strong>necti<strong>on</strong><br />

can be viewed as a ”geometric vector potential”. A free energy like Eq.(2) also<br />

describes the charged Cooper pairs implicit in the L<strong>on</strong>d<strong>on</strong> theory of a superc<strong>on</strong>ductor<br />

well below TC. In the superc<strong>on</strong>ductor analogy, the Gaussian curvature plays the role<br />

of a (spatially varying) external magnetic field. For the problem c<strong>on</strong>sidered here,<br />

however, there are interesting new n<strong>on</strong>linear effects associated with spatial variati<strong>on</strong>s<br />

in the metric.<br />

A detailed analysis of the free energy of Eq.(2.2) for a bumpy surface with<br />

free <str<strong>on</strong>g>and</str<strong>on</strong>g> circular boundary c<strong>on</strong>diti<strong>on</strong>s is presented in Appendices (C) <str<strong>on</strong>g>and</str<strong>on</strong>g> (D). Here<br />

we <strong>on</strong>ly sketch the main steps <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>clusi<strong>on</strong>s. The free energy can be readily c<strong>on</strong>-<br />

verted into a Coulomb gas model by using the relati<strong>on</strong><br />

γ αβ ∂α(∂βθ − Aβ) = s(u) − G(u) ≡ n(u) , (2.3)<br />

where γ αβ is the covariant antisymmetric tensor, G(u) is the Gaussian curvature <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

s(u) ≡ 1 �Nd √<br />

g i=1 qiδ(u − ui) is the disclinati<strong>on</strong> density with Nd defects of charge qi at<br />

positi<strong>on</strong>s ui. The final result is an effective free energy whose basic degrees of freedom<br />

are the defects themselves [35, 4]:<br />

F = KA<br />

2<br />

�<br />

�<br />

dA<br />

dA ′ n(u) Γ(u, u ′ ) n(u ′ ) , (2.4)


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 23<br />

where n(u) is defined in Eq.(E.18). The Green’s functi<strong>on</strong> Γ(u, u ′ ) is calculated (see<br />

Appendix A) by inverting the Laplacian defined <strong>on</strong> the surface<br />

Γ(u, u ′ � �<br />

1<br />

) ≡ −<br />

∆ uu ′<br />

, (2.5)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> we have suppressed for now defect core energy c<strong>on</strong>tributi<strong>on</strong>s which reflect the<br />

physics at microscopic length scales. Eq.(E.19) can be understood by analogy to two<br />

dimensi<strong>on</strong>al electrostatics, with the Gaussian curvature G(u) (with sign reversed)<br />

playing the role of a n<strong>on</strong>-uniform background charge distributi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> the topolog-<br />

ical defects appearing as point-like sources with electrostatic charges equal to their<br />

topological charge qi. As a result, the defects tend to positi<strong>on</strong> themselves so that<br />

the Gaussian curvature is screened: the positive <strong>on</strong>es <strong>on</strong> peaks <str<strong>on</strong>g>and</str<strong>on</strong>g> valleys <str<strong>on</strong>g>and</str<strong>on</strong>g> the<br />

negative <strong>on</strong>es <strong>on</strong> the saddles of the surface. However, this analogy does neglect<br />

positi<strong>on</strong>-dependent self-interacti<strong>on</strong>s [43], but since these are quadratic in the charge<br />

they are negligible for hexatics. Hence positive disclinati<strong>on</strong>s of minimal topological<br />

charge qi = 2π<br />

6<br />

c<strong>on</strong>tinue to be attracted to positive curvature (see Appendix C).<br />

More generally, we can c<strong>on</strong>sider p-fold symmetric order parameters with<br />

minimum charge defects ± 2π.<br />

The case p = 1 corresp<strong>on</strong>ds to tilt order of absorbed<br />

p<br />

molecules <str<strong>on</strong>g>and</str<strong>on</strong>g> p = 2 describes 2D nematics. The cases p = 4 <str<strong>on</strong>g>and</str<strong>on</strong>g> p = 6 describe<br />

tetradic <str<strong>on</strong>g>and</str<strong>on</strong>g> hexatic phases respectively [12]. Strictly speaking, Eq.(1) <strong>on</strong>ly describes<br />

the cases p = 1 <str<strong>on</strong>g>and</str<strong>on</strong>g> p = 2 in the <strong>on</strong>e-Frank-c<strong>on</strong>stant approximati<strong>on</strong> [38]. Most of<br />

our discussi<strong>on</strong> focuses <strong>on</strong> topography-driven transiti<strong>on</strong>s <strong>on</strong> a model surface shaped<br />

like a bell curve or ”Gaussian bump” (see Fig. 2.1), but the same mathematical<br />

approach can be readily carried over to study arbitrary surfaces of revoluti<strong>on</strong> that


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 24<br />

Figure 2.1: (a) The vector field n is c<strong>on</strong>fined to a surface shaped as a Gaussian.<br />

(b) Top view of (a) showing a schematic representati<strong>on</strong> of the positive <str<strong>on</strong>g>and</str<strong>on</strong>g> negative<br />

Gaussian curvature as a background ”charge” distributi<strong>on</strong> that switches sign at r =<br />

r0. Note that, according to the electrostatic analogy, a positive (negative) distributi<strong>on</strong><br />

of Gaussian curvature corresp<strong>on</strong>ds to negative (positive) topological charge density.<br />

are topologically equivalent to the plane. Furthermore, we do not expect the results<br />

of this analysis to depend qualitatively <strong>on</strong> the azimuthal symmetry of the surface,<br />

which is assumed purely for reas<strong>on</strong>s of mathematical c<strong>on</strong>venience.<br />

Points <strong>on</strong> our model surface embedded in three dimensi<strong>on</strong>al Euclidean space<br />

are specified by a three dimensi<strong>on</strong>al vector R(r, φ) given by<br />

⎛<br />

⎞<br />

⎜ r cos φ<br />

⎜<br />

R(r, φ) = ⎜<br />

r sin φ<br />

⎝<br />

h exp<br />

�<br />

− r2<br />

2r 2 0<br />

⎟ , (2.6)<br />

⎟<br />

� ⎠<br />

where r <str<strong>on</strong>g>and</str<strong>on</strong>g> φ are plane polar coordinates in the xy plane of Fig. 2.1. It is useful<br />

to characterize the deviati<strong>on</strong> of the bump from a plane in terms of a dimensi<strong>on</strong>less


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 25<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

l(r/r0)<br />

r/r0<br />

0.5 1 1.5 2 2.5 3<br />

Figure 2.2: Plot of l( r<br />

r<br />

) as a functi<strong>on</strong> of the dimensi<strong>on</strong>less radial coordinate r0 r0 for<br />

α = 1, 2, 3, 4. The arrow is oriented in the directi<strong>on</strong> of increasing α.<br />

aspect ratio<br />

α ≡ h<br />

r0<br />

. (2.7)<br />

The two orthog<strong>on</strong>al tangent vectors tr ≡ ∂R<br />

∂r <str<strong>on</strong>g>and</str<strong>on</strong>g> tφ ≡ ∂R can be normalized to define<br />

∂φ<br />

the Vierbein (orth<strong>on</strong>ormal basis vectors) Er <str<strong>on</strong>g>and</str<strong>on</strong>g> Eφ respectively. The comp<strong>on</strong>ents of<br />

the spin c<strong>on</strong>necti<strong>on</strong> introduced in Eq.(2.2) are given by Aα = Er · ∂αEφ [40, 42]. This<br />

leads to a vanishing radial comp<strong>on</strong>ent Ar <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

Aφ = − 1<br />

� l(r) , (2.8)<br />

where the important α-dependent functi<strong>on</strong> l(r) (see Fig. 2.2) is defined by<br />

l(r) ≡ 1 + α2 r 2<br />

r 2 0<br />

�<br />

exp − r2<br />

r2 �<br />

0<br />

, (2.9)


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 26<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> it is equal to the radial comp<strong>on</strong>ent of the diag<strong>on</strong>al metric tensor, gαβ,<br />

⎛<br />

⎜l(r)<br />

gαβ = ⎝<br />

0<br />

0 r2 ⎞<br />

⎟<br />

⎠ . (2.10)<br />

Note that the gφφ entry is equal to the flat space result r 2 in polar coordinates while<br />

grr = l(r) is modified in a way that depends <strong>on</strong> α but tends to the plane result grr = 1<br />

for small <str<strong>on</strong>g>and</str<strong>on</strong>g> large r, as illustrated in Fig. 2.2.<br />

The Gaussian curvature for the bump is readily found from the eigenvalues<br />

of the sec<strong>on</strong>d fundamental form [44],<br />

Gα(r) = α2 r2<br />

−<br />

r e 2 0<br />

r2 0 l(r) 2<br />

�<br />

1 − r2<br />

r2 �<br />

0<br />

. (2.11)<br />

Note that α c<strong>on</strong>trols the order of magnitude of G(r) <str<strong>on</strong>g>and</str<strong>on</strong>g> that G(r) changes sign at<br />

r = r0 (see Fig. 2.1b). The integrated Gaussian curvature ∆G(r) inside a cup of<br />

radius r centered <strong>on</strong> the bump is<br />

∆G(r) = 2π<br />

�<br />

1 − 1<br />

�<br />

�<br />

l(r)<br />

, (2.12)<br />

which vanishes as r → ∞. Eq.(2.12) also shows that the positive Gaussian curvature<br />

enclosed within the radius r0 (see Fig. 2.1) approaches 2π for α ≫ 1, half the<br />

integrated Gaussian curvature of a sphere.<br />

2.2.2 Defect free texture<br />

For small values of the aspect ratio α, the minimal energy texture for the<br />

hexatic will be free of defects. The ground state c<strong>on</strong>figurati<strong>on</strong> θo(u) satisfies the<br />

differential equati<strong>on</strong><br />

DαD α θ0 − D α Aα = 0 , (2.13)


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 27<br />

Figure 2.3: Projected ground state texture for an XY model <strong>on</strong> the bump, with the<br />

boundary c<strong>on</strong>diti<strong>on</strong> that the vector field is parallel to the y-axis at infinity. The two<br />

insets show the defect pairs suggested by two regi<strong>on</strong>s of large frustrati<strong>on</strong>, which lie<br />

close to a circle of radius r0.<br />

which results from minimizing the free energy in Eq.(2.2) with respect to the field θ(u)<br />

for fixed Aα. When expressed in terms of the coordinates in Eq.(2.6), the soluti<strong>on</strong> of<br />

Eq.(2.13) reads:<br />

θo(u) = −φ + c , (2.14)<br />

where c is an arbitrary c<strong>on</strong>stant. The smooth ground state texture is thus obtained<br />

if the director n forms an angle θo(u) = −φ + c with respect to the spatially varying<br />

basis vector Er. Note that a soluti<strong>on</strong> of the form θo(u) = c represents a defect of<br />

charge q = 2π in this ”rotating” system of coordinates.<br />

As an illustrati<strong>on</strong>, c<strong>on</strong>sider the projecti<strong>on</strong> <strong>on</strong> the plane of the minimal en-<br />

ergy texture of an XY model (p = 1) as shown in Fig. 2.3. The arrows represent<br />

the orientati<strong>on</strong> of tilted molecules <strong>on</strong> this surface in the <strong>on</strong>e Frank c<strong>on</strong>stant approxi-<br />

mati<strong>on</strong>. The field clearly displays str<strong>on</strong>g frustrati<strong>on</strong> al<strong>on</strong>g a directi<strong>on</strong> determined by


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 28<br />

the choice of the c<strong>on</strong>stant c in Eq.(2.14). If the bump is positi<strong>on</strong>ed within two very<br />

distant walls parallel to the y-axis which impose tangential boundary c<strong>on</strong>diti<strong>on</strong>s <strong>on</strong><br />

the molecular tilts, the ”preferred” directi<strong>on</strong> will be al<strong>on</strong>g ˆy 1 . The texture displayed<br />

in Fig. 2.3 can be interpreted as resulting from embry<strong>on</strong>ic pairs of defect dipoles<br />

al<strong>on</strong>g the line x = 0. The distorti<strong>on</strong> energy F0 of this ground state is given by:<br />

F0 = 1<br />

2 KA<br />

�<br />

dA g αβ (∂αθo − Aα)(∂βθo − Aβ) . (2.15)<br />

This expressi<strong>on</strong> can be evaluated for an infinitely large system by using Eq.(2.14) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

the explicit form of the spin c<strong>on</strong>necti<strong>on</strong> derived in Secti<strong>on</strong> 2.2.1, with the result:<br />

� ∞<br />

�<br />

1 −<br />

F0 = πKA dr<br />

� �2 l(r)<br />

r � l(r)<br />

. (2.16)<br />

0<br />

It follows from Eq.(2.16) that the ground state energy is a m<strong>on</strong>ot<strong>on</strong>ically increasing<br />

functi<strong>on</strong> of the aspect ratio, proporti<strong>on</strong>al to α 4 for small α. As we shall see, for large<br />

enough α, it can be energetically preferable to reduce this energy by introducing<br />

defect pairs into the texture. It is c<strong>on</strong>venient to rewrite Eq.(2.15) in terms of the<br />

Gaussian curvature G(r) <str<strong>on</strong>g>and</str<strong>on</strong>g> the Green functi<strong>on</strong> Γ(u, u ′ ) discussed in Appendix A<br />

[40]<br />

F0 = KA<br />

2<br />

�<br />

�<br />

dA<br />

dA ′ G(u) Γ(u, u ′ ) G(u ′ ) . (2.17)<br />

This result is what <strong>on</strong>e obtains by setting all qi = 0 in Eq.(E.19). The details of the<br />

mathematical derivati<strong>on</strong> are relegated to Appendix C.<br />

Although this result correctly represents the zero temperature limit of the<br />

vector model, correcti<strong>on</strong>s may be appropriate to describe the physics of ordered phases<br />

1 In case distant walls are present, the free soluti<strong>on</strong> of Eq.(2.14) is slightly modified to account for the<br />

new boundary c<strong>on</strong>diti<strong>on</strong>s. This is accomplished by the method of images or c<strong>on</strong>formal transformati<strong>on</strong>s [45].


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 29<br />

at finite temperature. ”Spin-wave” excitati<strong>on</strong>s (i.e., quadratic fluctuati<strong>on</strong>s of the<br />

order parameter about the ground state texture) can be accounted for by integrating<br />

out the l<strong>on</strong>gitudinal fluctuati<strong>on</strong>s θ ′ (u) around the ground state c<strong>on</strong>figurati<strong>on</strong> θo(u).<br />

By letting θ = θ0 + θ ′ in Eq.(2.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> using Eq.(E.18) we obtain [32, 35]:<br />

F = F0 + 1<br />

2 KA<br />

�<br />

dA g αβ ∂αθ ′ ∂βθ ′ , (2.18)<br />

The l<strong>on</strong>gitudinal variable θ ′ (u) appears <strong>on</strong>ly quadratically in F <str<strong>on</strong>g>and</str<strong>on</strong>g> the trace over<br />

θ ′ (u) can be explicitly performed with the result [46]:<br />

�<br />

Dθ ′ e − βKA 2<br />

where FL is the Liouville acti<strong>on</strong>,<br />

�<br />

βFL = c<br />

dA − KA<br />

24<br />

�<br />

R dA g αβ ∂αθ ′ ∂βθ ′<br />

�<br />

dA<br />

= e −βFL , (2.19)<br />

dA ′ G(u) Γ(u, u ′ ) G(u ′ ) . (2.20)<br />

The first term in this expressi<strong>on</strong> is a c<strong>on</strong>stant proporti<strong>on</strong>al to the fixed surface area<br />

of the frozen topography <str<strong>on</strong>g>and</str<strong>on</strong>g> will be suppressed in what follows. The remaining term<br />

causes a shift in the coupling c<strong>on</strong>stant appearing in Eq.(2.17) from KA to K ′ A =<br />

KA − kBT<br />

12π [32, 35]. This ”entropic” correcti<strong>on</strong> to the coupling c<strong>on</strong>stant KA at finite<br />

temperature also arises when defects are present.<br />

The energy in Eq.(2.17) represents an intrinsic, irreducible energy cost of<br />

geometric frustrati<strong>on</strong> for textures without defects. As we shall see, defects can reduce<br />

this frustrati<strong>on</strong>. However, for small values of α the energy cost of this frustrati<strong>on</strong><br />

will still be lower than the core energies associated with the creati<strong>on</strong> of the unbound<br />

defects <str<strong>on</strong>g>and</str<strong>on</strong>g> the work necessary to tear them apart.


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 30<br />

2.2.3 Energetics of defect pairs <strong>on</strong> a Gaussian bump<br />

A quantitative underst<str<strong>on</strong>g>and</str<strong>on</strong>g>ing of the energetics of defects <strong>on</strong> a fixed topog-<br />

raphy is essential to calculate the critical value(s) of the aspect ratio above which<br />

defect-unbinding becomes energetically favorable. The first step is to calculate the<br />

Green’s functi<strong>on</strong> Γ(u, u ′ ) that governs the ”coulombic” interacti<strong>on</strong> am<strong>on</strong>g defects <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

between each defect <str<strong>on</strong>g>and</str<strong>on</strong>g> the Gaussian curvature. The inversi<strong>on</strong> of the curved space<br />

Laplacian can be more easily accomplished by employing a set of ”isothermal coor-<br />

dinates”, such that the resulting Green’s functi<strong>on</strong> reduces to the familiar logarithm<br />

of two dimensi<strong>on</strong>al electrostatics. As shown in Appendix A the final result in terms<br />

of the original polar coordinates reads:<br />

Γ(u, u ′ ) = − 1<br />

4π ln[ℜ(r)2 + ℜ(r ′ ) 2<br />

− 2ℜ(r)ℜ(r ′ ) cos(φ − φ ′ )] + c , (2.21)<br />

where the functi<strong>on</strong> ℜ(r) can be thought of as a radial coordinate in the c<strong>on</strong>formal<br />

plane resulting from adopting an isothermal set of coordinates (see Appendix A)<br />

ℜ(r) = r e − R ∞<br />

r<br />

dr ′<br />

r ′<br />

“√ ”<br />

l(r ′ )−1<br />

, (2.22)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> l(r) is the α-dependent functi<strong>on</strong> introduced in Eq.(2.9). The c<strong>on</strong>stant c depends<br />

<strong>on</strong> the physics at short distances, which is discussed in Appendix A.<br />

The Green functi<strong>on</strong> in Eq.(2.21) corresp<strong>on</strong>ds to free boundary c<strong>on</strong>diti<strong>on</strong>s<br />

at infinity <str<strong>on</strong>g>and</str<strong>on</strong>g> preserves the cylindrical symmetry of the metric. It differs from<br />

the familiar result in flat space by a n<strong>on</strong>-linear radial stretch corresp<strong>on</strong>ding to a<br />

smooth deformati<strong>on</strong> of the bump into a flat disk. This Green’s functi<strong>on</strong> determines an


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 31<br />

attractive interacti<strong>on</strong> for the defect dipole pair. However, the Gaussian curvature of<br />

the bump also generates a geometric potential that tries to pull the disclinati<strong>on</strong> dipole<br />

apart. This geometric interacti<strong>on</strong> arises by combining cross terms between s(u) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

G(u) in Eq.(E.19) with the positi<strong>on</strong> dependent self-interacti<strong>on</strong>s derived in Appendix<br />

C. The resulting interacti<strong>on</strong> FG between defects <str<strong>on</strong>g>and</str<strong>on</strong>g> the Gaussian curvature takes<br />

the simple form<br />

Nd �<br />

FG = KA qi<br />

i=1<br />

where the geometrical potential V (u) is defined as<br />

�<br />

V (u) ≡ −<br />

�<br />

1 − qi<br />

�<br />

V (ui) , (2.23)<br />

4π<br />

dA G(u ′ ) Γ(u, u ′ ) . (2.24)<br />

The minus sign in fr<strong>on</strong>t of this geometric potential insures that defects of topological<br />

charge between zero <str<strong>on</strong>g>and</str<strong>on</strong>g> 4π are attracted by regi<strong>on</strong>s of positive Gaussian curvature<br />

[43]. For defects with a large topological charge of q > 4π, the sign of the geometric<br />

interacti<strong>on</strong>, FG, is reversed <str<strong>on</strong>g>and</str<strong>on</strong>g>, <str<strong>on</strong>g>and</str<strong>on</strong>g> defects of either sign are pushed away from the<br />

bump. This scenario does not affect the geometry-driven defect formati<strong>on</strong> discussed<br />

in this paper that relies <strong>on</strong> disclinati<strong>on</strong>s whose charge is well below 4π. However, as<br />

a result of the positi<strong>on</strong>-dependent self interacti<strong>on</strong>s, FG is no l<strong>on</strong>ger symmetric under<br />

the change q → −q, as <strong>on</strong>e would expect <strong>on</strong> the basis of the electrostatic analogy. The<br />

effect of this asymmetry is small for hexatic order but it increase for liquid crystals<br />

with p-fold order parameter, as p decreases (see Ref. [43], <str<strong>on</strong>g>and</str<strong>on</strong>g> references therein).<br />

Gauss’ law generalized to curved surfaces (see Appendix B) insures that the<br />

geometric force experienced by a defect of charge q with radial coordinate r will be<br />

determined <strong>on</strong>ly by the net curvature enclosed in a circle of radius r centered <strong>on</strong> the


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 32<br />

top of the bump. The resulting ”electric field” is radial as expected from electrostatics<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> is proporti<strong>on</strong>al to the gradient of the geometric potential, which for a Gaussian<br />

bump takes the form (see Appendix B)<br />

� ∞<br />

dr<br />

V (r) = −<br />

r<br />

′<br />

r ′<br />

�� �<br />

l(r ′ ) − 1<br />

. (2.25)<br />

For small values of the aspect ratio α the potential in Eq.(2.25) can be approximated<br />

by<br />

V (r) ≈ − α2 r2<br />

−<br />

r e 2 0<br />

4<br />

. (2.26)<br />

The resulting force is linear for small r (i.e. near the top of the bump) <str<strong>on</strong>g>and</str<strong>on</strong>g> decays like<br />

r2<br />

−<br />

r e 2 0 for r ≫ r0. As the aspect ratio α increases, the force generated by the curvature<br />

can overcome the attractive force binding the defect pair which varies logarithmically<br />

for short distances. As a result, oppositely charged defects that were originally tightly<br />

bound can be separated.<br />

This argument, however, neglects another complicati<strong>on</strong> resulting from the<br />

curvature of the surface: as the aspect ratio is increased, the Green’s functi<strong>on</strong> Γ(u, u ′ )<br />

in Eq.(2.21) <str<strong>on</strong>g>and</str<strong>on</strong>g> hence the force binding the defects together also increases. We<br />

illustrate this point in Fig. 2.4 for the special case of a positive defect pinned right<br />

<strong>on</strong> top of the bump <str<strong>on</strong>g>and</str<strong>on</strong>g> a negative <strong>on</strong>e free to move downhill at r. Up<strong>on</strong> invoking<br />

Equati<strong>on</strong>s (E.19) <str<strong>on</strong>g>and</str<strong>on</strong>g> (2.21), the potential binding the pair, Vpair (r), can be written<br />

down exactly as:<br />

Vpair (r) = KA q2 2π ln<br />

�<br />

r<br />

�<br />

a<br />

− KA q2 � ∞<br />

dr<br />

2π<br />

′<br />

r ′<br />

r<br />

+ 2q 2 Ec<br />

�� �<br />

l(r ′ ) − 1<br />

. (2.27)


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 33<br />

dr<br />

+<br />

+<br />

-<br />

Rz(r) -<br />

Figure 2.4: Effect of changing the aspect ratio of the bump <strong>on</strong> the work needed to pull<br />

apart two oppositely charged defects. Positive <str<strong>on</strong>g>and</str<strong>on</strong>g> negative defects are represented by<br />

open circles with their sign printed. The line elements corresp<strong>on</strong>ding to the projected<br />

length dr for the two aspect ratios are shown.<br />

The first term is the flat space Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> we have added two disclinati<strong>on</strong><br />

core energies. The last term represents a ”curvature correcti<strong>on</strong>” that has the same<br />

functi<strong>on</strong>al form of the geometric potential in Eq.(2.25), but it represents a distinct<br />

c<strong>on</strong>tributi<strong>on</strong> to the total free energy. As discussed in Appendix B, the pair potential<br />

energy, Vpair (r), can be understood by applying a generalized Gauss’ Law to the bump<br />

to determine a force which is equal to − q2<br />

. The potential follows by integrating<br />

2πr<br />

this force al<strong>on</strong>g the bump with the length element dr � l(r). Although the force is<br />

independent of the aspect ratio, the length element grows with α (see Fig. 2.4), which<br />

makes the pair more energetically bound for larger values of α. A careful calculati<strong>on</strong><br />

of these effects (including the c<strong>on</strong>tributi<strong>on</strong> associated with the positi<strong>on</strong>-dependent<br />

self-energies) reveals that the geometric force still overcomes the binding interacti<strong>on</strong><br />

r


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 34<br />

for sufficiently large values of α.<br />

An estimate of the critical value of α for which the dipole unbinds can<br />

be obtained by comparing the minimal free energy of the smooth frustrated field<br />

arising from Eq.(2.17) with the free energy in the presence of defects. The latter, as<br />

follows from Eq.(E.19), is composed of three c<strong>on</strong>tributi<strong>on</strong>s: the interacti<strong>on</strong>s am<strong>on</strong>g<br />

the defects, the interacti<strong>on</strong> between the defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature as given by Eq.(2.23)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the Gaussian curvature self-interacti<strong>on</strong>. The latter is equal to the minimal free<br />

energy of the smooth frustrated field <str<strong>on</strong>g>and</str<strong>on</strong>g> is renormalized at finite temperature in<br />

the same way. (Associated with the cutoff is a microscopic core energy, Ec, that we<br />

expect to be independent of the defect positi<strong>on</strong> <strong>on</strong> the bump, as l<strong>on</strong>g as the radius<br />

of curvature is much greater than any microscopic length scale.) The remaining two<br />

c<strong>on</strong>tributi<strong>on</strong>s are not renormalized by thermally induced spin wave fluctuati<strong>on</strong>s [35].<br />

As shown in Appendix C, the difference in free energy of a defected texture described<br />

by Eq.(E.19) relative to the defect-free result Eq.(2.17) can be written as:<br />

∆F (α)<br />

KA<br />

Nd �<br />

+<br />

i=1<br />

qi<br />

= 1<br />

2<br />

Nd Nd �<br />

i=1<br />

�<br />

qiqjΓa(ri, φi, rj, φj)<br />

j�=i<br />

�<br />

1 − qi<br />

�<br />

V (ri) +<br />

4π<br />

Ec<br />

KA<br />

Nd �<br />

i=1<br />

q 2 i . (2.28)<br />

where we have assumed overall charge neutrality for the defect c<strong>on</strong>figurati<strong>on</strong>. The<br />

subscript in Γa indicates that a c<strong>on</strong>stant microscopic core radius a has been absorbed<br />

in the definiti<strong>on</strong> of the Green functi<strong>on</strong> so that the argument of the logarithm in<br />

Eq.(2.21) becomes dimensi<strong>on</strong>less, as in Eq.(C.21),<br />

Γa(ri, φi, rj, φj) = Γ(ri, φi, rj, φj) + 1<br />

ln a . (2.29)<br />


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 35<br />

This microscopic cutoff a corresp<strong>on</strong>ds to a c<strong>on</strong>stant core radius for each defect <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

it is of the order of the spacing between the microscopic degrees of freedom. The<br />

sum of microscopic core energies in the fourth term of Eq.(2.28) needs to be fixed<br />

phenomenologically or from models that go bey<strong>on</strong>d simple elasticity theory [47]. Note<br />

that both Γa(ri, φi, rj, φj) <str<strong>on</strong>g>and</str<strong>on</strong>g> V (r) depend <strong>on</strong> α. If α becomes sufficiently large so<br />

that ∆F is less than zero, <strong>on</strong>e or more disclinati<strong>on</strong> dipoles unbind in the hexatic phase<br />

at a sequence of critical values αci . The analogous defects in XY-model textures of<br />

tilted liquid crystal molecules would be +/− vortex pairs.<br />

2.3 Curvature induced defect generati<strong>on</strong><br />

2.3.1 Onset of the defect-dipole instability<br />

If a dipole is created, say, al<strong>on</strong>g the line r = r0 of zero Gaussian curvature,<br />

the positive disclinati<strong>on</strong> will be pulled towards the center by the positive curvature<br />

while the negative <strong>on</strong>e will be repelled into the regi<strong>on</strong> of negative curvature. The net<br />

result is a reducti<strong>on</strong> of the total free energy of the order of the depth of the potential<br />

well since the logarithmic binding energy is approximately c<strong>on</strong>stant compared to the<br />

geometric potential. An approximate analytical treatment is obtained by assuming<br />

that the positive defect sits right at the center of the bump <str<strong>on</strong>g>and</str<strong>on</strong>g> the negative <strong>on</strong>e at a<br />

distance of the order of r0. The validity of this approximati<strong>on</strong> scheme can be checked<br />

by numerically minimizing the energy with respect to the positi<strong>on</strong> of the defects, as<br />

discussed in Secti<strong>on</strong> (2.3.2). We assume charge neutrality so that the two defects<br />

have equal <str<strong>on</strong>g>and</str<strong>on</strong>g> opposite topological charges of magnitude q. For order parameters


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 36<br />

V(r/r0 )<br />

0<br />

-0.5<br />

-1<br />

-1.5<br />

-2<br />

-2.5<br />

-3<br />

-3.5<br />

0.25 0.5 0.75 1 1.25 1.5 1.75 2<br />

Figure 2.5: Geometric potential V (r/r0) as a functi<strong>on</strong> of the dimensi<strong>on</strong>less ratio of r<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> r0 for α = 1, 2, 3, 4, 5. Note that | V (r) |≪ | V (0) | for r � r0. The arrow points<br />

in the directi<strong>on</strong> of increasing α.<br />

with a p-fold symmetry, the minimal topological charge is q = 2π.<br />

The approximate<br />

p<br />

free energy cost to generate this defect pair then follows from Equati<strong>on</strong>s (2.25),(2.27)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> (2.28):<br />

∆F (α)<br />

KA<br />

r/r0<br />

≈ q2<br />

� �<br />

r<br />

� � �<br />

ln + V (r) + q 1 −<br />

2π a<br />

q<br />

�<br />

V (0)<br />

4π<br />

�<br />

− q 1 + q<br />

�<br />

2 Ec<br />

V (r) + 2q . (2.30)<br />

4π<br />

KA<br />

The internal c<strong>on</strong>sistency of the formalism can be checked by investigating the limit<br />

r → a. As the negative defect approaches the positive <strong>on</strong>e at the center of the bump,<br />

we have V (a) ≈ V (0) <str<strong>on</strong>g>and</str<strong>on</strong>g> the energy tends to the expected flat space result 2q 2 Ec.<br />

The equilibrium positi<strong>on</strong> of the negative defect turns out to be for r equal<br />

to a few r0, so the terms c<strong>on</strong>taining V (r) in Eq.(2.30) can be dropped as a first<br />

approximati<strong>on</strong> because V (r) decays exp<strong>on</strong>entially (see Fig. 2.5):<br />

∆F (α)<br />

KA<br />

≈ q2<br />

2π ln<br />

�<br />

r<br />

� �<br />

+ q 1 −<br />

a<br />

q<br />

�<br />

2 Ec<br />

V (0) + 2q . (2.31)<br />

4π<br />

KA


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 37<br />

V(0)<br />

0<br />

-1<br />

- 2<br />

- 3<br />

- 4<br />

- 5<br />

- 6<br />

1 2 3 4 5<br />

Figure 2.6: Plot of the geometric potential evaluated at the center of the bump, V (0),<br />

for aspect ratios α between 0 <str<strong>on</strong>g>and</str<strong>on</strong>g> 5. The c<strong>on</strong>tinuous line is plotted using the exact<br />

form of V (r) while the dashed line is obtained from the low α expansi<strong>on</strong> given in<br />

Eq.(2.26).<br />

To estimate the critical value of the aspect ratio for which the first dipole unbinds,<br />

let r ∼ r0 <str<strong>on</strong>g>and</str<strong>on</strong>g> solve for αc in Eq.(2.31). Because a different choice for the core energy<br />

Ec can be accounted for by rescaling the core size a in Eq.(2.31), the c<strong>on</strong>diti<strong>on</strong> for<br />

unbinding is<br />

with<br />

If we know Ec<br />

KA<br />

|V (0)| ><br />

2q<br />

(4π − q) ln<br />

�<br />

r0<br />

a ′<br />

�<br />

, (2.32)<br />

a ′ = a e<br />

− 4πEc<br />

K A . (2.33)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> r0<br />

a , the critical aspect ratio αc can be obtained from inspecti<strong>on</strong><br />

of Fig. 2.6 where V (0) is plotted as a functi<strong>on</strong> of α. The Taylor expansi<strong>on</strong> of V (r)<br />

derived in Eq.(2.26) gives V (0) ≈ − α2<br />

4<br />

to leading order in α. Inspecti<strong>on</strong> of Fig. 2.6<br />

α


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 38<br />

shows that this approximati<strong>on</strong> works sufficiently well even for aspect ratios of order<br />

unity. Up<strong>on</strong> substituting for V (0) into Eq.(2.32), we obtain an estimate of how αc<br />

depends <strong>on</strong> r0<br />

a ′ :<br />

α 2 c ≈<br />

≈<br />

8q<br />

(4π − q) ln<br />

�<br />

r0<br />

a ′<br />

�<br />

�<br />

8q<br />

� �<br />

r0<br />

ln +<br />

(4π − q) a<br />

4πEc<br />

�<br />

. (2.34)<br />

KA<br />

Note that a critical height hc = αcr0 for defect unbinding is predicted for fixed r0<br />

a ′<br />

with defect charge q = 2π<br />

p<br />

relati<strong>on</strong> is tested in Secti<strong>on</strong> 2.3.2.<br />

for all integer values of p. The validity of this approximate<br />

The c<strong>on</strong>tinuum theory adopted here is valid in the limit r0 ≫ a. If Ec can<br />

be neglected compared to KA, the defect unbinding instability is triggered when the<br />

energy gain derived from letting the defects screen the Gaussian curvature (approx-<br />

imately given by qV (0)) overcomes the work needed to pull them apart a distance<br />

r0. This work, of order q2<br />

2π ln � �<br />

r0<br />

r0<br />

, increases very slowly with large , hence the c<strong>on</strong>-<br />

a<br />

a<br />

tinuum approximati<strong>on</strong> can be satisfied while keeping the work finite. Note that the<br />

result does not depend <strong>on</strong> the size of the system R because we assume overall discli-<br />

nati<strong>on</strong> charge neutrality <str<strong>on</strong>g>and</str<strong>on</strong>g> the assumpti<strong>on</strong> that R ≫ r0. In this limit, boundary<br />

effects can be ignored provided that they do not impose a topological c<strong>on</strong>straint <strong>on</strong><br />

the phase of the order parameter. An aligning outer wall in a circular hexatic sample,<br />

for example, would force the b<strong>on</strong>d angle field to rotate by 2π, leading to six defects in<br />

the ground state even in flat space. The interesting physics which results is addressed<br />

in Secti<strong>on</strong> 2.4.


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 39<br />

2.3.2 Numerical investigati<strong>on</strong> of defect-unbinding transiti<strong>on</strong>s<br />

The disclinati<strong>on</strong> unbinding transiti<strong>on</strong>s can be investigated more quantita-<br />

tively by minimizing numerically ∆F (α) in Eq.(2.28) with respect to the positi<strong>on</strong>s of<br />

the defects. The aspect ratio above which ∆F (α) becomes negative corresp<strong>on</strong>ds to<br />

the threshold value αc (analogous to a first order transiti<strong>on</strong>) for which the singular<br />

field is energetically favored with respect to the smooth texture of Fig. 2.3. We<br />

emphasize that the energy l<str<strong>on</strong>g>and</str<strong>on</strong>g>scape can have two minima. The first occurs when<br />

two oppositely charged defects form a closely bound dipole (with separati<strong>on</strong> of the<br />

order of the cutoff a) <str<strong>on</strong>g>and</str<strong>on</strong>g> hence annihilate each other leaving a smooth texture. The<br />

sec<strong>on</strong>d minimum corresp<strong>on</strong>ds to an unbound pair (with separati<strong>on</strong> of a few r0) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

it disappears when the geometric force is too weak to overcome the binding force of<br />

the pair. This scenario occurs for a finite value of α characteristic of the geometry<br />

of the substrate above which the formati<strong>on</strong> of an unbound dipole is possible, albeit<br />

energetically unfavored (see Fig. 2.7). As α is increased above αc, the smooth-texture<br />

minimum becomes metastable <str<strong>on</strong>g>and</str<strong>on</strong>g> the unbinding of a defect pair is the most likely<br />

scenario.<br />

It is useful to parameterize ∆F (α) in terms of the dimensi<strong>on</strong>less radial co-<br />

ordinates ¯ri ≡ ri . The geometric potential is defined in Eq.(A.9) as a functi<strong>on</strong> of ¯ri,<br />

r0<br />

V (r) ≡ ˜ Vα( r ), where<br />

r0<br />

� ∞<br />

˜Vα(x)<br />

dy<br />

�� �<br />

= −<br />

1 + α2y2 exp(−y2 ) − 1<br />

x y<br />

. (2.35)<br />

In order to write the defect-defect interacti<strong>on</strong> in terms of the dimensi<strong>on</strong>less radial


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 40<br />

- 0.15<br />

- 0.2<br />

- 0.25<br />

- 0.3<br />

0.5<br />

1<br />

1.5<br />

Figure 2.7: Plot of ∆F/KA versus x1 <str<strong>on</strong>g>and</str<strong>on</strong>g> x2 the positi<strong>on</strong>s of the negative <str<strong>on</strong>g>and</str<strong>on</strong>g> pos-<br />

itive defects respectively in units of r0, for α = 1.2. A c<strong>on</strong>stant energy offset equal<br />

to − q2<br />

2π ln � r0<br />

a ′<br />

�<br />

has been neglected. The metastable minimum at x1 � 1.3 r0 <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

x2 � 0.2 r0 corresp<strong>on</strong>ds to an unbound pair (see Fig. 2.8a). As α is decreased<br />

further the energy barrier that separates this minimum from the smooth texture soluti<strong>on</strong><br />

(corresp<strong>on</strong>ding to x1 approaching x2) disappears <str<strong>on</strong>g>and</str<strong>on</strong>g> the two opposite defects<br />

annihilate.<br />

2<br />

2.5<br />

coordinate ¯ri, we introduce a new functi<strong>on</strong> ˜ ℜ(¯ri) defined by<br />

˜ℜ(¯ri) ≡ ri<br />

exp[V (ri)] = ri<br />

r0<br />

= ℜ(ri)<br />

r0<br />

r0<br />

�<br />

exp ˜Vα<br />

� ri<br />

r0<br />

0.1<br />

��<br />

0.2<br />

, (2.36)<br />

where Eq.(A.8) was used in the last step. We can now transform Γa(¯ri, φi, ¯rj, φj) by<br />

eliminating ℜ(ri) in favor of ˜ ℜ(¯ri). Thus we have using Equati<strong>on</strong>s (C.21) <str<strong>on</strong>g>and</str<strong>on</strong>g> (A.12)<br />

Nd<br />

Nd<br />

�<br />

�<br />

− qiqjΓa(ri, φi, rj, φj) = − qiqjΓ(¯ri, φi, ¯rj, φj)<br />

j�=i<br />

+ 1<br />

2π<br />

j�=i<br />

Nd �<br />

i=1<br />

q 2 i ln( r0<br />

) , (2.37)<br />

a<br />

where we have exploited charge neutrality <str<strong>on</strong>g>and</str<strong>on</strong>g> Eq.(C.22). The free energy minimized


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 41<br />

�<br />

∆F<br />

with respect to the positi<strong>on</strong>s of the defects, min<br />

to Eq.(2.28), as<br />

where<br />

� �<br />

∆F<br />

min<br />

KA<br />

= f(α) + 1<br />

4π<br />

f(α) ≡ min [ 1<br />

2<br />

+<br />

Nd �<br />

qi<br />

i=1<br />

Nd �<br />

j�=i<br />

Nd �<br />

i=1<br />

KA<br />

q 2 i ln<br />

�<br />

, can now be written, according<br />

�<br />

r0<br />

a ′<br />

�<br />

qiqjΓ(¯ri, φi, ¯rj, φj)<br />

, (2.38)<br />

�<br />

1 − qi<br />

�<br />

V (¯ri)] . (2.39)<br />

4π<br />

Note that in the sec<strong>on</strong>d term in Eq.(2.38) the core energy of each defect has been<br />

absorbed in the modified core radius a ′ , as defined in Eq.(2.33). This generates an<br />

energy cost for unbinding that can be overcome if f(α) assumes sufficiently large<br />

negative values. Hence it is sufficient to study numerically how ∆F (α) varies as a<br />

functi<strong>on</strong> of a single parameter, eg. r0<br />

a ′ . As an illustrati<strong>on</strong> of this approach, we study<br />

explicitly the unbinding of <strong>on</strong>e <str<strong>on</strong>g>and</str<strong>on</strong>g> two disclinati<strong>on</strong>s pairs leading to the ground states<br />

represented schematically in Fig. 2.8. The smooth ground state becomes unstable<br />

to the formati<strong>on</strong> of <strong>on</strong>e defect dipole first. The critical aspect ratio above which this<br />

scenario occurs can be determined with the aid of Fig. 2.9 where the functi<strong>on</strong> f(α)<br />

introduced in Eq.(2.39) is plotted as a functi<strong>on</strong> of the aspect ratio. From Eq.(2.38)<br />

we see that αc1 is determined by<br />

f(αc1) = − q2<br />

2π ln<br />

�<br />

r0<br />

a ′<br />

�<br />

. (2.40)<br />

For r0<br />

a ′ = 10 4 <str<strong>on</strong>g>and</str<strong>on</strong>g> q = 2π<br />

6 , we obtain a critical aspect ratio αc1 ≈ 3.2. As a comparis<strong>on</strong>,<br />

the approximate c<strong>on</strong>diti<strong>on</strong> derived in Eq.(2.32) gives αc1 ≈ 3 when used in c<strong>on</strong>juncti<strong>on</strong>


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 42<br />

(a)<br />

(b)<br />

+<br />

r 0<br />

- + +<br />

-<br />

Figure 2.8: The equilibrium defects positi<strong>on</strong>s are illustrated schematically in the case<br />

of <strong>on</strong>e (a) <str<strong>on</strong>g>and</str<strong>on</strong>g> two dipoles (b). We assume free boundary c<strong>on</strong>diti<strong>on</strong>s at infinity, as in<br />

Fig. 2.3, so that the effect of image charges can be neglected.<br />

with Fig. 2.6. The rougher estimate in Eq.(2.34) leads (for q = 2π<br />

6 ) to αc1 ≈ 2.6. This<br />

discrepancy is easily understood c<strong>on</strong>sidering that Eq.(2.34) was derived by means of<br />

a low α expansi<strong>on</strong>. The critical aspect ratio αc1 is too low for the two-dipole defect<br />

c<strong>on</strong>figurati<strong>on</strong> to become energetically favorable with respect to the smooth ground<br />

state. Indeed, inspecti<strong>on</strong> of Fig. 2.10 reveals that the critical aspect ratio αc2 for<br />

which the ”two-dipole instability” sets in is approximately equal to 3.6 for the same<br />

choice of parameters used in the single pair case. Note that, in the presence of two<br />

dipoles, the energy cost arising from the sec<strong>on</strong>d term in Eq.(2.38) is twice as large<br />

because there are four defects rather than two. However, for α � 4.2 generating<br />

two dipoles becomes more energetically favored than a single dipole (see Fig. 2.11).<br />

The approach illustrated here can be used to calculate a cascade of defect unbinding<br />

r 0<br />

x<br />

-


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 43<br />

0<br />

f(!)<br />

- 1<br />

- 2<br />

- 3<br />

- 4<br />

2 3 !c1 4 5<br />

Figure 2.9: Plot of f(α) versus the aspect ratio α obtained by minimizing over the<br />

single-dipole defect c<strong>on</strong>figurati<strong>on</strong> represented schematically in Fig. 2.8a. As discussed<br />

in the text, the first unbinding transiti<strong>on</strong> occurs when f(αc1) is equal to − q2<br />

2π ln � r0<br />

a ′<br />

�<br />

.<br />

The value αc1 is indicated by the dashed line for r0<br />

a ′ = 104 <str<strong>on</strong>g>and</str<strong>on</strong>g> q = 2π . Note that no<br />

6<br />

minimum (corresp<strong>on</strong>ding to an unbound pair) exists for α less than 1 (approximately),<br />

hence the curve cannot be c<strong>on</strong>tinued to the origin (see Fig 2.7).<br />

instabilities at critical aspect ratios αci<br />

involving higher number of dipoles <str<strong>on</strong>g>and</str<strong>on</strong>g> their<br />

equilibrium c<strong>on</strong>figurati<strong>on</strong>s in the ground state. Note that the unbinding eventually<br />

stops since the integrated Gaussian curvature in the top cup cannot exceed 2π. We<br />

expect the qualitative features of our analysis to be independent of the exact shape<br />

of the bumpy substrate although the specific values αci<br />

!<br />

depend <strong>on</strong> the geometry <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

the choice of the microscopic parameters a <str<strong>on</strong>g>and</str<strong>on</strong>g> Ec. Finally, we emphasize that the<br />

curvature-induced unbinding is similar to a first order transiti<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> occurs for rather<br />

pr<strong>on</strong>ounced deviati<strong>on</strong>s from flatness (ie. large α).


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 44<br />

0<br />

f(!)<br />

- 1<br />

- 2<br />

-<br />

-<br />

-<br />

-<br />

-<br />

3<br />

4<br />

5<br />

6<br />

7<br />

!c2<br />

2 3 4 5<br />

Figure 2.10: Plot of f(α) versus the aspect ratio α obtained by minimizing the twodipole<br />

defect c<strong>on</strong>figurati<strong>on</strong> represented schematically in Fig. 2.8b . The sec<strong>on</strong>d<br />

unbinding transiti<strong>on</strong> occurs when f(αc2) is equal to − q2<br />

π ln � r0<br />

a ′<br />

�<br />

. The value αc2 is<br />

indicated by the dashed line for r0<br />

a ′ = 104 <str<strong>on</strong>g>and</str<strong>on</strong>g> q = 2π (compare with Fig. 2.9).<br />

6<br />

2.3.3 Single vortex instability<br />

The unbinding of defect pairs may not be the most likely scenario if the size<br />

of the system R is sufficiently small. In this case, the creati<strong>on</strong> of a single vortex at<br />

the center of the bump may become energetically favorable for lower aspect ratios<br />

than required by the defect dipole instability. The equati<strong>on</strong> for the b<strong>on</strong>d angle field<br />

θs(u) for a single defect of charge q at the center of the bump is given by:<br />

θs(φ) =<br />

�<br />

q<br />

�<br />

− 1 φ , (2.41)<br />

2π<br />

where the b<strong>on</strong>d angle is measured with respect to the rotating basis vectors corre-<br />

sp<strong>on</strong>ding to the polar coordinates discussed in Secti<strong>on</strong> 2.2.2. Up<strong>on</strong> substituting θs(φ)<br />

in Eq.(2.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> subtracting the free energy F0 corresp<strong>on</strong>ding to the defect-free texture<br />

!


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 45<br />

2<br />

1<br />

0<br />

- 1<br />

- 2<br />

- 3<br />

!F(")<br />

"c1<br />

"c2<br />

2 2.5 3 3.5 4 4.5 5 5.5<br />

∆F (α)<br />

Figure 2.11: Plot of versus α corresp<strong>on</strong>ding to a single dipole (c<strong>on</strong>tinuous line)<br />

KA<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> two dipoles (dotted line) for r0<br />

a ′ = 104 . The critical aspect ratios αc1 <str<strong>on</strong>g>and</str<strong>on</strong>g> αc2<br />

are indicated by dashed lines. Note that the aspect ratio for which the two dipole<br />

c<strong>on</strong>figurati<strong>on</strong> becomes energetically favored occurs for α > 4.2.<br />

we obtain:<br />

∆F (α)<br />

KA<br />

= q2<br />

4π ln<br />

� �<br />

ℜ(R)<br />

a<br />

"<br />

�<br />

+ q 1 − q<br />

�<br />

V (0)<br />

4π<br />

2 Ec<br />

+ q , (2.42)<br />

KA<br />

where Ec was added by h<str<strong>on</strong>g>and</str<strong>on</strong>g>. The same result is obtained by using the more general<br />

formalism developed in Appendix D. Indeed, by letting the positi<strong>on</strong> of an isolated<br />

defect tend to the center of the bump in Eq.(D.24) we obtain the energy of the singular<br />

field in the case of free boundary c<strong>on</strong>diti<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> the result matches Eq.(2.42). As<br />

discussed in Appendix D, a defect located at ri is attracted to the boundary at R<br />

for free boundary c<strong>on</strong>diti<strong>on</strong>s. One can think of this interacti<strong>on</strong> as resulting from an<br />

image defect of opposite sign behind the edge of the sample at positi<strong>on</strong> r ′ i such that


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 46<br />

the following relati<strong>on</strong> holds in terms of the c<strong>on</strong>formal radius ℜ(r ′ )<br />

ℜ(r ′ i) = ℜ(R)2<br />

ℜ(ri)<br />

. (2.43)<br />

This result can be understood by analogy to the familiar electrostatic problem of a<br />

charged line located a distance ri from the center of a cylindrical grounded c<strong>on</strong>ductor<br />

whose axis is parallel to it [48]. The analogy becomes precise if <strong>on</strong>e lets ri → ℜ(ri)<br />

as explained in Appendix D.<br />

If the geometric potential is not str<strong>on</strong>g enough (as in the flat space limit<br />

α = 0), the defect will migrate to the edge of the sample <str<strong>on</strong>g>and</str<strong>on</strong>g> annihilate with its<br />

image leaving a smooth field. On the other h<str<strong>on</strong>g>and</str<strong>on</strong>g>, when the aspect ratio is sufficiently<br />

large, the defect can lower its energy by sitting at the center of the bump. Comparis<strong>on</strong><br />

of Eq.(2.42) with Eq.(2.31) shows that, unless R ≫ r0, the energy of the single vortex<br />

instability will be lower or at least comparable to the unbinding of a defect dipole.<br />

In fact, the threshold αs that α needs to exceed to trigger the single defect instability<br />

is easily obtained if the values of the geometric potential at the origin are tabulated<br />

for different aspect ratios, as illustrated in Fig. 2.5. The c<strong>on</strong>diti<strong>on</strong> for single vortex<br />

generati<strong>on</strong> reads<br />

q<br />

|V (0)| ><br />

(q − 4π) ln<br />

�<br />

R<br />

a ′<br />

�<br />

. (2.44)<br />

Using the same method adopted to derive Eq.(2.34) we obtain an estimate of how αs<br />

depends <strong>on</strong> R<br />

a ′ (compare with Eq.(2.34)):<br />

α 2 4q<br />

s ≈<br />

(4π − q) ln<br />

�<br />

R<br />

a ′<br />

�<br />

. (2.45)


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 47<br />

The single vortex instability is reminiscent of vortex generati<strong>on</strong> in rotating superfluid<br />

helium with α playing the role of the angular speed Ω. For a volume of helium<br />

c<strong>on</strong>tained in a cylindrical vessel of radius R <str<strong>on</strong>g>and</str<strong>on</strong>g> rotating uniformly with c<strong>on</strong>stant<br />

angular speed, the critical value Ωc1 above which defect generati<strong>on</strong> occurs is given by<br />

[49]<br />

where K = 2π�<br />

mHe<br />

Ωc1 ≈ K<br />

� �<br />

R<br />

ln , (2.46)<br />

2πR2 a<br />

is the magnitude of the quantum of circulati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> a the core radius<br />

2 . Note that Ωc1 decreases as R increases, unlike αs which diverges logarithmically.<br />

Thus, the single defect instability studied here is a finite size effect. In c<strong>on</strong>trast, the<br />

disclinati<strong>on</strong> unbinding studied earlier in this secti<strong>on</strong> does not depend <strong>on</strong> the system<br />

size because of charge neutrality. Hence the thermodynamic limit can be safely taken,<br />

provided the characteristic length over which the curvature varies (ie. r0) is not too<br />

large compared to a (see Eq.(2.34)).<br />

In c<strong>on</strong>sidering the case of small system size, it is important to keep in mind<br />

two assumpti<strong>on</strong>s implicit in the present treatment. The radius of curvature r0<br />

α must<br />

be much larger than the core radius everywhere for the c<strong>on</strong>tinuum approach to be<br />

valid, that is r0 ≫ α a. Additi<strong>on</strong>ally, the Gaussian curvature must be vanishing small<br />

at the edge of the system which requires R to be larger than a few r0.<br />

2 A similar mechanism applies to superc<strong>on</strong>ductors in a uniform magnetic field.


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 48<br />

2.3.4 Lattice of bumps, valleys <str<strong>on</strong>g>and</str<strong>on</strong>g> saddle points<br />

In some experimental realizati<strong>on</strong>s perhaps modelled <strong>on</strong> those of Ref. [30]<br />

the topography will be periodic. In this Secti<strong>on</strong> we discuss qualitatively how the<br />

results described above generalize to a 2-dimensi<strong>on</strong>al lattice of bumps with variable<br />

aspect ratio for both square <str<strong>on</strong>g>and</str<strong>on</strong>g> triangular lattices. A more quantitative approach to<br />

this problem would involve finding c<strong>on</strong>formal set of coordinates for periodic boundary<br />

c<strong>on</strong>diti<strong>on</strong>s. This is possible in principle but more involved since cylindrical symmetry<br />

is now lost. N<strong>on</strong>etheless, the intuiti<strong>on</strong> gained by studying the single bump allows us<br />

to make some guesses for the ground state. We first note that the geometric potential<br />

generated by the lattice of bumps is not simply the superpositi<strong>on</strong> of results for single<br />

bump potentials. This is caused by the n<strong>on</strong> linear relati<strong>on</strong> between the surface height<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the Gaussian curvature acting as a source for the geometric potential. To explore<br />

this point further, c<strong>on</strong>sider what happens when four bumps are placed at the vertices<br />

of a square. At the center of the square a minimum of the height functi<strong>on</strong> occurs<br />

corresp<strong>on</strong>ding to a new regi<strong>on</strong> of positive Gaussian curvature. This effect is particu-<br />

larly acute for r0 ≤ L, where L is the bump spacing. In general interference between<br />

bumps creates a dual lattice of valleys. A similar breakdown of the superpositi<strong>on</strong><br />

principle arises for triangular lattices.<br />

As the aspect ratio of hilly l<str<strong>on</strong>g>and</str<strong>on</strong>g>scapes such as those shown in Fig. 2.12 is<br />

increased, defects can be created to screen the Gaussian curvature. Their positi<strong>on</strong>s<br />

can be guessed by c<strong>on</strong>sidering a unit cell of the lattice such that the integrated Gaus-<br />

sian curvature vanishes. For a square lattice, we c<strong>on</strong>jecture that the first topography


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 49<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-3 -2.5 -2 -1.5 -1 -0.5 0<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

1<br />

0.75<br />

0.5<br />

0.25<br />

0<br />

-3<br />

0<br />

-3 -2.5 -2 -1.5 -1 -0.5 0<br />

-2<br />

-1<br />

0 0<br />

1<br />

2<br />

3<br />

3<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

2<br />

1<br />

0<br />

-1<br />

-5<br />

-6 -4 -2 0 2 4 6<br />

-6 -4 -2 0 2 4 6<br />

Figure 2.12: (Top) Ground states for square (left) <str<strong>on</strong>g>and</str<strong>on</strong>g> triangular (right) arrays of<br />

bumps. The first <str<strong>on</strong>g>and</str<strong>on</strong>g> sec<strong>on</strong>d rows corresp<strong>on</strong>d to moderate values of the aspect ratio<br />

α respectively. For simplicity, we assume that r0, the bump width is comparable to<br />

the lattice spacing. Positive defects (red dots) ”screen” regi<strong>on</strong>s of positive Gaussian<br />

curvature while negative <strong>on</strong>es (blue dots) are located <strong>on</strong> the saddles of the ”hilly”<br />

l<str<strong>on</strong>g>and</str<strong>on</strong>g>scape.<br />

0<br />

5<br />

-5<br />

0<br />

5


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 50<br />

induced transiti<strong>on</strong> is associated with the appearance of positive defects at the top<br />

of the bumps <str<strong>on</strong>g>and</str<strong>on</strong>g> negative <strong>on</strong>es half way between them in the vertical or horiz<strong>on</strong>tal<br />

directi<strong>on</strong> (see Fig. 2.12). This two-fold degeneracy is compatible with the symme-<br />

try of the lattice <str<strong>on</strong>g>and</str<strong>on</strong>g> analogous to the freedom in choosing the axis al<strong>on</strong>g which the<br />

first disclinati<strong>on</strong>-dipole appears <strong>on</strong> the single bump. The negative defects are shared<br />

between two adjacent cells while the positive <strong>on</strong>es are shared am<strong>on</strong>g four cells thus<br />

ensuring overall charge neutrality. As the value of α increases even more, <strong>on</strong>e might<br />

expect an additi<strong>on</strong>al positive defect appears in the valley located at the center of<br />

each cell <str<strong>on</strong>g>and</str<strong>on</strong>g> two additi<strong>on</strong>al negative defects shared with the adjacent cells are cre-<br />

ated between the bumps at right angles to the directi<strong>on</strong> discussed above (see Fig.<br />

2.12).<br />

For the triangular lattice, we c<strong>on</strong>jecture that the first transiti<strong>on</strong> corresp<strong>on</strong>ds<br />

to positive defects <strong>on</strong> top of the bumps <str<strong>on</strong>g>and</str<strong>on</strong>g> negative <strong>on</strong>es between the bumps al<strong>on</strong>g<br />

<strong>on</strong>e of the three axis of symmetry of the unit cell. As the value of the aspect ratio<br />

is increased, additi<strong>on</strong>al positive defects appear <strong>on</strong> the six minima of the surface <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

negative <strong>on</strong>es are generated al<strong>on</strong>g the remaining two axes of symmetry of the unit<br />

cell (see Fig. 2.12). A simple count of the total defect charges enclosed in the unit<br />

cell shows that this scenario also satisfies the requirement of defect charge neutrality.<br />

2.4 Defect Dec<strong>on</strong>finement<br />

With potential experiments in mind [37], it is interesting to c<strong>on</strong>sider the<br />

case of hexatic order <strong>on</strong> a bump encircled by a circular wall of radius R ≫ r0 which


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 51<br />

aligns the hexatic b<strong>on</strong>d angles. As a simple model, imagine an array of hexag<strong>on</strong>s<br />

which locally achieve a comm<strong>on</strong> orientati<strong>on</strong> tangential to the wall (see Fig. D.1b).<br />

The hexatic order parameter will thus rotate by 2π up<strong>on</strong> making a circuit of the<br />

wall, insuring that at least six defects of ”charge” 2π<br />

6<br />

must be included in the ground<br />

state for all values of the aspect ratio. These boundary-c<strong>on</strong>diti<strong>on</strong> induced defects will<br />

interact with the Gaussian curvature of the bump <str<strong>on</strong>g>and</str<strong>on</strong>g> with the wall. The Nd defects<br />

c<strong>on</strong>tribute large (c<strong>on</strong>stant) self energies of the form �Nd i=1 KAq2 i ln � �<br />

R that dominate<br />

a<br />

the total energy for sufficiently large systems. Since � Nd<br />

i=1 qi must be equal to 2π, the<br />

energy is minimized when the defects split up into the smallest possible charges.<br />

The equilibrium defect c<strong>on</strong>figurati<strong>on</strong> must minimize the free energy taking<br />

into account the c<strong>on</strong>fining potential generated by the Gaussian curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> the<br />

interacti<strong>on</strong>s of the defects with the boundary <str<strong>on</strong>g>and</str<strong>on</strong>g> am<strong>on</strong>g themselves. The repulsive<br />

force exercised by the wall <strong>on</strong> a defect located at ℜ(ri) in the c<strong>on</strong>formal plane can<br />

be computed by placing an image defect of same charge outside the wall at positi<strong>on</strong><br />

ℜ(R) 2<br />

. The mathematics resembles the problem of finding the magnetic field of a line<br />

ℜ(ri)<br />

current located at a given distance ri from the center of a cylinder of high permeability<br />

material <str<strong>on</strong>g>and</str<strong>on</strong>g> whose radius R is greater than ri [48]. The analogy is complete up<strong>on</strong><br />

performing the change of coordinates r → ℜ(r) <str<strong>on</strong>g>and</str<strong>on</strong>g> identifying the gradient of the<br />

b<strong>on</strong>d angle ∂αθ(u) with the magnetic field. This is explained in detail in Appendix C<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> D where we introduce a c<strong>on</strong>jugate functi<strong>on</strong> χ(u) analogous to the vector potential<br />

that simplifies the analysis of this problem. Thus, each of the Nd defects will also<br />

interact with an equal number of image defects. This situati<strong>on</strong> can be described


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 52<br />

mathematically by deriving an appropriate Green’s functi<strong>on</strong> Γ N that includes the<br />

images, as discussed in Appendix D (see Eq.(D.19)). The resulting free energy F N<br />

reads:<br />

F N<br />

KA<br />

= 1<br />

2<br />

+<br />

−<br />

Nd �<br />

j�=i<br />

Nd �<br />

i=1<br />

Nd �<br />

i=1<br />

qiqj Γ N (xi; xj) + F0<br />

qi(1 − qi<br />

4π )V (ri) +<br />

Nd �<br />

i=1<br />

qi 2<br />

4π ln<br />

� �<br />

ℜ(R)<br />

qi 2<br />

4π ln � 1 − x 2� i , (2.47)<br />

where F0 is defined in Eq.(2.17) <str<strong>on</strong>g>and</str<strong>on</strong>g> the Green’s functi<strong>on</strong> Γ N (xi; xj) is given by<br />

Γ N (xi; xj) = − 1<br />

4π ln � x 2 i + x 2 j − 2xixj cos (φi − φj) �<br />

− 1<br />

4π ln � x 2 i x 2 j + 1 − 2xixj cos (φi − φj) � .<br />

a<br />

(2.48)<br />

The last term accounts for the interacti<strong>on</strong> with the image defects <str<strong>on</strong>g>and</str<strong>on</strong>g> the superscript<br />

N indicates Neumann boundary c<strong>on</strong>diti<strong>on</strong>s <strong>on</strong> an appropriate potential functi<strong>on</strong>.<br />

Here, we use scaled coordinates in the c<strong>on</strong>formal plane xi ≡ ℜ(ri)<br />

. The interacti<strong>on</strong> of<br />

ℜ(R)<br />

the defects with the curvature is not affected by the presence of the distant wall.<br />

To provide an illustrati<strong>on</strong> of the combined effect of curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> boundary<br />

c<strong>on</strong>diti<strong>on</strong>s <strong>on</strong> tangential vector order, we first c<strong>on</strong>sider the simpler case of a nematic<br />

order parameter with periodicity equal to π. This simplified model neglects differences<br />

in the elastic c<strong>on</strong>stants for bend <str<strong>on</strong>g>and</str<strong>on</strong>g> splay <str<strong>on</strong>g>and</str<strong>on</strong>g> does not incorporate any effect due to<br />

the uniaxial coupling of the nematogens to the curvature. In this case, minimizati<strong>on</strong><br />

of the logarithmically diverging part of the free energy (fourth term in Eq. 2.47)


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 53<br />

suggests that there will be <strong>on</strong>ly two disclinati<strong>on</strong>s of charge q = π displaced al<strong>on</strong>g a<br />

radial directi<strong>on</strong> (see Fig. 2.13b). By applying Eq.(2.47), we can parameterize the<br />

energy of the system in terms of the scaled radial coordinates, x1 <str<strong>on</strong>g>and</str<strong>on</strong>g> x2 of the two<br />

disclinati<strong>on</strong>s. The resulting energy l<str<strong>on</strong>g>and</str<strong>on</strong>g>scape is plotted in Fig. 2.13 (for α = 2 <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

R = 7r0) <str<strong>on</strong>g>and</str<strong>on</strong>g> clearly reveals two minimal-energy c<strong>on</strong>figurati<strong>on</strong>s. The first minimum<br />

corresp<strong>on</strong>ds to <strong>on</strong>e disclinati<strong>on</strong> c<strong>on</strong>fined at the top of the bump (slightly shifted<br />

from the center) <str<strong>on</strong>g>and</str<strong>on</strong>g> the other at a radial distance approximately 70% of R (see<br />

Fig. 2.13b bottom panel). The sec<strong>on</strong>d minimum corresp<strong>on</strong>ds to a fully dec<strong>on</strong>fined<br />

state with both disclinati<strong>on</strong>s placed symmetrically at approximately 67% of R (see<br />

Fig. 2.13b top panel). As the aspect ratio is raised even further, the saddle in the<br />

energy l<str<strong>on</strong>g>and</str<strong>on</strong>g>scape of Fig. 2.13a becomes a minimum corresp<strong>on</strong>ding to a c<strong>on</strong>figurati<strong>on</strong><br />

in which both disclinati<strong>on</strong>s are c<strong>on</strong>fined in the cup of positive Gaussian curvature by<br />

the geometric potential.<br />

As illustrated in Fig. 2.14, there is a critical value of the aspect ratio,<br />

αD � 1.5, above which it is energetically favorable for the system to have <strong>on</strong>e discli-<br />

nati<strong>on</strong> c<strong>on</strong>fined at the top of the bump. For α < αD the fully dec<strong>on</strong>fined c<strong>on</strong>figu-<br />

rati<strong>on</strong> becomes energetically favorable, but the two minima can still coexist. As α<br />

is decreased even further, the repulsi<strong>on</strong> between the two disclinati<strong>on</strong>s overcomes the<br />

c<strong>on</strong>fining force of the geometric potential <str<strong>on</strong>g>and</str<strong>on</strong>g> makes the sec<strong>on</strong>d minimum in Fig.<br />

2.14a (corresp<strong>on</strong>ding to the partially c<strong>on</strong>fined c<strong>on</strong>figurati<strong>on</strong>) disappear altogether.<br />

This ”spinodal point” occurs for α ≈ 0.9 <strong>on</strong> the Gaussian bump. The specific values<br />

of the critical aspect ratios are geometry dependent, but the generic mechanism of


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 54<br />

F/k A<br />

2<br />

1<br />

0<br />

2<br />

4<br />

x1 6<br />

2<br />

4<br />

6<br />

x2<br />

(a) (b)<br />

Figure 2.13: (a) The free energy for a nematic (double headed vector field) living<br />

<strong>on</strong> a Gaussian bump surrounded by an aligning circular wall is plotted for α = 2 as<br />

a functi<strong>on</strong> of the scaled radial coordinates x1 <str<strong>on</strong>g>and</str<strong>on</strong>g> x2 of the two disclinati<strong>on</strong>s. The<br />

radial coordinates have been scaled by r0 <str<strong>on</strong>g>and</str<strong>on</strong>g> the size of the system is R = 7r0. Note<br />

that the energy plot is symmetric with respect to the line x1 = x2. (b) Schematic<br />

illustrati<strong>on</strong> of the positi<strong>on</strong>s of the two disclinati<strong>on</strong>s (black dots) corresp<strong>on</strong>ding to the<br />

deep energy minima at positi<strong>on</strong>s x1 = 0.04 <str<strong>on</strong>g>and</str<strong>on</strong>g> x2 = 4.9 (or viceversa) <str<strong>on</strong>g>and</str<strong>on</strong>g> to a<br />

shallow minimum at x1 = x2 = 4.7. The two defects are <strong>on</strong> opposite sides of the<br />

bump. The c<strong>on</strong>tinuous line corresp<strong>on</strong>ds to the circular boundary while the dashed<br />

<strong>on</strong>e to the circle of zero Gaussian curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> radius r0 (drawing not to scale).


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 55<br />

-<br />

-<br />

-<br />

0.4<br />

0.2<br />

0<br />

0.2<br />

0.4<br />

0.6<br />

F/k A<br />

! D<br />

0 0.25 0.5 0.75 1 1.25 1.5 1.75 !<br />

Figure 2.14: Plot of the free energy of a nematic (double headed vector field) <strong>on</strong> a<br />

Gaussian bump encircled by an aligning wall as a functi<strong>on</strong> of α. The dotted line<br />

represents the energy of the fully dec<strong>on</strong>fined c<strong>on</strong>figurati<strong>on</strong> in Fig. 2.13b top panel<br />

while the c<strong>on</strong>tinuous line corresp<strong>on</strong>ds to the defect pattern illustrated in the bottom<br />

panel of Fig. 2.13b. The energy of the fully dec<strong>on</strong>fined c<strong>on</strong>figurati<strong>on</strong> is approximately<br />

independent of α because the two disclinati<strong>on</strong>s are far away from the bump.<br />

dec<strong>on</strong>finement depends <strong>on</strong>ly <strong>on</strong> a large separati<strong>on</strong> of the length scales r0 <str<strong>on</strong>g>and</str<strong>on</strong>g> R that<br />

c<strong>on</strong>trol the interacti<strong>on</strong> with the curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> the boundary respectively.<br />

The analysis for the hexatic case is complicated by the fact that more defect<br />

c<strong>on</strong>figurati<strong>on</strong>s are possible when six defects are present. We start by noting that even<br />

in flat space (α = 0) there are two natural low energy defect c<strong>on</strong>figurati<strong>on</strong>s with high<br />

symmetry: the ground state corresp<strong>on</strong>ding to the six defects sitting at the vertices of<br />

an hexag<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> a higher energy state given by a pentag<strong>on</strong>al distributi<strong>on</strong> of defects<br />

with the sixth defect sitting at the center of the circular sample (see Fig. 2.15b). As<br />

the aspect ratio is raised, the pentag<strong>on</strong>al arrangement becomes energetically favored<br />

since it pays to have a defect c<strong>on</strong>fined in the (geometric) potential well at the origin<br />

(see Fig. 2.15a). To study the transiti<strong>on</strong>, it is useful to derive expressi<strong>on</strong>s for the<br />

energy of the two defect c<strong>on</strong>figurati<strong>on</strong>s as a functi<strong>on</strong> of the radius of the outer defect


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 56<br />

ring r. Every defect (except the <strong>on</strong>e at the origin, possibly) has the same scaled<br />

coordinate xi = x <str<strong>on</strong>g>and</str<strong>on</strong>g> the angles between two defects are integer multiples of 2π<br />

n<br />

where n is the number of defects in the outer ring (n = 5 for the pentag<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> n = 6<br />

for the hexag<strong>on</strong>). In this case, the sums involved in the first (interacti<strong>on</strong>) term in<br />

Eq.(2.47) can be efficiently evaluated using the following identity:<br />

1 �n−1<br />

�<br />

ln p<br />

2<br />

2 + 1 − 2p cos( 2πi<br />

n )<br />

�<br />

i<br />

= ln (1 − p n ) − ln (1 − p) . (2.49)<br />

Up<strong>on</strong> using Eq.(2.49) with p = 1 <str<strong>on</strong>g>and</str<strong>on</strong>g> p = x 2 to evaluate the sums arising from the<br />

first <str<strong>on</strong>g>and</str<strong>on</strong>g> the sec<strong>on</strong>d term of the Green’s functi<strong>on</strong> in Eq.(2.48) respectively, we obtain<br />

the free energy FH for the hexag<strong>on</strong>al c<strong>on</strong>figurati<strong>on</strong>:<br />

FH(α)<br />

KA<br />

= − π<br />

6 ln<br />

�� �5 � � �<br />

17<br />

ℜ(r) ℜ(r)<br />

−<br />

−<br />

ℜ(R) ℜ(R)<br />

π<br />

+<br />

ln 6<br />

6 11π π<br />

V (r) +<br />

6 6 ln<br />

� �<br />

ℜ(R)<br />

.<br />

a<br />

(2.50)<br />

The free energy for the pentag<strong>on</strong>al c<strong>on</strong>figurati<strong>on</strong> FP is readily obtained after similar<br />

manipulati<strong>on</strong>s<br />

FP (α)<br />

KA<br />

= − 5π<br />

36 ln<br />

�� �6 � � �<br />

16<br />

ℜ(r) ℜ(r)<br />

−<br />

−<br />

ℜ(R) ℜ(R)<br />

π<br />

ln 2<br />

2<br />

+ 11π 55π π<br />

V (0) + V (r) +<br />

36 36 6 ln<br />

� �<br />

ℜ(R)<br />

. (2.51)<br />

a<br />

Note that these manipulati<strong>on</strong>s are very similar to the <strong>on</strong>es necessary to describe<br />

superfluid helium in a cylinder of radius R [50]. In fact, the superfluid problem is<br />

analogous to the case of hexatic order with free boundary c<strong>on</strong>diti<strong>on</strong>s <strong>on</strong> a circular<br />

boundary of radius R (see Appendix D). The rather unusual form of the argument


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 57<br />

F/k A<br />

- 0.4<br />

- 0.6<br />

- 0.8<br />

- 1<br />

! D<br />

(a)<br />

0 0.5 1 1.5<br />

!<br />

2<br />

R<br />

(b)<br />

Figure 2.15: Plot of the free energy of an hexatic phase (draped <strong>on</strong> the Gaussian bump<br />

encircled by a wall) as a functi<strong>on</strong> of α. The dotted line represents the energy of the<br />

hexag<strong>on</strong>al c<strong>on</strong>figurati<strong>on</strong> illustrated in the top panel <strong>on</strong> the left while the c<strong>on</strong>tinuous<br />

line corresp<strong>on</strong>ds to the pentag<strong>on</strong>al arrangement in the bottom panel. The outer defect<br />

rings in both c<strong>on</strong>figurati<strong>on</strong>s are approximately 90% of R. The critical aspect ratio<br />

αD corresp<strong>on</strong>ding to the dec<strong>on</strong>finement transiti<strong>on</strong> discussed in the text is indicated<br />

by the dashed line.<br />

of the logarithm in Equati<strong>on</strong>s (2.50) <str<strong>on</strong>g>and</str<strong>on</strong>g> (2.51) arises from the sum over the image<br />

defects whose positi<strong>on</strong>s depend n<strong>on</strong>-linearly <strong>on</strong> the positi<strong>on</strong> of the defects themselves.<br />

Minimizati<strong>on</strong> of Equati<strong>on</strong>s (2.50) <str<strong>on</strong>g>and</str<strong>on</strong>g> (2.51) with respect to r fixes the<br />

distance of the outer defects. The resulting minimal energies FP <str<strong>on</strong>g>and</str<strong>on</strong>g> FH are plotted<br />

as a functi<strong>on</strong> of α in Fig. 2.15a. The critical value αD, for which FP < FH can be<br />

easily estimated by realizing that FH is approximately independent of α because the<br />

disclinati<strong>on</strong>s are far from the bump. On the other end, FP decreases with α because<br />

the c<strong>on</strong>fined disclinati<strong>on</strong> is trapped in a potential well whose depth is approximately<br />

given by − 11<br />

144 πα2 (see the sec<strong>on</strong>d term of Eq.(2.51) <str<strong>on</strong>g>and</str<strong>on</strong>g> the low α expansi<strong>on</strong> for<br />

V (0) derived in Eq.(2.26)). The critical aspect ratio, αD, for which the dec<strong>on</strong>finement<br />

transiti<strong>on</strong> occurs can be estimated by setting the depth of this potential well equal to<br />

the energy difference between the hexag<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> pentag<strong>on</strong> c<strong>on</strong>figurati<strong>on</strong>s in flat space.<br />

R<br />

r0<br />

r0


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 58<br />

The latter can be read off from the energy diagram in Fig. 2.15a <str<strong>on</strong>g>and</str<strong>on</strong>g> the result is<br />

approximately 0.3KA which leads αD ∼ 1.1 in agreement with the value indicated in<br />

Fig. 2.15a.<br />

As the aspect ratio is raised even further, less symmetric defect c<strong>on</strong>figura-<br />

ti<strong>on</strong>s become energetically favored corresp<strong>on</strong>ding to a larger number of disclinati<strong>on</strong>s<br />

c<strong>on</strong>fined in the cup of positive Gaussian curvature. For example, when two discli-<br />

nati<strong>on</strong>s are c<strong>on</strong>fined within r = r0, the outer defect ring is given by four defects<br />

approximately located at the vertices of a square. We note that these defect c<strong>on</strong>figu-<br />

rati<strong>on</strong>s cease to exist at low aspect ratios because they require the geometric potential<br />

to overcome the str<strong>on</strong>g repulsive interacti<strong>on</strong> between the c<strong>on</strong>fined defects. As dis-<br />

cussed earlier for Fig.(2.14), the actual values of the aspect ratios involved depend<br />

<strong>on</strong> the specific geometry of the substrate. However the basic mechanism behind the<br />

dec<strong>on</strong>finement transiti<strong>on</strong> is more general.<br />

Note that, as α increases, the geometric mechanism of defect dipole unbind-<br />

ing discussed in the last secti<strong>on</strong> may also set in. Because of the presence of <strong>on</strong>e or<br />

more positive defects at the top of the bump, the critical aspect ratio necessary to<br />

unbind <strong>on</strong>e dipole will be larger than what was calculated before. If dipole unbinding<br />

does occur, the new defects will ”decorate” the existing patterns by adding new posi-<br />

tively charged disclinati<strong>on</strong>s in the regi<strong>on</strong> of positive Gaussian curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> expelling<br />

the negative <strong>on</strong>es in the external regi<strong>on</strong> of the bump (r > r0) where the Gaussian<br />

curvature is also negative (see Fig.2.1).


Chapter 2: B<strong>on</strong>d-orientati<strong>on</strong>al order <strong>on</strong> a corrugated substrate. 59<br />

2.5 C<strong>on</strong>clusi<strong>on</strong><br />

We have discussed how the varying curvature of a surface such as a ”Gaus-<br />

sian bump” can trigger the generati<strong>on</strong> of single defects or the unbinding of dipoles,<br />

even if no topological c<strong>on</strong>straints or entropic arguments require their presence. This<br />

mechanism is independent of temperature if the system is kept well below its Koster-<br />

litz Thouless transiti<strong>on</strong> temperature. It would be interesting to revisit Kosterlitz-<br />

Thouless defect unbinding transiti<strong>on</strong>s <strong>on</strong> surfaces of varying Gaussian curvature in<br />

the presence of a quenched topography [15] in the light of the present work. One<br />

might also explore the dynamics of the delocalizati<strong>on</strong> transiti<strong>on</strong> that occurs when a<br />

bump is c<strong>on</strong>fined by a circular edge <str<strong>on</strong>g>and</str<strong>on</strong>g> the aspect ratio is lowered until the defects,<br />

initially c<strong>on</strong>fined <strong>on</strong> top of the bump by the geometric potential, are forced to ”slide”<br />

towards the boundary. Quantitative studies of periodic arrangements of bumps would<br />

be interesting <str<strong>on</strong>g>and</str<strong>on</strong>g> could be inspired by fruitful analogies with methods <str<strong>on</strong>g>and</str<strong>on</strong>g> ideas from<br />

solid state physics.<br />

We also hope to extend this work by c<strong>on</strong>sidering crystalline order <strong>on</strong> bumpy<br />

topographies <str<strong>on</strong>g>and</str<strong>on</strong>g> taking explicitly into account the screening of clouds of dislocati<strong>on</strong>s<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> possible generati<strong>on</strong> of grain boundaries [4]. Such an analysis would facilitate com-<br />

paris<strong>on</strong> with experiments performed with a single grain of block copolymer spherical<br />

domains 3 <strong>on</strong> a suitably patterned substrate [30, 37].<br />

3 Block copolymers are formed by blocks of the same m<strong>on</strong>omer unit (labeled by A) covalently bound to<br />

sequences of an unlike type (labeled by B). By c<strong>on</strong>trolling the relative volume fracti<strong>on</strong> of the two blocks,<br />

<strong>on</strong>e can engineer a self-assembled ordered phase composed of spheres of block A in a sea of B m<strong>on</strong>omers.<br />

The typical radius of the resulting block-copolymer spherical cores is of the order of a few nanometers <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

their spacing tens of nanometers. These values can be tuned by suitably choosing the block-copolymers <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

varying their volume fracti<strong>on</strong>.


Chapter 3<br />

<str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick<br />

spherical shells.<br />

Based <strong>on</strong> V. Vitelli <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>, c<strong>on</strong>d-mat/0604293<br />

60


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 61<br />

3.1 Introducti<strong>on</strong><br />

The study of liquid crystal phases benefits from geometrical reas<strong>on</strong>ing in<br />

two important ways. Firstly, liquid crystal elasticity can often be cast in terms of the<br />

curvature of equipotential lines (or surfaces) that map out the corresp<strong>on</strong>ding textures.<br />

Sec<strong>on</strong>d, the observed textures are str<strong>on</strong>gly affected by geometric <str<strong>on</strong>g>and</str<strong>on</strong>g> topological<br />

c<strong>on</strong>straints imposed by the presence of boundaries c<strong>on</strong>fining the system. The liquid<br />

crystal ground state results from the competiti<strong>on</strong> between the energetic requirement<br />

of minimizing the ”curvature of the texture” <str<strong>on</strong>g>and</str<strong>on</strong>g> the geometric frustrati<strong>on</strong> introduced<br />

by boundaries that impart a preferred curvature at the edge of the sample that often<br />

cannot propagate across the system [38, 47, 51].<br />

The boundary c<strong>on</strong>diti<strong>on</strong>s can be c<strong>on</strong>trolled experimentally with the possibil-<br />

ity of designing molecular systems with intriguing technological applicati<strong>on</strong>s [52]. For<br />

example, colloidal particles coated by a very thin nematic layer have in their ground<br />

state four disclinati<strong>on</strong>s sitting at the vertices of a tetrahedr<strong>on</strong>. Each coated colloidal<br />

particle can then be viewed as the fundamental building block of a self assembled<br />

lattice with tetravalent coordinati<strong>on</strong>. The ”b<strong>on</strong>ds” between the colloidal particles<br />

could be provided by chemical linkers attached at the four ”bald spots” at the cores<br />

of the four disclinati<strong>on</strong>s present in each colloid [53]. A sec<strong>on</strong>d example, is provided by<br />

self-assembled systems of block copolymers [30] which are a promising tool for ”soft<br />

lithography” <strong>on</strong> both flat <str<strong>on</strong>g>and</str<strong>on</strong>g> curved substrates [54]. In additi<strong>on</strong>, liquid crystals in<br />

c<strong>on</strong>fined geometries provide an arena for physicists <str<strong>on</strong>g>and</str<strong>on</strong>g> mathematicians interested in<br />

applicati<strong>on</strong>s of geometrical <str<strong>on</strong>g>and</str<strong>on</strong>g> topological ideas to material science [42, 55, 56].


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 62<br />

In this work we present a theoretical study of liquid crystal phases (focusing<br />

<strong>on</strong> vector, nematic <str<strong>on</strong>g>and</str<strong>on</strong>g> hexatic order) c<strong>on</strong>fined in a spherical shell of varying thickness<br />

with the director assumed to be tangent to the two interfaces. We first c<strong>on</strong>sider the<br />

two dimensi<strong>on</strong>al regime where a nematic film coats a quenched spherical surface<br />

such as a colloidal particle in soluti<strong>on</strong> or the interface of, say, a water droplet in<br />

oil. The presence of topological defects in the ground state for ordered states <strong>on</strong><br />

spherical surfaces is unavoidable [31, 57, 12] . Recent experimental <str<strong>on</strong>g>and</str<strong>on</strong>g> theoretical<br />

investigati<strong>on</strong>s of spherical crystallography have provided an alternative c<strong>on</strong>text to<br />

study the c<strong>on</strong>straints posed by the compactness of the underlying curved space [4,<br />

3]. More recent explorati<strong>on</strong>s have c<strong>on</strong>centrated <strong>on</strong> 2D ordered phases c<strong>on</strong>fined to<br />

interfaces of varying Gaussian curvature [39, 58, 59] as well as dynamically fluctuating<br />

surfaces [35, 36].<br />

As the thickness of a nematic film increases, an escaped three dimensi<strong>on</strong>al<br />

texture, also str<strong>on</strong>gly influenced by the spherical topology <str<strong>on</strong>g>and</str<strong>on</strong>g> the boundary c<strong>on</strong>di-<br />

ti<strong>on</strong>s, become energetically favored with respect to planar textures. This instability<br />

destabilizes the tetravalent nematic texture <strong>on</strong> colloids. In this paper, we estimate<br />

the thickness of the nematic film above which a texture with four radial disclinati<strong>on</strong><br />

lines of charge s = 1<br />

2<br />

becomes unstable to four half-hedgehogs. The two competing<br />

textures studied in this paper are shown in Fig. 3.1. We also discuss the possibility<br />

of hysteresis between the two textures.<br />

The organizati<strong>on</strong> of this paper is as follows. In secti<strong>on</strong> 3.2 we derive exact<br />

soluti<strong>on</strong>s for the ground state of spherical films of tilted molecules <str<strong>on</strong>g>and</str<strong>on</strong>g> nematogens


<str<strong>on</strong>g>Liquid</str<strong>on</strong>g><br />

1mm<br />

Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 63<br />

Figure 3.1: (Top panel) Two-dimensi<strong>on</strong>al texture characterized by four short disclinati<strong>on</strong><br />

lines at the vertices of a tetrahedr<strong>on</strong> inscribed in the sphere. The surface texture<br />

shown (inscribed <strong>on</strong> the surface of a baseball) is invariant throughout the thickness<br />

of the shell. (Bottom panel) Cut view of the escaped three dimensi<strong>on</strong>al texture given<br />

by two pairs of half hedgehogs located at the north <str<strong>on</strong>g>and</str<strong>on</strong>g> south poles of the sphere.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 64<br />

within isotropic elasticity by using the method of c<strong>on</strong>formal mappings. In the nota-<br />

ti<strong>on</strong> of References [53, 12], these situati<strong>on</strong>s corresp<strong>on</strong>d to order parameters described<br />

by a b<strong>on</strong>d angle with p = 1, 2 fold symmetry in the tangent plane of the sphere (see<br />

Appendix A). A mathematical justificati<strong>on</strong> for our approach is provided in Appendix<br />

B, where the same technique is illustrated in the c<strong>on</strong>text of a more familiar flat space<br />

problem. In secti<strong>on</strong> 3.3 we study the stability of liquid crystal textures to thermal<br />

fluctuati<strong>on</strong>s by means of a normal mode analysis whose details are relegated to Ap-<br />

pendices A <str<strong>on</strong>g>and</str<strong>on</strong>g> C. The stability of the valence-four texture against escaped soluti<strong>on</strong>s<br />

is c<strong>on</strong>sidered in secti<strong>on</strong> 3.4 where a phase diagram is derived with the thickness as a<br />

c<strong>on</strong>trol parameter. The texture distorti<strong>on</strong>s caused by the elastic anisotropy between<br />

bend <str<strong>on</strong>g>and</str<strong>on</strong>g> splay deformati<strong>on</strong>s are briefly c<strong>on</strong>sidered in secti<strong>on</strong> 3.5.<br />

3.2 Textures<br />

The liquid crystal free energy for molecules embedded in an arbitrary frozen<br />

surface can be written in the <strong>on</strong>e c<strong>on</strong>stant approximati<strong>on</strong> as<br />

F = K<br />

2<br />

�<br />

dA Din j (u)D i nj(u) , (3.1)<br />

where u = {u1, u2} is a set of internal coordinates, n(u) is the liquid crystal director<br />

defined in the tangent plane, Di is the covariant derivative with respect to the metric<br />

of the surface <str<strong>on</strong>g>and</str<strong>on</strong>g> dA is the infinitesimal surface area [33, 32, 35, 40]. The free energy<br />

of Eq.(3.1) is invariant up<strong>on</strong> rotating each molecule n(u) by the same (arbitrary)<br />

angle with respect to any axis of rotati<strong>on</strong> perpendicular to the local tangent plane.<br />

The treatment of systems with a p-fold symmetry is straightforward provided that the


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 65<br />

<strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong> is used for p = 1 <str<strong>on</strong>g>and</str<strong>on</strong>g> p = 2 <str<strong>on</strong>g>and</str<strong>on</strong>g> the c<strong>on</strong>sequences<br />

of any additi<strong>on</strong>al couplings to curvature neglected [32]. This choice of free energy<br />

implies that the minimal energy c<strong>on</strong>figurati<strong>on</strong> will be given locally by neighboring<br />

n(u) vectors which differ <strong>on</strong>ly by parallel transport. The curvature of the surface<br />

induces ”frustrati<strong>on</strong>” in the texture. In fact, by Gauss’ ”Theorema egregium” [42],<br />

tangent vectors parallel transported al<strong>on</strong>g a closed loop are rotated by an amount<br />

equal to the Gaussian curvature integrated over the enclosed area. On a sphere, this<br />

theorem insures that the nematic ground state always has four excess disclinati<strong>on</strong>s<br />

[31, 12]. More generally, the sum of the topological charges <strong>on</strong> any closed surface<br />

is equal to the integrated Gaussian curvature, implying a minimum of 2 <str<strong>on</strong>g>and</str<strong>on</strong>g> 12<br />

disclinati<strong>on</strong>s in the ground state of tilted molecules <str<strong>on</strong>g>and</str<strong>on</strong>g> hexatics, respectively.<br />

We introduce a local angle field α(u), corresp<strong>on</strong>ding to the angle between<br />

n(u) <str<strong>on</strong>g>and</str<strong>on</strong>g> an arbitrary local reference frame, we can rewrite the free energy introduced<br />

in Eq.(3.1) as:<br />

F = 1<br />

2 K<br />

�<br />

dS g ij (∂iα − Ai)(∂jα − Aj) , (3.2)<br />

where dS = d 2 u √ g, g is the determinant of the metric tensor gij <str<strong>on</strong>g>and</str<strong>on</strong>g> Ai is the spin-<br />

c<strong>on</strong>necti<strong>on</strong> whose curl is the Gaussian curvature G(u) [40, 42]. On a sphere of radius<br />

R parametrized by polar coordinates (θ, φ), the <strong>on</strong>ly n<strong>on</strong> vanishing comp<strong>on</strong>ents of the<br />

(inverse) metric tensor are g rr = 1<br />

R sin θ <str<strong>on</strong>g>and</str<strong>on</strong>g> gφφ = 1<br />

R<br />

. A c<strong>on</strong>venient choice of of the<br />

spin c<strong>on</strong>necti<strong>on</strong> (which plays the role of the vector potential) is discussed in Appendix<br />

A. The simplified free energy in Eq.(4.3) is the starting point of our analysis.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 66<br />

3.2.1 Tilted molecules <strong>on</strong> a sphere<br />

The orientati<strong>on</strong>al order of molecules tilted by a c<strong>on</strong>stant angle with respect<br />

to a spherical interface can be modelled by a vector field n(θ, φ) defined in the local<br />

tangent plane <strong>on</strong> which the molecule has a fixed length projecti<strong>on</strong> [57]. To determine<br />

the ground state of the liquid crystal texture, we minimize the Frank free energy of<br />

Eq.(4.3). As discussed above, the topological charges must sum up to 4π, the inte-<br />

grated Gaussian curvature of the sphere [40, 42]. For a vector field (p = 1) the texture<br />

with <strong>on</strong>ly two defects of charges +2π minimizes the Frank free energy <str<strong>on</strong>g>and</str<strong>on</strong>g> satisfies<br />

the topological c<strong>on</strong>straint. Since the defects repel each other they preferentially sit<br />

at two antipodal points that we can designate as the north <str<strong>on</strong>g>and</str<strong>on</strong>g> south pole of the<br />

sphere. If the splay <str<strong>on</strong>g>and</str<strong>on</strong>g> bend coupling c<strong>on</strong>stants of the nematic are equal, then there<br />

is a large degeneracy in the ground state arising from the invariance of the vector<br />

free energy in Eq.(4.3) under global rotati<strong>on</strong>s α(u)→α(u)+c, where u ≡ θ, φ. One<br />

representative texture is a ”sink” <str<strong>on</strong>g>and</str<strong>on</strong>g> a ”source” of n(u) at the two poles. In this<br />

splay rich texture n(u) is parallel to the lines of l<strong>on</strong>gitude <strong>on</strong> a sphere. In a bend rich<br />

texture, related to the previous by a π<br />

2<br />

rotati<strong>on</strong> about the local normal to the sur-<br />

face, n(u) is everywhere parallel to the lines of latitude. Any other rotati<strong>on</strong> of n(u)<br />

that makes an arbitrary c<strong>on</strong>stant angle with respect to this texture is an acceptable<br />

soluti<strong>on</strong> for the ground state of the molecules.<br />

As we now show, this degeneracy is lifted when K3 �= K1. Indeed the effect<br />

of distinct splay <str<strong>on</strong>g>and</str<strong>on</strong>g> bend elastic c<strong>on</strong>stants K1 <str<strong>on</strong>g>and</str<strong>on</strong>g> K3 (the twist elastic c<strong>on</strong>stant<br />

K2 is absent in two dimensi<strong>on</strong>s) is to select the bend-rich texture if K1 > K3 or


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 67<br />

the splay-rich <strong>on</strong>e if K3 > K1. The intermediate c<strong>on</strong>figurati<strong>on</strong>s obtained by a global<br />

rotati<strong>on</strong> of the director are now unstable. Assume for simplicity that K3 > K1. In<br />

this case, it is c<strong>on</strong>venient to recast the Frank free energy (see Appendix A) as follows:<br />

F = 1<br />

�<br />

2<br />

d 2 x √ �<br />

g [K1 Di n j ) ( D i n j�<br />

+(K3 − K1)(D × n) 2 ] , (3.3)<br />

where the covariant derivatives is expressed in terms of the Christoffel c<strong>on</strong>necti<strong>on</strong>,<br />

Γ j<br />

i t ,<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the covariant form of the curl squared is [40, 60],<br />

Di n j = ∂i n j + Γ j<br />

i t nt , (3.4)<br />

( � D × n) 2 ≡ � Di nj − Dj ni ) ( D i n j − D j n i� . (3.5)<br />

The first term in Eq.(3.3) resembles the Frank free energy in the <strong>on</strong>e coupling c<strong>on</strong>stant<br />

approximati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> is minimized by choosing the sink-source (or ”lines of l<strong>on</strong>gitude”)<br />

soluti<strong>on</strong>. The sec<strong>on</strong>d term (which is positive definite) will vanish for this texture since<br />

the sink-source texture is bend free. All other textures have a higher energy.<br />

A similar argument can be used to prove that the two vortex-c<strong>on</strong>figurati<strong>on</strong><br />

which follows the lines of latitude is the minimum of the free energy when K1 > K3<br />

by rewriting the Frank free energy as<br />

F = 1<br />

�<br />

2<br />

d 2 x √ �<br />

g [K3 Di n j ) ( D i �<br />

nj<br />

+ (K1 − K3) (D · n) 2 ] , (3.6)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 68<br />

where the covariant form of the divergence reads<br />

D · n ≡ 1 �√ i<br />

√ ∂i g n<br />

g � . (3.7)<br />

The latitudinal texture minimizes the first term of Eq.(3.6) while the sec<strong>on</strong>d vanishes<br />

because this texture is splay free. Any deviati<strong>on</strong> from the splay-free latitudinal texture<br />

will <strong>on</strong>ly increase the energy.<br />

The energy of both textures can be expressed as a functi<strong>on</strong> of the anisotropy<br />

parameter, ɛ, <str<strong>on</strong>g>and</str<strong>on</strong>g> the mean of the elastic c<strong>on</strong>stants, K,<br />

ɛ ≡ K3 − K1<br />

K3 + K1<br />

K ≡ K3 + K1<br />

2<br />

, (3.8)<br />

, (3.9)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the radius of the sphere, R, scaled by the short distance cutoff a. The resulting<br />

free energy for arbitrary ɛ reads (see Eq.(E.16))<br />

� � � �<br />

R<br />

F = 2πK (1 − |ɛ |) ln − 0.3<br />

a<br />

. (3.10)<br />

The c<strong>on</strong>clusi<strong>on</strong>s of this secti<strong>on</strong> are summarized in Fig. 3.2 which suggests that there<br />

is a disc<strong>on</strong>tinuous first order transiti<strong>on</strong> when ɛ passes through zero. This analysis<br />

mirrors similar arguments valid in the plane [61].<br />

3.2.2 Nematic Texture<br />

The nematic texture of very thin spherical shells of nematic liquid crystal<br />

with tangential boundary c<strong>on</strong>diti<strong>on</strong>s can be analyzed within the <strong>on</strong>e Frank c<strong>on</strong>stant<br />

approximati<strong>on</strong> by using the method of c<strong>on</strong>formal mappings whose mathematical jus-<br />

tificati<strong>on</strong> is illustrated in Appendix B by means of a simpler example.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 69<br />

-1 1<br />

Figure 3.2: Schematic illustrati<strong>on</strong> of the phase diagram of the texture of tilted<br />

molecules <strong>on</strong> a sphere as a functi<strong>on</strong> of the anisotropy parameter ɛ superimposed<br />

<strong>on</strong> a plot of the free energy, F , stored in the texture versus ɛ. The two competing<br />

ground states (top view) are characterized by either pure bend (lines of l<strong>on</strong>gitude<br />

c<strong>on</strong>figurati<strong>on</strong> at left) or pure splay (lines of latitude c<strong>on</strong>figurati<strong>on</strong> at right).<br />

F<br />

ε


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 70<br />

An elegant argument introduced by Lubensky <str<strong>on</strong>g>and</str<strong>on</strong>g> Prost in Ref.[12] shows<br />

that the ground state of nematogens <strong>on</strong> a sphere is given by 4 disclinati<strong>on</strong>s of topo-<br />

logical charge s = 1/2 sitting at the vertexes of a tetrahedr<strong>on</strong>. The energy of single<br />

disclinati<strong>on</strong>s is proporti<strong>on</strong>al to the square of its strength. As a result, the l<strong>on</strong>gitudi-<br />

nal <str<strong>on</strong>g>and</str<strong>on</strong>g> latitudinal textures derived for tilted molecules in secti<strong>on</strong> 3.2.1 are unstable<br />

since their energies can be lowered by splitting each s = 1 defect at the north <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

south pole into two s = 1/2 disclinati<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> letting them relax to their equilibrium<br />

positi<strong>on</strong>s at the vertexes of a tetrahedr<strong>on</strong> where they are as far away from each other<br />

as possible. According to a calculati<strong>on</strong> in Ref.[12], the energy Fs of a sphere of radius<br />

R with in plane orientati<strong>on</strong>al order <str<strong>on</strong>g>and</str<strong>on</strong>g> 2p interacting minimal disclinati<strong>on</strong>s for a<br />

p-fold order parameter is given by<br />

�<br />

1<br />

Fs = 2πK h<br />

p ln<br />

� � �<br />

2 4p R<br />

+ cp . (3.11)<br />

a<br />

where the {cp} are c<strong>on</strong>stants depending <strong>on</strong> the symmetry of the order parameter <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

the defect core energy while h is the thickness of the liquid crystal layer 1 . When<br />

p = 2 is chosen in Eq.(3.11) the elastic energy is indeed smaller than the corresp<strong>on</strong>ding<br />

value for p = 1 in the limit R ≫ a, in agreement with related arguments given in<br />

Ref.[53].<br />

To obtain an algebraic expressi<strong>on</strong> for the texture we proceed as illustrated in<br />

Appendix B <str<strong>on</strong>g>and</str<strong>on</strong>g> seek a functi<strong>on</strong> Ω(x, y, z) = Φ(x, y, z)+iΨ(x, y, z) which is harm<strong>on</strong>ic<br />

<strong>on</strong> the sphere except for two arcs c<strong>on</strong>necting the defects in pairs. The calculati<strong>on</strong><br />

for nematogens described below was suggested to us by F. Dys<strong>on</strong> [62]. The functi<strong>on</strong><br />

1 The numerical values of the relevant c<strong>on</strong>stants are c1 = 0 <str<strong>on</strong>g>and</str<strong>on</strong>g> c2 � −0.2.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 71<br />

Figure 3.3: Schematic illustrati<strong>on</strong> of the baseball texture of a thin nematic shell. The<br />

same texture is reproduced from a different perspective in the top panel of Fig. 3.1.<br />

Φ(x, y, z), which we can interpret as an electrostatic potential, takes equal <str<strong>on</strong>g>and</str<strong>on</strong>g> op-<br />

posite values <strong>on</strong> the two arcs <str<strong>on</strong>g>and</str<strong>on</strong>g> is equal to zero <strong>on</strong> a baseball-like seam (see Fig.<br />

3.3) which divides the sphere into two c<strong>on</strong>gruent regi<strong>on</strong>s. The nematic director is<br />

then oriented (up to a global rotati<strong>on</strong>) al<strong>on</strong>g the c<strong>on</strong>tour lines of Φ(x, y, z), that is<br />

the equipotential lines of this ”curved space capacitor”. In this analogy, the c<strong>on</strong>tour<br />

lines of Ψ(x, y, z) are electric field lines, hence they corresp<strong>on</strong>d to a valid texture<br />

where the director is rotated locally by π<br />

2<br />

with respect to the equipotential lines. The<br />

arcs can be either great-circle arcs extending more than half way round the sphere or<br />

short great-circles arcs c<strong>on</strong>necting the same pair of defects al<strong>on</strong>g the shortest path.<br />

The first choice leads to equipotential lines whose seam resembles in shape that of a


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 72<br />

C<strong>on</strong>formal plane<br />

N<br />

S<br />

Figure 3.4: Graphical c<strong>on</strong>structi<strong>on</strong> of the stereographic projecti<strong>on</strong>. Regi<strong>on</strong>s close to<br />

the north pole have larger images in the c<strong>on</strong>formal plane than regi<strong>on</strong>s of equal areas<br />

close to the south pole. The stereographic projecti<strong>on</strong> preserves the topology of the<br />

surface provided all points at infinity are identified with the north pole.<br />

baseball. If the sec<strong>on</strong>d choice is made the pattern of equipotential lines would not<br />

deviate much from c<strong>on</strong>centric circles <str<strong>on</strong>g>and</str<strong>on</strong>g> the seam would look more like the seam of<br />

a cricket ball. We will explicitly show that the two choices are equivalent since the<br />

equipotential lines of the first soluti<strong>on</strong> are field lines of the sec<strong>on</strong>d <str<strong>on</strong>g>and</str<strong>on</strong>g> vice versa.<br />

We choose the arcs c<strong>on</strong>necting the defect pairs al<strong>on</strong>g great circles <str<strong>on</strong>g>and</str<strong>on</strong>g> we<br />

take the four defects labelled by A, B, C, D to lie at the vertices of a tetrahedr<strong>on</strong><br />

inscribed <strong>on</strong> a sphere of radius 1 <str<strong>on</strong>g>and</str<strong>on</strong>g> whose north <str<strong>on</strong>g>and</str<strong>on</strong>g> south poles are N = (0, 0, 1)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> S = (0, 0, −1) respectively<br />

A = 1<br />

√ 3 (1, 1, 1) , B = 1<br />

√ 3 (−1, −1, 1) ,<br />

C = 1<br />

√ 3 (−1, 1, −1) , D = 1<br />

√ 3 (1, −1, −1) . (3.12)<br />

We now perform a stereographic projecti<strong>on</strong> (see Fig. A.1) that maps every point <strong>on</strong>


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 73<br />

a unit sphere centered <strong>on</strong> the origin <strong>on</strong>to the plane z = −1 according to the rule<br />

⎛ ⎞ ⎛ ⎞<br />

⎜ x ⎟ ⎜<br />

⎜ ⎟ ⎜<br />

⎜ ⎟ ⎜<br />

⎜<br />

y ⎟ → ⎜<br />

⎟ ⎜<br />

⎝ ⎠ ⎝<br />

a<br />

b<br />

⎟<br />

⎠<br />

z −1<br />

. (3.13)<br />

The coordinates of the image points (c<strong>on</strong>nected to points <strong>on</strong> the sphere by dashed<br />

lines in Fig. A.1) are given by<br />

a = 2x<br />

1 − z ,<br />

b =<br />

2y<br />

1 − z<br />

. (3.14)<br />

Up<strong>on</strong> transforming to a complex coordinate w = a + ib, the four tetrahedral points<br />

of Eq.(3.12) are mapped <strong>on</strong>to<br />

where<br />

A ′ = p (1 + i) , B ′ = p (−1 − i) ,<br />

C ′ = q (−1 + i) , D ′ = q (1 − i) , (3.15)<br />

p = √ 3 + 1 , q = √ 3 − 1 . (3.16)<br />

(In this secti<strong>on</strong> p does not refer to the symmetry of the order parameter). The<br />

great arc passing through the south pole (corresp<strong>on</strong>ding to <strong>on</strong>e capacitor plate in the<br />

electrostatic analogy) maps <strong>on</strong>to the segment A ′ B ′ , as illustrated schematically in the<br />

top panel of Fig. 3.5, while the great arc through the north pole maps <strong>on</strong>to the two<br />

semi-infinite segments of the line a + b = 0 which bracket C ′ D ′ . We can fold the two<br />

cuts in the w plane back <strong>on</strong> top of each other by mapping the w plane <strong>on</strong>to the u


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 74<br />

-1<br />

B'<br />

C'<br />

C”<br />

w<br />

u<br />

D'<br />

A'<br />

-2q 2 2p 2<br />

D”<br />

v<br />

A”<br />

B”<br />

-k k 1<br />

Figure 3.5: Illustrati<strong>on</strong> of the change in the branch cut structures of the complex<br />

functi<strong>on</strong> Ω(v) describing the nematic texture after performing a series of c<strong>on</strong>formal<br />

transformati<strong>on</strong>s. In the top panel we have simply performed a stereographic projecti<strong>on</strong><br />

from the sphere. The middle panel shows the ”fold up” transformati<strong>on</strong> of<br />

Eq.(3.17) whereas the bottom panel corresp<strong>on</strong>ds to the transformati<strong>on</strong> in Eq.(3.21)<br />

that symmetrizes the positi<strong>on</strong>s of the cuts.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 75<br />

plane via<br />

u = −iw 2 . (3.17)<br />

As shown in the middle panel of Fig. 3.5, the images à <str<strong>on</strong>g>and</str<strong>on</strong>g> ˜ B of A’ <str<strong>on</strong>g>and</str<strong>on</strong>g> B’ now both<br />

lie <strong>on</strong> the real axis at 2p 2 while the images of C’ <str<strong>on</strong>g>and</str<strong>on</strong>g> D’ now lie at −2q 2 . The two cuts<br />

in the u plane are both <strong>on</strong> the real axis, running from zero to 2p 2 <str<strong>on</strong>g>and</str<strong>on</strong>g> from −2q 2 to<br />

minus infinity. On the sphere, these corresp<strong>on</strong>d to geodesics c<strong>on</strong>necting defects which<br />

stretch more than halfway around the sphere. In order to make the cuts symmetric<br />

with respect to the imaginary axis (see the bottom panel of Fig. 3.5) we search for a<br />

c<strong>on</strong>formal transformati<strong>on</strong> that maps the following four points in the complex u-plane<br />

to four points <strong>on</strong> the real axis of a complex v-plane<br />

u0 = 0 → v0 = k ,<br />

u1 = −2q 2 → v1 = −k ,<br />

u2 = −∞ → v2 = −1 ,<br />

u3 = 2p 2 → v3 = 1 . (3.18)<br />

In order to fully determine the c<strong>on</strong>formal transformati<strong>on</strong> we need to determine the<br />

value of k. This can be d<strong>on</strong>e by using a st<str<strong>on</strong>g>and</str<strong>on</strong>g>ard relati<strong>on</strong> in the theory of c<strong>on</strong>formal<br />

transformati<strong>on</strong>s [45]<br />

u0 − u1<br />

u0 − u2<br />

u3 − u2<br />

u3 − u1<br />

= v0 − v1<br />

v0 − v2<br />

v3 − v2<br />

v3 − v1<br />

. (3.19)<br />

Up<strong>on</strong> inserting the points of Eq.(3.18) into Eq.(3.19), we determine the value of k<br />

(less than <strong>on</strong>e)<br />

k = 2√ 2 − p<br />

p + 2 √ 2<br />

. (3.20)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 76<br />

Equati<strong>on</strong> (3.18) c<strong>on</strong>tains four independent relati<strong>on</strong>s so we are still left with three<br />

c<strong>on</strong>diti<strong>on</strong>s to determine the three independent coefficients {α, β, δ} of the bilinear<br />

c<strong>on</strong>formal transformati<strong>on</strong> that implements the mapping illustrated pictorially in the<br />

bottom plate of Fig. 3.5<br />

v =<br />

u + δ<br />

α u + β<br />

. (3.21)<br />

The required coefficients needed to implement the mapping in Eq.(3.18) are<br />

α = −1 ,<br />

β = 2p (2 √ 2 + p) ,<br />

δ = 2p (2 √ 2 − p) , (3.22)<br />

To solve Laplace’s equati<strong>on</strong>, we desire a functi<strong>on</strong> Ω(v) which is analytic except <strong>on</strong><br />

the two cuts <strong>on</strong> the real axis, <str<strong>on</strong>g>and</str<strong>on</strong>g> whose real part takes c<strong>on</strong>stant values <strong>on</strong> the cuts.<br />

By symmetry, Ω(v) is an odd functi<strong>on</strong> of v, <str<strong>on</strong>g>and</str<strong>on</strong>g> its real part Φ(v) is zero when the<br />

real part of v is zero. Therefore the image of the seam in the v plane is simply the<br />

imaginary axis ℜe v = 0. Up<strong>on</strong> substituting for v using Equati<strong>on</strong>s (3.21) <str<strong>on</strong>g>and</str<strong>on</strong>g> (3.17),<br />

the c<strong>on</strong>diti<strong>on</strong> ℜe v = 0 becomes<br />

16 + 4p 2 ℑm(w 2 ) − |w| 4 = 0 , (3.23)<br />

With the help of Equati<strong>on</strong>s (3.14) <str<strong>on</strong>g>and</str<strong>on</strong>g> (3.13), we can now write down the equati<strong>on</strong><br />

of the seam explicitly in the original cartesian coordinates [62]<br />

z = (2 + √ 3) xy , (3.24)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 77<br />

a b<br />

c<br />

Figure 3.6: Different views of a track of parallel nematogens which partiti<strong>on</strong>s the<br />

sphere into two equal areas, each c<strong>on</strong>taining two s = 1 disclinati<strong>on</strong> defects. It resem-<br />

2<br />

bles a ”fattened” versi<strong>on</strong> of the seam of a baseball.<br />

or in spherical polar coordinates {φ, θ} as<br />

cos θ<br />

sin2 θ =<br />

�<br />

1 +<br />

d<br />

√ �<br />

3<br />

sin 2φ . (3.25)<br />

2<br />

The seam defined by the line of zero potential, is represented for different orientati<strong>on</strong>s<br />

of the sphere in Fig. 3.6. Its c<strong>on</strong>tour length l, <strong>on</strong> a unit radius ball, is readily<br />

calculated up<strong>on</strong> integrating the expressi<strong>on</strong> for the infinitesimal arc of the seam<br />

�<br />

dl = sin2 � �2 dφ<br />

θ + 1 dθ , (3.26)<br />

dθ<br />

from θmin ≈ 0.69 radians to θmax ≈ 2.44 radians <str<strong>on</strong>g>and</str<strong>on</strong>g> multiplying the result by four<br />

in view of the symmetry of the seam. The values of θmin <str<strong>on</strong>g>and</str<strong>on</strong>g> θmax are obtained<br />

from Eq.(3.25) by setting φ equal to π<br />

4<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> 3π<br />

4<br />

respectively. Up<strong>on</strong> using Eq.(3.25) to<br />

substitute φ(θ) in Eq.(3.26), we obtain l � 9.09 for a sphere of unit radius. The seam


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 78<br />

is l<strong>on</strong>ger than the equatorial circumference by slightly less than 50%.<br />

The branch cut structure in the v plane is sufficiently simple to allow a guess<br />

of the corresp<strong>on</strong>ding analytic functi<strong>on</strong> Ω(v). A functi<strong>on</strong> with cuts from k to 1 <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

−k to −1, whose real part is equal <str<strong>on</strong>g>and</str<strong>on</strong>g> opposite <strong>on</strong> the two cuts <str<strong>on</strong>g>and</str<strong>on</strong>g> with a single<br />

imaginary period around any curve separating the cuts is easily identified to be a<br />

st<str<strong>on</strong>g>and</str<strong>on</strong>g>ard elliptic integral<br />

Ω(v) =<br />

� v<br />

0<br />

� (k 2 − t 2 )(1 − t 2 ) � − 1<br />

2 dt . (3.27)<br />

with v given in terms of w by Equati<strong>on</strong>s (3.17) <str<strong>on</strong>g>and</str<strong>on</strong>g> (3.21). The nematic director is<br />

oriented (up to a global rotati<strong>on</strong>) al<strong>on</strong>g the c<strong>on</strong>tour lines of the imaginary or (real<br />

part) of Ω(v). The equipotential (red) <str<strong>on</strong>g>and</str<strong>on</strong>g> field lines (black) of Ω(v) are c<strong>on</strong>veniently<br />

plotted using the stereographic projecti<strong>on</strong> plane w = a + ib in Fig. 3.7 al<strong>on</strong>g with the<br />

positi<strong>on</strong>s of the disclinati<strong>on</strong>s (green dots). It is easy to switch from the stereographic-<br />

projecti<strong>on</strong> plane w = a + ib of Fig. 3.7 to spherical polar coordinates (θ, φ) by using<br />

the relati<strong>on</strong> (reviewed in Appendix B),<br />

w = 2R cot<br />

where R is the radius of the sphere.<br />

� �<br />

θ<br />

e<br />

2<br />

iφ , (3.28)<br />

If we had c<strong>on</strong>structed the base-ball with cuts al<strong>on</strong>g the short geodesics<br />

c<strong>on</strong>necting the defects, then the form of the texture given in Eq.(3.27) would be the<br />

same, but the parameter k in the elliptic integral would be given by<br />

k = ( √ 3 − √ 2) 2 ( √ 2 + 1) 2 , (3.29)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 79<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-4 -2 0 2 4<br />

Figure 3.7: Illustrati<strong>on</strong> of the nematic texture in stereographic projecti<strong>on</strong> (Color<br />

<strong>on</strong>line). The red <str<strong>on</strong>g>and</str<strong>on</strong>g> black field lines corresp<strong>on</strong>d to two energetically degenerate (in<br />

the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>) families of bend <str<strong>on</strong>g>and</str<strong>on</strong>g> splay rich textures. The<br />

dots indicate the four tetrahedral s = 1 disclinati<strong>on</strong> defects.<br />

2


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 80<br />

instead of Eq.(5.13). The equati<strong>on</strong> for the seam becomes<br />

z = −(2 − √ 3) xy (3.30)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the corresp<strong>on</strong>ding equipotential <str<strong>on</strong>g>and</str<strong>on</strong>g> field lines are plotted in black <str<strong>on</strong>g>and</str<strong>on</strong>g> red re-<br />

spectively in Fig. 3.7. The expressi<strong>on</strong> reads<br />

cos θ<br />

sin2 � √ �<br />

3<br />

= − 1 − sin 2φ , (3.31)<br />

θ 2<br />

in spherical polar coordinates, <str<strong>on</strong>g>and</str<strong>on</strong>g> leads to the same c<strong>on</strong>tour length l as Eq.(3.25).<br />

Thus the two choices of arcs lead to two equivalent textures differing <strong>on</strong>ly by a π<br />

2<br />

about the local normal. As in Secti<strong>on</strong> 3.2.1, the degeneracy in energy between the<br />

red <str<strong>on</strong>g>and</str<strong>on</strong>g> black flow lines in Fig. 3.7 is lifted up<strong>on</strong> c<strong>on</strong>sidering the effect of elastic<br />

anisotropy generated by a different energy cost for bend <str<strong>on</strong>g>and</str<strong>on</strong>g> splay.<br />

3.3 Stability of liquid crystal textures to thermal fluctuati<strong>on</strong>s<br />

In this secti<strong>on</strong>, we study the stability of liquid crystal ground states to<br />

thermal fluctuati<strong>on</strong>s [53]. To explore the fidelity of directi<strong>on</strong>al b<strong>on</strong>ds at finite tem-<br />

peratures, we employ a Coulomb gas representati<strong>on</strong> of the liquid crystal free energy<br />

(in the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>) obtained by substituting in Eq.(4.3) the<br />

relati<strong>on</strong><br />

γ αβ ∂α(∂βθ − Aβ) = s(u) − G(u) ≡ n(u) , (3.32)<br />

where γ αβ is the covariant antisymmetric tensor [40], G(u) is the Gaussian curvature<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> s(u) ≡ 1 �Nd √<br />

g i=1 qiδ(u − ui) is the disclinati<strong>on</strong> density with Nd defects of charge


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 81<br />

qi at positi<strong>on</strong>s ui. The final result is an effective free energy whose basic degrees of<br />

freedom are the defects themselves [35, ?, 33]:<br />

F = K<br />

2<br />

�<br />

�<br />

dA<br />

dA ′ n(u) Γ(u, u ′ ) n(u ′ ) . (3.33)<br />

The Green’s functi<strong>on</strong> Γ(u, u ′ ) is calculated (see Appendix A) by inverting the Lapla-<br />

cian defined <strong>on</strong> the sphere<br />

Γ(u, u ′ � �<br />

1<br />

) ≡ −<br />

∆ uu ′<br />

, (3.34)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> we have suppressed for now defect core energy c<strong>on</strong>tributi<strong>on</strong>s which reflect the<br />

physics at microscopic length scales. Equati<strong>on</strong>s (3.32) <str<strong>on</strong>g>and</str<strong>on</strong>g> (3.33) can be understood<br />

by analogy to two dimensi<strong>on</strong>al electrostatics, with the Gaussian curvature G(u) (with<br />

sign reversed) playing the role of a uniform background charge distributi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> the<br />

topological defects appearing as point-like sources with electrostatic charges equal<br />

to their topological charge qi 2 . On a generic surface, the defects tend to positi<strong>on</strong><br />

themselves so that the Gaussian curvature is screened: typically, the positive <strong>on</strong>es are<br />

attracted to peaks <str<strong>on</strong>g>and</str<strong>on</strong>g> valleys while the negative <strong>on</strong>es to the saddles of the surface<br />

[58, 59]. This geometric potential is ruled out by symmetry <strong>on</strong> an undeformed sphere<br />

since the Gaussian curvature is c<strong>on</strong>stant. The Gaussian curvature plays the role of<br />

a uniform background charge fixing the net charge of the defects c<strong>on</strong>sistent with the<br />

topological c<strong>on</strong>straint imposed by the Poincare-Hopf theorem (see secti<strong>on</strong> 3.2 <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

References [42, 63]). The equilibrium positi<strong>on</strong>s of the defects are then determined<br />

<strong>on</strong>ly by defect-defect interacti<strong>on</strong>s which are proporti<strong>on</strong>al to the logarithm of their<br />

2 The charge q can be defined by the amount θ increases al<strong>on</strong>g a counterclockwise path enclosing the<br />

defect’s core.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 82<br />

chordal distance (see Appendix A) according to<br />

F = − πK<br />

2p 2<br />

�<br />

ninj ln [1 − cos βij] , (3.35)<br />

i�=j<br />

where the integers ni <str<strong>on</strong>g>and</str<strong>on</strong>g> nj describing the singularities associated with each defect,<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the integer p c<strong>on</strong>trols the period 2π<br />

p<br />

of the orientati<strong>on</strong>al order parameter. The<br />

”topological charge” describing the rotati<strong>on</strong> of the order parameter around each defect<br />

is given by sj = n<br />

p . The geodesic angle βij subtended by the two defects at positi<strong>on</strong>s<br />

ui = {θi, φi} <str<strong>on</strong>g>and</str<strong>on</strong>g> uj = {θj, φj} can be c<strong>on</strong>veniently recast in terms of their spherical<br />

polar coordinates<br />

cos βij = cos θi cos θj + sin θi sin θj cos [φi − φj] . (3.36)<br />

As a simple example, we first c<strong>on</strong>sider the case of a Z = 2 colloidal particle<br />

(p=1) with two antipodal defects of index 1. We can study the effect of thermal<br />

disrupti<strong>on</strong> of the ground state by setting βij = π + θ in Eq.(3.35) <str<strong>on</strong>g>and</str<strong>on</strong>g> exp<str<strong>on</strong>g>and</str<strong>on</strong>g>ing in<br />

the bending angle θ. The resulting free energy, apart from an additive c<strong>on</strong>stant, reads<br />

F ≈ πK<br />

4 θ2 . (3.37)<br />

Up<strong>on</strong> applying the equipartiti<strong>on</strong> theorem we obtain in the limit K ≫ kBT [53]<br />

〈cos βij〉 ≈ −1 + 1<br />

2 〈θ2 〉 ,<br />

≈ −1 + kBT<br />

πK<br />

, (3.38)<br />

which describes the fidelity of π antipodal ”b<strong>on</strong>ds” of a divalent colloidal particle.<br />

The effect of thermal fluctuati<strong>on</strong>s <strong>on</strong> the tetrahedral ground state of nematic<br />

molecules c<strong>on</strong>fined <strong>on</strong> the sphere (p=2 <str<strong>on</strong>g>and</str<strong>on</strong>g> all sj = 1)<br />

can be studied by means of a<br />

2


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 83<br />

normal mode analysis. The basic results sketched in [53] were obtained by a slightly<br />

different method. Here we describe an alternative treatment in some detail <str<strong>on</strong>g>and</str<strong>on</strong>g> extend<br />

our analysis to hexatic <str<strong>on</strong>g>and</str<strong>on</strong>g> tetratic defect arrays (see Appendix C).<br />

We start by defining a generalized array of defect coordinates {qi} as a 2N<br />

dimensi<strong>on</strong>al vector, where N is the number of defects in the ground state (or, equiv-<br />

alently, the valence of the colloidal molecule). N = 4 in the case of the tetrahedr<strong>on</strong>.<br />

The first N entries of the vector q are the l<strong>on</strong>gitudinal deviati<strong>on</strong>s of the N defects<br />

from a perfect tetrahedral c<strong>on</strong>figurati<strong>on</strong> while the remaining N comp<strong>on</strong>ents describe<br />

defect displacements al<strong>on</strong>g the lines of latitude of a sphere of unit radius. As a re-<br />

sult, the deviati<strong>on</strong>s of the i th defect from its equilibrium c<strong>on</strong>figurati<strong>on</strong> {θi 0 , φi 0 } are<br />

parameterized by the two independent comp<strong>on</strong>ents of the vector q<br />

qi = δθi ,<br />

qN+i = δφi sin(θi 0 ) . (3.39)<br />

The relati<strong>on</strong>s in Eq.(3.39) can be used to reexpress Eq.(3.36) in terms of the compo-<br />

nents of the displacements vector qi, with the result,<br />

sin � θi 0 � �<br />

+ qi sin θj 0 �<br />

+ qj cos<br />

cos βij = cos � θi 0 � �<br />

+ qi cos θj 0 �<br />

+ qj +<br />

�<br />

φi 0 − φj 0 + qN+i qN+j<br />

−<br />

sin θi<br />

0 sin θj 0<br />

�<br />

. (3.40)<br />

Up<strong>on</strong> substituting Eq.(3.40) in Eq.(3.35), the free energy F can be exp<str<strong>on</strong>g>and</str<strong>on</strong>g>ed around<br />

the equilibrium c<strong>on</strong>figurati<strong>on</strong> to quadratic order in qi with the result (apart from an<br />

additive c<strong>on</strong>stant)<br />

F ≈ 1 �<br />

2<br />

ij<br />

Mij qi qj , (3.41)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 84<br />

where the matrix, Mij, describing the deformati<strong>on</strong> of the tetrahedral molecule is<br />

naturally defined as<br />

� � 2 ∂ F<br />

Mij =<br />

∂qi ∂qj<br />

qi,qj=0<br />

. (3.42)<br />

The eigenvalues of this matrix can be classified according to the irreducible represen-<br />

tati<strong>on</strong> of the symmetry group of the tetrahedr<strong>on</strong>; their degeneracies can be determined<br />

purely from the group theoretical relati<strong>on</strong> [64, 65]<br />

n (γ) = 1 �<br />

g<br />

i<br />

gi χ (γ)∗<br />

i χ (Σ)<br />

i , (3.43)<br />

where n (γ) is the number of frequency degenerate normal modes that transform like<br />

the irreducible representati<strong>on</strong> labeled by γ, gi is the number of symmetry operati<strong>on</strong>s<br />

of the tetrahedral point group in the i th class, g = �<br />

i gi = 24 is the total number of<br />

symmetry operati<strong>on</strong> in the group, χ (γ)<br />

i is the character of the i th class in the irreducible<br />

representati<strong>on</strong> labelled by γ while χ (Σ)<br />

i<br />

representati<strong>on</strong> formed by the defects’ displacements.<br />

is the corresp<strong>on</strong>ding character for the reducible<br />

The informati<strong>on</strong> necessary to apply Eq.(3.43) to a tetravalent colloid is col-<br />

lected in Table 3.1. The top row c<strong>on</strong>tains the five symmetry class {E, C3, C2, S4, σd}<br />

c<strong>on</strong>tained in the tetrahedral point group ℑd, corresp<strong>on</strong>ding respectively to the iden-<br />

tity, three <str<strong>on</strong>g>and</str<strong>on</strong>g> two fold rotati<strong>on</strong>s, four fold rotatory-reflecti<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> reflecti<strong>on</strong> through<br />

a plane of symmetry [64]. The number of symmetry operati<strong>on</strong>s gi included in the i th<br />

class also appears in the top row: thus, {gi} = {1, 8, 3, 6, 6} where the same ordering<br />

used above to list the classes has been adopted.<br />

The left most column of Table 3.1 lists the <strong>on</strong>e, two <str<strong>on</strong>g>and</str<strong>on</strong>g> three dimensi<strong>on</strong>al<br />

irreducible representati<strong>on</strong>s of the tetrahedral group {A1, A2, E, F1, F2}, al<strong>on</strong>g with


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 85<br />

Table 3.1: Character for the irreducible representati<strong>on</strong>s of the tetrahedral point group<br />

together with the character of the eight-dimensi<strong>on</strong>al representati<strong>on</strong> Σ generated by<br />

the defect displacements of a tetravalent colloid.<br />

ℑd E 8C3 3C2 6S4 6σd<br />

A1 1 1 1 1 1<br />

A2 1 1 1 −1 −1<br />

E 2 −1 2 0 0<br />

F1 3 0 −1 1 −1<br />

F2 3 0 −1 −1 1<br />

Σ 8 −1 0 0 0<br />

the eight dimensi<strong>on</strong>al representati<strong>on</strong> Σ generated by the defect displacements. The<br />

entries of the table list the characters corresp<strong>on</strong>ding to each class of the five irreducible<br />

representati<strong>on</strong>s, χ (γ)<br />

i , <str<strong>on</strong>g>and</str<strong>on</strong>g> in the last row the corresp<strong>on</strong>ding characters, χ (Σ)<br />

i , for the<br />

eight dimensi<strong>on</strong>al representati<strong>on</strong>. The former are tabulated from st<str<strong>on</strong>g>and</str<strong>on</strong>g>ard group<br />

theoretical treatments while the latter needs to be worked out from the traces of the<br />

transformati<strong>on</strong> matrices that describe how the displacement coordinates qi transform<br />

under the acti<strong>on</strong> of each symmetry element in the group. These manipulati<strong>on</strong>s are<br />

rather cumbersome, especially for the ”icosahedral molecule” arising when a spherical<br />

surface is coated with a pure hexatic layer (see Appendix C).<br />

In the rich literature <strong>on</strong> molecular vibrati<strong>on</strong>s a set of empirical rules has<br />

been developed to write down the characters by examining <strong>on</strong>ly the transformati<strong>on</strong><br />

of the three dimensi<strong>on</strong>al cartesian displacements of the few atoms whose equilibrium


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 86<br />

positi<strong>on</strong>s are not altered by the symmetry operati<strong>on</strong>. In Appendix C we provide<br />

analogous rules that simplify the task of finding the χ (Σ)<br />

i<br />

characters by incorporating<br />

the c<strong>on</strong>straint that each atom is c<strong>on</strong>fined <strong>on</strong> a sphere <str<strong>on</strong>g>and</str<strong>on</strong>g> hence <strong>on</strong>ly two orthog<strong>on</strong>al<br />

displacements need to be c<strong>on</strong>sidered as shown in Eq.(3.39).<br />

The interested reader is referred to Appendix C for a more comprehensive<br />

mathematical justificati<strong>on</strong> of the normal mode analysis applied to the tetrahedral<br />

colloid <str<strong>on</strong>g>and</str<strong>on</strong>g> to the more complicated cases of hexatic Z = 12 <str<strong>on</strong>g>and</str<strong>on</strong>g> tetratic order<br />

Z = 8. Here, we simply summarize the results of applying Eq.(3.43) in c<strong>on</strong>juncti<strong>on</strong><br />

with Table 3.1 to find the degeneracies of the eigenvalue spectrum of the matrix Mij.<br />

The representati<strong>on</strong> Σ c<strong>on</strong>tains (<strong>on</strong>ly <strong>on</strong>ce) the three dimensi<strong>on</strong>al representati<strong>on</strong>s F2<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> F1 as well as the two dimensi<strong>on</strong>al representati<strong>on</strong> E.<br />

Σ = F2 + F1 + E . (3.44)<br />

The three normal coordinates with vanishing frequency corresp<strong>on</strong>d to the three rigid<br />

body rotati<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> bel<strong>on</strong>g to the F1 irreducible representati<strong>on</strong> [64, 65]. We are left<br />

with a doublet (E) <str<strong>on</strong>g>and</str<strong>on</strong>g> a triplet (F2) corresp<strong>on</strong>ding respectively to two shear-like<br />

twisting deformati<strong>on</strong>s of the tetrahedr<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> to three stretching <str<strong>on</strong>g>and</str<strong>on</strong>g> bending modes<br />

of the cords joining neighboring defects.<br />

This symmetry analysis is c<strong>on</strong>firmed by direct diag<strong>on</strong>alizati<strong>on</strong> of the matrix<br />

Mij which leads the following set of eigenvalues λi<br />

{λi} =<br />

3 π K<br />

8<br />

{0, 0, 0, 1, 1, 2, 2, 2} . (3.45)<br />

In Secti<strong>on</strong> C, we also list the eigenvectors wi of Mij. The displacement coordinates


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 87<br />

are readily expressed in terms of the eigenvectors<br />

qi = U −1<br />

ij wj , (3.46)<br />

where the unitary matrix U diag<strong>on</strong>alizes M <str<strong>on</strong>g>and</str<strong>on</strong>g> hence the free energy of Eq.(3.41)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> is defined by<br />

U M U −1 = Diag (λi) . (3.47)<br />

Its c<strong>on</strong>structi<strong>on</strong> is easily achieved by the st<str<strong>on</strong>g>and</str<strong>on</strong>g>ard Gram-Schmidt orthog<strong>on</strong>alizati<strong>on</strong><br />

procedure to the eigenvectors {wi} = {w1, ..., w8}, where the same ordering chosen in<br />

listing the eigenvalues in Eq.(3.45) is implicitly assumed. The resulting orthog<strong>on</strong>al<br />

basis vectors are the rows of the 8 × 8 matrix U.<br />

We are now in a positi<strong>on</strong> to evaluate 〈cos βij〉 where the thermal average is<br />

performed with the Boltzman weight obtained from the free energy in Eq.(3.41) which<br />

is now diag<strong>on</strong>al. Note that for the tetrahedr<strong>on</strong> any choice of pair of defects labelled<br />

by i <str<strong>on</strong>g>and</str<strong>on</strong>g> j (where i �= j) will lead to the same answer, unlike the less symmetric cases<br />

of the twisted cube (p=4) <str<strong>on</strong>g>and</str<strong>on</strong>g> the icosahedr<strong>on</strong> (p=6) c<strong>on</strong>sidered in Appendix C. The<br />

bending angle cos βij in Eq.(3.40) can be Taylor exp<str<strong>on</strong>g>and</str<strong>on</strong>g>ed in the qi. The resulting<br />

expressi<strong>on</strong> is rather cumbersome, but <strong>on</strong>ce the displacements {qi} are reexpressed in<br />

terms of the normal coordinates {wi} (by means of Eq.(3.46)), cos βij reduces to<br />

cos βij = − 1 2 � 2<br />

+ w6 + w<br />

3 9<br />

2 7 + w 2� 8 , (3.48)<br />

where the <strong>on</strong>ly eigenmodes {w6, w7, w8} appearing in Eq.(3.48) corresp<strong>on</strong>d to the<br />

bending triplet of Eq.(3.45).<br />

It is now easy to perform the thermal average by Gaussian integrati<strong>on</strong> of


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 88<br />

the energetically degenerate eigenmodes, with the result [53]<br />

〈cos βij〉 = − 1 2 � 2<br />

+ 〈w<br />

3 9<br />

6〉 + 〈w 2 7〉 + 〈w 2 8〉 �<br />

= − 1 16 kBT<br />

+<br />

3 9πK<br />

. (3.49)<br />

3.4 Valence transiti<strong>on</strong>s in thick nematic shells<br />

In this secti<strong>on</strong>, we study the crossover from a two to three dimensi<strong>on</strong>al<br />

regime as the thickness of the spherical shell, h, increases. For thicker shells, three di-<br />

mensi<strong>on</strong>al defect c<strong>on</strong>figurati<strong>on</strong>s (”escaped” in the third dimensi<strong>on</strong>) compete with the<br />

planar textures described in the previous secti<strong>on</strong>s, leading to a structural transiti<strong>on</strong><br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> a change in valence from Z = 4 to Z = 2 bey<strong>on</strong>d a critical value of h.<br />

We first c<strong>on</strong>sider the case of a cylindrical slab (or disk) of radius R <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

thickness h filled with a nematic whose director is tangent to the two circular faces<br />

[66] (see Fig. 3.8). This simpler geometry captures the essential features of the<br />

problem <str<strong>on</strong>g>and</str<strong>on</strong>g> provides a suitable starting point for underst<str<strong>on</strong>g>and</str<strong>on</strong>g>ing thin spherical shells<br />

(see Fig. 3.9).<br />

3.4.1 Slab geometry<br />

To estimate the energy stored in the texture of Fig. 3.8, we coarse grain<br />

the system to ”blobs” of size h. The elastic energy arises from two sources: a l<strong>on</strong>g<br />

distance c<strong>on</strong>tributi<strong>on</strong> from a radial texture associated with an s = 1 disclinati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

a local energy cost for the elastic deformati<strong>on</strong>s inside the spherical blob in Fig. 3.8.<br />

In the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>, the former can be estimated as the energy


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 89<br />

Figure 3.8: Side view of the 2D nematic texture ansatz soluti<strong>on</strong> in Eq.(3.54). The<br />

cylindrical slab has height h <str<strong>on</strong>g>and</str<strong>on</strong>g> radius R. The arrows can be interpreted as flow<br />

lines of a fluid entering a narrow <str<strong>on</strong>g>and</str<strong>on</strong>g> l<strong>on</strong>g channel from a point source located <strong>on</strong><br />

the top plate. To obtain the nematic texture the vectors need to be normalized <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

viewed as rods with both ends identified as shown in Fig. 3.9.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 90<br />

Figure 3.9: Nematic texture in a thin spherical shell. The nematic director near the<br />

pair of half hedgehogs indicated in the figure is well described by the <strong>on</strong>e calculated<br />

for the slab in Fig. 3.8.<br />

πKh ln � �<br />

R of a disclinati<strong>on</strong> whose enlarged ”core” of size h is given by the spherical<br />

h<br />

blob while the latter is roughly 4πKh, the energy of two half hedgehogs living inside<br />

the blob [67]. In view of the azimuthal symmetry of our c<strong>on</strong>figurati<strong>on</strong>, the director,<br />

n(r, z), can be parameterized by the angle Θ(r, z) formed with respect to the �z axis<br />

of the circular slab al<strong>on</strong>g the centers of the two half hedgehogs shown in Fig. 3.8<br />

n(r, z) = sin Θ(r, z)er + cos Θ(r, z)ez . (3.50)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 91<br />

The energy density f(Θ) = 1<br />

2 K1 (∇ · n) 2 + 1<br />

2 K3 (∇ × n) 2 expressed in terms of the<br />

b<strong>on</strong>d angle reads<br />

f(Θ) = K1<br />

2<br />

�<br />

sin Θ<br />

r + Θr<br />

�2 cos Θ − Θz sin Θ<br />

+ K3<br />

2 (Θz cos Θ + Θr sin Θ) 2 . (3.51)<br />

where K1 <str<strong>on</strong>g>and</str<strong>on</strong>g> K3 are the splay <str<strong>on</strong>g>and</str<strong>on</strong>g> bend c<strong>on</strong>stants <str<strong>on</strong>g>and</str<strong>on</strong>g> Θr ≡ ∂Θ<br />

∂r <str<strong>on</strong>g>and</str<strong>on</strong>g> Θz ≡ ∂Θ<br />

∂z<br />

the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>, minimizati<strong>on</strong> of the free energy leads to a n<strong>on</strong><br />

linear partial differential equati<strong>on</strong> for the b<strong>on</strong>d angle Θ(r, z)<br />

1<br />

r<br />

∂<br />

∂r<br />

�<br />

r ∂Θ<br />

�<br />

∂r<br />

. In<br />

+ ∂2Θ sin 2Θ<br />

=<br />

∂z2 r2 . (3.52)<br />

The operator <strong>on</strong> the left arise from the Laplacian in cylindrical coordinates. 3 .<br />

Instead of solving this partial differential equati<strong>on</strong>, we follow a route anal-<br />

ogous to that in Ref.[68], that is we c<strong>on</strong>struct an exact 2D soluti<strong>on</strong> for the liquid<br />

crystal problem <str<strong>on</strong>g>and</str<strong>on</strong>g> then rotate it al<strong>on</strong>g the axis �z to retrieve an ansatz for the 3D<br />

director c<strong>on</strong>figurati<strong>on</strong> 4 .<br />

To solve the 2D problem we adopt the method of c<strong>on</strong>formal mappings that<br />

simplifies the study of many complicated boundary problems in fluid dynamics <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

electrostatics [45]. Think of Fig. 3.8 as a source of fluid c<strong>on</strong>fined to flow in a narrow<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> l<strong>on</strong>g channel (R ≫ h). The spherical ”half-hedgehog” corresp<strong>on</strong>ds to the source<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the hyperbolic <strong>on</strong>e at the top to the stagnati<strong>on</strong> point of the flow. The complex<br />

3<br />

Note there is no need to explicitly c<strong>on</strong>sider the azimuthal angle φ as an additi<strong>on</strong>al independent variable<br />

in view of the symmetry of the problem.<br />

4<br />

Note that the 2D soluti<strong>on</strong> is the true minimum of the two dimensi<strong>on</strong>al liquid crystal elastic free energy<br />

while the 3D Ansatz does not minimize f(Θ) nor satisfy Eq.(3.52)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 92<br />

potential Ω(w) of the desired flow is<br />

� �<br />

πw<br />

� �<br />

Ω(w) = Log sin − 1<br />

h<br />

, (3.53)<br />

where the complex variable is denoted by w = r + iz to distinguish it from the z<br />

coordinate al<strong>on</strong>g the axis of the cylinder. The velocity field is given by Ω ′ (w) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

the corresp<strong>on</strong>ding complex nematic director n(w) = cos Θ + i sin Θ is obtained by<br />

normalizing this vector field. Note that the prime in Ω ′ (w) denotes the derivative<br />

with respect to w of the complex functi<strong>on</strong> Ω = Φ + iΨ <str<strong>on</strong>g>and</str<strong>on</strong>g> we define Ω ≡ Φ − iΨ. A<br />

straightforward but tedious calculati<strong>on</strong> (see Appendix B) leads the functi<strong>on</strong>al form<br />

of Θ(r, z) for a source at z = − h<br />

2<br />

tan Θ(r, z) = sec<br />

h<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> a stagnati<strong>on</strong> point at z = + , namely<br />

2<br />

�<br />

πz<br />

� �<br />

πr<br />

�<br />

sinh<br />

h h<br />

. (3.54)<br />

This trial soluti<strong>on</strong> respects the boundary c<strong>on</strong>diti<strong>on</strong>s of the problem, has the correct<br />

short <str<strong>on</strong>g>and</str<strong>on</strong>g> l<strong>on</strong>g distance behavior in r <str<strong>on</strong>g>and</str<strong>on</strong>g> the expected functi<strong>on</strong>al form near the<br />

source 5 .<br />

Up<strong>on</strong> inserting Θ(r, z) in Eq.(3.51) <str<strong>on</strong>g>and</str<strong>on</strong>g> integrating the energy density f(Θ)<br />

over the cylindrical volume of the slab, we obtain the total elastic energy E1 stored<br />

in the field 6<br />

� � � �<br />

R<br />

E1 = πKh ln + c<br />

h<br />

, (3.55)<br />

where c ≈ 4.2. Note that the functi<strong>on</strong>al dependence of E1 <strong>on</strong> R <str<strong>on</strong>g>and</str<strong>on</strong>g> h matches<br />

the expectati<strong>on</strong>s from the blob argument. Indeed the prefactor of the logarithm is<br />

5 Similar problems were recently investigated in the c<strong>on</strong>text of complex magnetic patterns of ferromagnetic<br />

nanostructures shaped as strips <str<strong>on</strong>g>and</str<strong>on</strong>g> rings. See O. Tchernyshyov <str<strong>on</strong>g>and</str<strong>on</strong>g> Gia-Wei Chern, c<strong>on</strong>d-mat/0506744 <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

references therein.<br />

6 The resulting energy E1 does not rely <strong>on</strong> the assumpti<strong>on</strong> R ≫ h, as can be explicitly checked numerically.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 93<br />

universal in the sense that it does not depend <strong>on</strong> the details of the trial soluti<strong>on</strong> near<br />

the hedgehogs, but <strong>on</strong>ly <strong>on</strong> its l<strong>on</strong>g distance behavior. By c<strong>on</strong>trast, we expect the<br />

result quoted above for the coefficient c to be <strong>on</strong>ly an estimate (an upper bound)<br />

since it relies <strong>on</strong> a trial soluti<strong>on</strong> for Θ(r, z) that is not the absolute minimum of the<br />

free energy. Numerical studies carried out in Ref.[66] for the slab geometry reported<br />

that c ≈ 4.19.<br />

with two s = 1<br />

2<br />

A competing energy minimum for the nematic is given by a planar texture<br />

disclinati<strong>on</strong> lines. Note that + 1<br />

2<br />

lines cannot escape in the third<br />

dimensi<strong>on</strong> [61]. A tedious but straightforward calculati<strong>on</strong> shows that the energy Ep<br />

of the pair of disclinati<strong>on</strong> lines is given by<br />

Ep = π<br />

2 Kh<br />

�<br />

ln<br />

� �<br />

R<br />

− 0.06 +<br />

a<br />

�<br />

4 Ec<br />

πKh<br />

. (3.56)<br />

where a is a macroscopic cutoff typically of the order of the molecular length. These<br />

two disclinati<strong>on</strong> repel each other, <str<strong>on</strong>g>and</str<strong>on</strong>g> are repelled by the circular boundary, leading<br />

to a separati<strong>on</strong> of order R. The sec<strong>on</strong>d term of Eq.(3.56), corresp<strong>on</strong>ding to the<br />

interacti<strong>on</strong> of the two disclinati<strong>on</strong>s with the boundary <str<strong>on</strong>g>and</str<strong>on</strong>g> am<strong>on</strong>g themselves, is<br />

negligibly small. The third term accounts for the core energies of the two disclinati<strong>on</strong><br />

lines 2Ec. The combinati<strong>on</strong> Kh is equal to the two dimensi<strong>on</strong>al coupling c<strong>on</strong>stant<br />

K2D. The relevant dimensi<strong>on</strong>less ratio is Ec<br />

K2D .<br />

By setting Ep = E1 we obtain the critical thickness h ∗ above which the<br />

escaped ”half-hedgehogs” become energetically favored compared to a single s = 1<br />

2<br />

disclinati<strong>on</strong> line:<br />

h ∗ = e c′√ R a . (3.57)


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 94<br />

The core energy terms in Eq.(3.56) reduce the previous estimate of c to c ′ = c − 2Ec<br />

πKh .<br />

A similar analysis applies to spherical shells which we now discuss.<br />

3.4.2 Extrapolati<strong>on</strong> to thin spherical shells<br />

In principle, <strong>on</strong>e could proceed al<strong>on</strong>g the same route as in the previous<br />

secti<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> find a c<strong>on</strong>formal mapping that provides a trial functi<strong>on</strong> for the b<strong>on</strong>d<br />

angle Θ(θ, r) 7 corresp<strong>on</strong>ding to the texture shown in Fig. 3.9. This is possible but<br />

rather cumbersome. In the case of very thin shells <strong>on</strong>e can adapt the slab calculati<strong>on</strong><br />

by noting again that the energy is composed of two parts. There is a l<strong>on</strong>g distance<br />

piece arising from ”combing the hair” of the nematic texture in the tangent plane of<br />

the sphere that we can read off from a suitable 2D calculati<strong>on</strong> (see Appendix A) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

the short distance c<strong>on</strong>tributi<strong>on</strong> arising from the short distance c<strong>on</strong>tributi<strong>on</strong> arising<br />

from the two pairs of half-hedgehogs at the north <str<strong>on</strong>g>and</str<strong>on</strong>g> south pole.<br />

The energy Ed of two short +1 disclinati<strong>on</strong> lines placed at antipodal points<br />

<strong>on</strong> a sphere (the north <str<strong>on</strong>g>and</str<strong>on</strong>g> south pole, say) can be estimated by performing a 2D<br />

calculati<strong>on</strong> <strong>on</strong> the curved surface <str<strong>on</strong>g>and</str<strong>on</strong>g> simply multiplying the result by the thickness<br />

of the layer h<br />

� � �<br />

R<br />

Ed = 2πKh ln − 0.3 +<br />

a<br />

Ec<br />

�<br />

πKh<br />

, (3.58)<br />

where the middle term accounts for the interacti<strong>on</strong> between the two disclinati<strong>on</strong>s in<br />

their equilibrium positi<strong>on</strong>s. Note that this result is accurate <strong>on</strong>ly up to factors of<br />

the order of � �<br />

h since the explicit integrati<strong>on</strong> over the volume of the thin shell was<br />

R<br />

7 In this case spherical coordinates are adopted for the independent variables.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 95<br />

bypassed. To obtain the energy of the escaped soluti<strong>on</strong>, the core size a in Eq.(3.58) is<br />

rescaled to h. This will account for the integrati<strong>on</strong> of the energy density at distances<br />

of the order of a few h ≪ R from the two hedgehogs 8 .<br />

The energy stored in the remaining porti<strong>on</strong>s of the thin shell is approx-<br />

imately given by twice the energy 4.2 πKh of the yellow blob of Fig. 3.8. This<br />

estimate neglects curvature correcti<strong>on</strong>s of the order of � �<br />

h <str<strong>on</strong>g>and</str<strong>on</strong>g> arises because at dis-<br />

R<br />

tances of the order h the spherical shell looks locally like a flat circular slab as l<strong>on</strong>g<br />

as h ≪ R. The resulting energy E2 of the escaped c<strong>on</strong>figurati<strong>on</strong> reads<br />

� � � �<br />

R<br />

E2 = 2πKh ln − 0.3 + 4.2<br />

h<br />

. (3.59)<br />

Although the prefactor of the sub-leading term linear in h has <strong>on</strong>ly been estimated, we<br />

expect that the coefficients of the logarithm, which arises from large scales compared<br />

to h, is exact. For a spherical shell whose radius R is a hundred times its thickness,<br />

the correcti<strong>on</strong>s from higher powers of h<br />

R<br />

are indeed negligible. However for reas<strong>on</strong>able<br />

values of R,<br />

the logarithmic term of Eq.(3.59) is still comparable in magnitude to the<br />

h<br />

”sub-leading” <strong>on</strong>e linear in h.<br />

The energy E4 of the tetravalent c<strong>on</strong>figurati<strong>on</strong> can be evaluated using similar<br />

c<strong>on</strong>siderati<strong>on</strong>s, with the result<br />

� � �<br />

R<br />

E4 = πKh ln − 0.4 +<br />

a<br />

4Ec<br />

�<br />

Kπh<br />

. (3.60)<br />

Up<strong>on</strong> setting E4 = E2 we obtain the critical thickness h ∗ below which the tetravalent<br />

8 In these porti<strong>on</strong>s of the shell the integr<str<strong>on</strong>g>and</str<strong>on</strong>g> reduces to the energy density of the two disclinati<strong>on</strong> problem<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> hence the integrati<strong>on</strong> can be easily carried out leading the result in Eq.(3.58) with a lower cutoff of the<br />

order h.


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 96<br />

c<strong>on</strong>figurati<strong>on</strong> becomes energetically favored<br />

h ∗ 2Ec<br />

(4.1−<br />

= e Kπh )√ R a . (3.61)<br />

The exp<strong>on</strong>ential prefactor arises from the terms linear in h in Eq.(3.59), which cannot<br />

be ignored in estimating h ∗ even in the limit R ≫ h. Note that an accurate determi-<br />

nati<strong>on</strong> of the argument in the exp<strong>on</strong>ent would require knowledge of the core energies<br />

of the disclinati<strong>on</strong> lines 9 .<br />

The energy barrier between these two coexisting minima of the free energy<br />

f(Θ) can be estimated by splitting the path c<strong>on</strong>necting them in Θ space in two steps.<br />

First, c<strong>on</strong>sider a c<strong>on</strong>tinuous deformati<strong>on</strong> of the escaped texture of Fig. 3.9 obtained<br />

by appropriately rotating each nematigen until the n<strong>on</strong> − escaped soluti<strong>on</strong> (with two<br />

disclinati<strong>on</strong>s lines of index <strong>on</strong>e at the north <str<strong>on</strong>g>and</str<strong>on</strong>g> south pole) is recovered. This part<br />

of the path must be uphill in energy if the escaped soluti<strong>on</strong> was allowed to escape in<br />

the first place. The corresp<strong>on</strong>ding energy barrier ∆E is given approximately by the<br />

difference between Ed as calculated in Eq.(3.58) <str<strong>on</strong>g>and</str<strong>on</strong>g> E2 in Eq.(3.59)<br />

split in two + 1<br />

2<br />

� � � �<br />

h<br />

∆E = 2πKh ln − 4.2<br />

a<br />

. (3.62)<br />

The sec<strong>on</strong>d step c<strong>on</strong>sists in letting each of the unstable disclinati<strong>on</strong> lines<br />

defects <str<strong>on</strong>g>and</str<strong>on</strong>g> subsequently separate them until they sit at the vertexes<br />

of a tetrahedr<strong>on</strong> inscribed in the sphere. This porti<strong>on</strong> of the path is downhill because<br />

the ”n<strong>on</strong>-escaped” texture of valence 2 is unstable 10 . As a result the energy barrier<br />

is simply the energy difference calculated in Eq.(3.62).<br />

9<br />

In fact, the exp<strong>on</strong>ential prefactor can be interpreted as a numerically significant rescaling of the core<br />

radius.<br />

10<br />

This can be proved by writing down the energy of the pair <str<strong>on</strong>g>and</str<strong>on</strong>g> show that it decreases m<strong>on</strong>ot<strong>on</strong>ically as<br />

<strong>on</strong>e separates them because of the ”electrostatic-like” repulsi<strong>on</strong> [12, 58].


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 97<br />

Up<strong>on</strong> inserting K ≈ 10 −6 dyn in Eq.(3.62) <str<strong>on</strong>g>and</str<strong>on</strong>g> taking kBT ≈ 4 10 −14 erg,<br />

as in Ref.[69, 67], we obtain<br />

∆E<br />

kBT<br />

� � �<br />

15h h<br />

≈ ln − 4.2 +<br />

nm a<br />

Ec<br />

�<br />

Kπh<br />

For shells with critical thickness h ∗ Eq.(3.63) reduces to<br />

∆E<br />

kBT<br />

�<br />

R<br />

≈ 103<br />

a ln<br />

� �<br />

R<br />

a<br />

. (3.63)<br />

, (3.64)<br />

where a core size of the order of 10 nm was assumed <str<strong>on</strong>g>and</str<strong>on</strong>g> the core energies were set<br />

to zero [12]. This estimate indicates that the energy barrier is very high around h ∗<br />

suggesting that exchange between the two minima is unlikely to happen by thermal<br />

activati<strong>on</strong>. In a m<strong>on</strong>odisperse soluti<strong>on</strong> of shells with thickness h, the ratio between<br />

shells of the two valence will be given by their Boltzman factors as l<strong>on</strong>g as equilibrium<br />

is reached. If <strong>on</strong>e engineers shells with thickness below h ∗ , the Z=4 c<strong>on</strong>figurati<strong>on</strong><br />

would be more likely.<br />

3.5 C<strong>on</strong>clusi<strong>on</strong><br />

We have studied the crossover from the two dimensi<strong>on</strong>al regime of liquid<br />

crystals c<strong>on</strong>fined <strong>on</strong> a spherical surface to the full three dimensi<strong>on</strong>al problem in a<br />

spherical shell. For very thin shells, the nematic ground state has four disclinati<strong>on</strong><br />

lines sitting at the vertices of a tetrahedr<strong>on</strong> inscribed in the ball <str<strong>on</strong>g>and</str<strong>on</strong>g> whose texture<br />

approximately track the seam of a baseball. As the thickness increases, a competing<br />

three dimensi<strong>on</strong>al defected texture characterized by two pairs of half hedgehogs at<br />

the north <str<strong>on</strong>g>and</str<strong>on</strong>g> south pole becomes energetically favorable. For ultra-thin shells this


Chapter 3: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures in thick spherical shells. 98<br />

instability is suppressed <str<strong>on</strong>g>and</str<strong>on</strong>g> <strong>on</strong>e expects a defected ground state with tetravalent<br />

symmetry. Estimates of the stability of this texture to thermal fluctuati<strong>on</strong>s indicate<br />

that the vibrati<strong>on</strong>s around the equilibrium c<strong>on</strong>figurati<strong>on</strong>s of the defects should not be<br />

significant. The present analysis has been carried out primarily in the limit in which<br />

the elastic anisotropy parameter, ɛ = K3−K1<br />

K3+K1<br />

≪ 1.<br />

We hope to extend our investigati<strong>on</strong> with a systematic study of the effect of<br />

elastic anisotropy <strong>on</strong> the nematic texture. It is interesting to note that in the case of<br />

pure bend or splay (ie. ɛ = ±1) the ground state is given by <strong>on</strong>ly two disclinati<strong>on</strong>s of<br />

unit index at the north <str<strong>on</strong>g>and</str<strong>on</strong>g> south pole. This suggests the possibility that the effect<br />

of the elastic anisotropy may not be limited to locally adjusting the orientati<strong>on</strong> of the<br />

director but may induce a change in the inter-defect interacti<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> hence a distorti<strong>on</strong><br />

of the tetravalent equilibrium c<strong>on</strong>figurati<strong>on</strong>. The limit of str<strong>on</strong>g elastic anisotropy is<br />

also relevant to studies of the nematic to smectic transiti<strong>on</strong> in a spherical geometry<br />

for which the ratio of the bend to splay coupling c<strong>on</strong>stants K3<br />

K1<br />

is expected to diverge.<br />

Additi<strong>on</strong>al experimental complicati<strong>on</strong>s include the possibility of having a<br />

nematic layer of n<strong>on</strong> c<strong>on</strong>stant thickness that would induce trapping of the defects in<br />

the regi<strong>on</strong>s where the layer is thinner. This effect may also induce a local transiti<strong>on</strong><br />

to an escaped texture where the layer thickens in just <strong>on</strong>e hemisphere so that two<br />

disclinati<strong>on</strong> lines of index 1<br />

2<br />

shells with three-fold symmetry could be observed.<br />

are traded for a pair of half hedgehogs. If that happens


Chapter 4<br />

<str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved<br />

surface.<br />

Based <strong>on</strong> V. Vitelli <str<strong>on</strong>g>and</str<strong>on</strong>g> A. M. Turner PRL 93, 215301(2004).<br />

99


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 100<br />

The physics of topological defects <strong>on</strong> curved surfaces plays an increasingly<br />

significant role in the engineering of devices based <strong>on</strong> coated interfaces. [52, 53,<br />

3]. Defects also affect the mechanical properties of some biological systems, such as<br />

spherical viruses, whose shape is dependent <strong>on</strong> the presence of disclinati<strong>on</strong>s in their<br />

protein shell [70]. Furthermore, the effects induced by a curved substrate <strong>on</strong> the<br />

distributi<strong>on</strong> of defects are not fully understood even in well studied systems such as<br />

thin superfluid or superc<strong>on</strong>ducting films. In this paper, we study simple c<strong>on</strong>tinuum<br />

generalizati<strong>on</strong>s of the plane XY model to frozen surfaces of varying curvature to gain<br />

a broad underst<str<strong>on</strong>g>and</str<strong>on</strong>g>ing of the interacti<strong>on</strong> between topological defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature.<br />

The XY model is a simple setting in which particle-like objects emerge from<br />

a more fundamental theory. The basic degree of freedom is an angle-valued functi<strong>on</strong><br />

<strong>on</strong> the plane whose values vary from 0 to 2π. These angles could represent the<br />

orientati<strong>on</strong>s of interacting arrows. The interacti<strong>on</strong>, which tends to align neighboring<br />

arrows, results from the c<strong>on</strong>tinuum free energy F given by<br />

F = K<br />

2<br />

�<br />

d 2 u (∇θ (u )) 2 , (4.1)<br />

where the set of coordinates u = (x, y) label points <strong>on</strong> the plane. Despite its simplicity,<br />

this model captures the main properties of vortices in layers of superfluid 4 He or thin<br />

superc<strong>on</strong>ducting films when the field θ (u ) is identified with the phase of the collective<br />

wave functi<strong>on</strong>. In additi<strong>on</strong>, the elastic energy of Eq.(4.1) correctly describes liquid<br />

crystalline phases for which the b<strong>on</strong>d angle, θ (u ), has periodicity 2π<br />

p<br />

with p ≥ 3.<br />

For a soluti<strong>on</strong> of nematigens (p = 2) <str<strong>on</strong>g>and</str<strong>on</strong>g> tilted molecules in a a Langmuir film<br />

(p = 1) two different elastic c<strong>on</strong>stants are necessary to account for bend <str<strong>on</strong>g>and</str<strong>on</strong>g> splay


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 101<br />

deformati<strong>on</strong>s [38], but these are renormalized to the same value at finite temperatures<br />

[71]. Besides its experimental significance, the XY model is the cornerst<strong>on</strong>e of our<br />

c<strong>on</strong>ceptual underst<str<strong>on</strong>g>and</str<strong>on</strong>g>ing of topological defects, singular c<strong>on</strong>figurati<strong>on</strong>s of the field<br />

θ (u ).<br />

Like particles, defects have charges <str<strong>on</strong>g>and</str<strong>on</strong>g> a characteristic Coulomb-like inter-<br />

acti<strong>on</strong>. The charge q, a multiple of 2π,<br />

can be defined by the amount θ increases al<strong>on</strong>g<br />

p<br />

a path enclosing the defect’s core. The force between two defects located at positi<strong>on</strong>s<br />

ui <str<strong>on</strong>g>and</str<strong>on</strong>g> uj is determined by the energy stored in the θ field, K qi qj U(ui, uj), where<br />

the inter-particle potential U(ui, uj) is proporti<strong>on</strong>al to the logarithm of the distance<br />

in the plane. On a flat surface, thin layers of superfluids, superc<strong>on</strong>ductors <str<strong>on</strong>g>and</str<strong>on</strong>g> liquid<br />

crystals can all be analyzed within the framework of Eq.(4.1) [72]. However, there<br />

is a crucial difference between, say, the phase of the superfluid order parameter <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

the angle that describes the local orientati<strong>on</strong> of liquid crystal molecules. The for-<br />

mer transforms like a scalar since the quantum mechanical phase does not change<br />

when the system is rotated, while the latter represents a vector aligned to the local<br />

directi<strong>on</strong> of the molecules. Thus, a comm<strong>on</strong> boundary c<strong>on</strong>diti<strong>on</strong> for a liquid crystal<br />

is for the director to be tangent to the boundary of the substrate. By c<strong>on</strong>trast, no<br />

such c<strong>on</strong>straint exists for a 4 He film because its wave functi<strong>on</strong> is defined in a different<br />

space from the <strong>on</strong>e in which the superfluid is c<strong>on</strong>fined. This distincti<strong>on</strong> is crucial<br />

<strong>on</strong> a curved surface. In the ground state of a 4 He film, the phase θ(u) is c<strong>on</strong>stant<br />

throughout the surface <str<strong>on</strong>g>and</str<strong>on</strong>g> the corresp<strong>on</strong>ding energy vanishes. The free energy Fs to


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 102<br />

be minimized is a scalar generalizati<strong>on</strong> of Eq.(4.1):<br />

Fs = K<br />

2<br />

�<br />

d 2 u √ g g αβ ∂αθ(u) ∂βθ(u) . (4.2)<br />

Here the set of coordinates u = (u1, u2) label points <strong>on</strong> the surface while √ g is the<br />

determinant of the metric tensor gαβ. On the other h<str<strong>on</strong>g>and</str<strong>on</strong>g>, a c<strong>on</strong>stant b<strong>on</strong>d angle θ(u)<br />

is not the ground state of the liquid crystal because it is measured with respect to<br />

an arbitrary basis vector Eα(u) with α = 1, 2. Indeed, it is not possible to make the<br />

directi<strong>on</strong>s of the molecules parallel everywhere <strong>on</strong> a curved space; the lowest energy<br />

state is attained by optimally distributing the unavoidable bend <str<strong>on</strong>g>and</str<strong>on</strong>g> splay of the<br />

vectors over the whole surface. The free energy functi<strong>on</strong>al Fv to be minimized is a<br />

vector generalizati<strong>on</strong> of Eq.(4.1) [40]:<br />

Fv = K<br />

2<br />

�<br />

d 2 u √ gg αβ (∂αθ(u) − Ωα(u))(∂βθ(u) − Ωβ(u)) , (4.3)<br />

where Ωα(u), the c<strong>on</strong>necti<strong>on</strong>, compensates for the rotati<strong>on</strong> of the 2D basis vectors<br />

Eα(u) in the directi<strong>on</strong> of uα. Since the curl of Ωα(u) is equal to G(u) [33], the<br />

integr<str<strong>on</strong>g>and</str<strong>on</strong>g> in Eq.(4.3) cannot be made to vanish <strong>on</strong> a surface with n<strong>on</strong>-zero Gaussian<br />

curvature 1 . As the substrate becomes more curved, the energy cost of this geometric<br />

frustrati<strong>on</strong> can be lowered by generating defects in the ground state even in the<br />

absence of topological c<strong>on</strong>straints [73, 74].<br />

In this letter, we introduce a novel coupling between a defect <str<strong>on</strong>g>and</str<strong>on</strong>g> the varying<br />

curvature of the substrate which originates in a c<strong>on</strong>formal anomaly of the free energies<br />

of Equati<strong>on</strong>s (4.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> (4.3). This anomaly arises, even at zero temperature, from<br />

1 Ωα(u) is a n<strong>on</strong>-c<strong>on</strong>servative field <str<strong>on</strong>g>and</str<strong>on</strong>g> hence cannot be equal to ∂αθ everywhere.


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 103<br />

imposing a c<strong>on</strong>stant cutoff, a, localized at the core of each defect 2 . By c<strong>on</strong>trast,<br />

finite temperature c<strong>on</strong>formal anomalies [46] are generated by the presence of a short<br />

wavelength cutoff for the fluctuati<strong>on</strong>s in θ(u) at every point <strong>on</strong> the surface. A physical<br />

c<strong>on</strong>sequence of the anomalous coupling is that topological defects in superfluids <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

superc<strong>on</strong>ductors interact with the curvature in a radically different way from the case<br />

of liquid crystal order 3 .<br />

For thin layers of superfluids <str<strong>on</strong>g>and</str<strong>on</strong>g> superc<strong>on</strong>ductors, we prove that the geo-<br />

metric interacti<strong>on</strong> E s (ui) is given by:<br />

E s (ui) = − K<br />

4π q2 i V (ui) , (4.4)<br />

where ui <str<strong>on</strong>g>and</str<strong>on</strong>g> qi are respectively the positi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> topological charge of the defect. The<br />

geometric potential V (u) satisfies a covariant versi<strong>on</strong> of Poiss<strong>on</strong>’s equati<strong>on</strong> where the<br />

negative of the Gaussian curvature G(u) plays the role of the charge density:<br />

DαD α V (u) = G(u) . (4.5)<br />

For an azimuthally symmetric surface such as the bump represented in Fig. 4.1,<br />

we can explicitly obtain V (u), as a functi<strong>on</strong> of the radial distance from the top, by<br />

employing a two dimensi<strong>on</strong>al analogue of Gauss’ law [74]. The resulting potential well<br />

V (r) vanishes at infinity <str<strong>on</strong>g>and</str<strong>on</strong>g> its width <str<strong>on</strong>g>and</str<strong>on</strong>g> depth are given respectively by the linear<br />

size of the bump <str<strong>on</strong>g>and</str<strong>on</strong>g> its aspect ratio squared. Eq.(C.7) has an elegant geometrical<br />

interpretati<strong>on</strong> if a set of coordinates is chosen so that the metric tensor is cast in<br />

2<br />

We assume that a is much smaller than the radius of curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> hence independent of the defect<br />

positi<strong>on</strong>.<br />

3<br />

In both cases, the electrostatic interacti<strong>on</strong> between the defects U(ui, uj) is given by the Green’s functi<strong>on</strong><br />

of the covariant Laplacian that is not translati<strong>on</strong>ally invariant <str<strong>on</strong>g>and</str<strong>on</strong>g> depends <strong>on</strong> the shape of the surface [74].


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 104<br />

the form gαβ = δαβ exp (−2ω(u)) [40]. The c<strong>on</strong>formal factor ω (u) is c<strong>on</strong>trolled by<br />

the overall shape of the surface <str<strong>on</strong>g>and</str<strong>on</strong>g> it satisfies the same Poiss<strong>on</strong> Eq.(C.7) as the<br />

geometric potential [40]. We therefore proceed with the identificati<strong>on</strong> of V (u) with<br />

ω(u) 4 . This observati<strong>on</strong> will be the basis of our proof of Eq.(4.4) which results in<br />

the novel predicti<strong>on</strong> that a vortex in a superfluid or superc<strong>on</strong>ducting film is repelled<br />

(attracted) by positive (negative) Gaussian curvature irrespective of its charge <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

sign.<br />

For liquid crystals, the geometric interacti<strong>on</strong> E v (ui) c<strong>on</strong>tains an additi<strong>on</strong>al<br />

term discussed in previous investigati<strong>on</strong>s of hexatic membranes [35], which arises<br />

from the geometric frustrati<strong>on</strong> of the vector field. This term, linear in q, happens to<br />

c<strong>on</strong>tain the same functi<strong>on</strong> V (u) as the c<strong>on</strong>formal anomaly of Eq.(4.4). When both<br />

c<strong>on</strong>tributi<strong>on</strong>s are included, E v (ui) acquires an unexpected dependence <strong>on</strong> the charge<br />

of the defect:<br />

E v (ui) = K qi<br />

�<br />

1 − qi<br />

�<br />

V (ui) . (4.6)<br />

4π<br />

The interpretati<strong>on</strong> of V (u) as a geometric potential <str<strong>on</strong>g>and</str<strong>on</strong>g> the linear dependence <strong>on</strong> q in<br />

the first term of Eq.(4.6) are c<strong>on</strong>sistent with the general belief that a defect interacts<br />

logarithmically with the Gaussian curvature, as an electrostatic particle would with<br />

a background charge distributi<strong>on</strong>. However, E v (ui) does not grow linearly with the<br />

charge of the defect, as expected from the electrostatic analogy. Instead, the geometric<br />

interacti<strong>on</strong> peaks for a defect of charge 2π <str<strong>on</strong>g>and</str<strong>on</strong>g> eventually becomes repulsive for q<br />

greater than the critical charge qc = 4π 5 .<br />

4<br />

This identificati<strong>on</strong> is possible for corrugated planes flat at infinity. The details of the analysis valid for<br />

deformed spheres <str<strong>on</strong>g>and</str<strong>on</strong>g> torii will be presented elsewhere.<br />

5 q−2π<br />

The symmetry around q = 2π is reflected in the pattern of field lines around a defect. There are | π |


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 105<br />

Figure 4.1: A corrugated substrate <str<strong>on</strong>g>and</str<strong>on</strong>g> its downward projecti<strong>on</strong> <strong>on</strong> a flat plane. The<br />

shaded strip surrounding P is more stretched than the <strong>on</strong>e surrounding Q despite<br />

their projecti<strong>on</strong>s <strong>on</strong>to the plane having the same area. The energy stored in the field<br />

will be lower if the core of the defect is located at Q rather than P .<br />

The quadratic coupling has an intuitive explanati<strong>on</strong> in the case of azimuthally<br />

symmetric surfaces. C<strong>on</strong>sider a very thin superfluid film deposited <strong>on</strong> the surface il-<br />

lustrated in Fig. 4.1 with a vortex of charge q placed <strong>on</strong> top of the bump. In order<br />

to calculate the energy stored in the field, we <strong>on</strong>ly need to know that the superfluid<br />

phase θ(u) changes uniformly by q al<strong>on</strong>g a circumference of length 2πr centered <strong>on</strong><br />

the defect. Inspecti<strong>on</strong> of Eq.(4.2) reveals that the energy density of the field in the<br />

shaded strip at distance r is proporti<strong>on</strong>al to � � q 2,<br />

where r is the distance to the sin-<br />

r<br />

gularity measured in the plane of projecti<strong>on</strong> (see Fig. 4.1). By vertically stretching<br />

the surface, the amount of area in the shaded strip is increased with respect to its<br />

lobes.


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 106<br />

projecti<strong>on</strong> <strong>on</strong> the plane, while the energy density is unchanged. As a result, the total<br />

energy stored in the field is greater when a vortex sits <strong>on</strong> top of a bumpy surface than<br />

when the same vortex is located at the center of a flat disk of the same area. Hence, it<br />

is energetically favorable for the vortex to migrate to the flat porti<strong>on</strong>s of the surface.<br />

In this case, the vortex is far away from the bump so that the total energy stored<br />

in the field does not differ much from the flat plane result [75]. For less symmetric<br />

surfaces, the resulting geometric interacti<strong>on</strong> will depend <strong>on</strong> the shape of the entire<br />

surface as embedded in the metric tensor.<br />

The physical origin of the linear coupling between defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature in<br />

Eq.(4.6) is illustrated in Fig. 4.2 for a disclinati<strong>on</strong> of charge 2π centered <strong>on</strong> a bump.<br />

As the curvature of the bump is increased, the bend or splay of the director of the<br />

liquid crystal decreases <str<strong>on</strong>g>and</str<strong>on</strong>g> hence the energy stored in the vector field is reduced.<br />

As a result, this linear coupling causes positively (negatively) charged defects to be<br />

attracted by positive (negative) Gaussian curvature [35]. However, this mechanism<br />

competes with the repulsive geometric interacti<strong>on</strong> illustrated in Fig. 4.1 that is at<br />

work also in the case of liquid crystal order. We note that the linear coupling is absent<br />

for superfluids because, in Eq.(4.2), ∂αθ(u) is not coupled to a curvature dependent<br />

c<strong>on</strong>necti<strong>on</strong>, Ωα(u), as it is in Eq.(4.3). The critical value qc = 4π, where the single<br />

defect potential E v (ui) changes sign, can be determined from simple geometrical ar-<br />

guments. C<strong>on</strong>sider an isolated disclinati<strong>on</strong> of charge q <strong>on</strong> a hemispherical cup placed<br />

<strong>on</strong> a flat plane. On account of azimuthal symmetry, the force acting <strong>on</strong> the defect<br />

depends <strong>on</strong>ly <strong>on</strong> the net Gaussian curvature enclosed by the circle <strong>on</strong> which it is


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 107<br />

Figure 4.2: Disclinati<strong>on</strong>s of charge 2π located <strong>on</strong> top of bumps with different aspect<br />

ratios. The amount of splay in the liquid crystal director <strong>on</strong> the taller bump is reduced<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> hence the energy density is lower.<br />

placed, see Fig. 4.3a [74]. This interacti<strong>on</strong> is unchanged if we deform the outer regi<strong>on</strong><br />

of the plane <str<strong>on</strong>g>and</str<strong>on</strong>g> eventually compactify it to form a sphere as illustrated in Fig. 4.3b.<br />

In order to satisfy topological c<strong>on</strong>straints 6 , we still need to place a shadow defect of<br />

charge (4π − q) at the south pole (the <strong>on</strong>ly positi<strong>on</strong> outside the circle that does not<br />

destroy the azimuthal symmetry of the initial problem). The curvature-defect interac-<br />

ti<strong>on</strong> <strong>on</strong> the hemisphere is thus reduced to the well known defect-defect interacti<strong>on</strong> <strong>on</strong><br />

the sphere [12]. The latter is proporti<strong>on</strong>al to q(4π − q) <str<strong>on</strong>g>and</str<strong>on</strong>g> so is the curvature-defect<br />

interacti<strong>on</strong> <strong>on</strong> the deformed plane of Fig. 4.3a, in agreement with Eq.(4.6). This<br />

provides evidence that a disclinati<strong>on</strong> of charge greater than 4π will be repelled from<br />

6 The sum of the defects charges <strong>on</strong> a surface of genus g is equal to 4π(1 − g) for a vector field <str<strong>on</strong>g>and</str<strong>on</strong>g> 0 for<br />

a scalar [40].


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 108<br />

(a) (b)<br />

Figure 4.3: (a) An isolated disclinati<strong>on</strong> <strong>on</strong> a deformed plane feels a force that depends<br />

<strong>on</strong>ly <strong>on</strong> the enclosed Gaussian curvature. (b) The deformed plane is compactified to<br />

the sphere by placing a shadow defect at the south pole.<br />

regi<strong>on</strong>s of positive curvature. We now present a derivati<strong>on</strong> of the coupling between<br />

curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> defects in helium <str<strong>on</strong>g>and</str<strong>on</strong>g> superc<strong>on</strong>ducting films that employs the method<br />

of c<strong>on</strong>formal mapping, often adopted in electromagnetism <str<strong>on</strong>g>and</str<strong>on</strong>g> fluid mechanics to sim-<br />

plify the boundary of complicated planar regi<strong>on</strong>s. In this c<strong>on</strong>text, we use c<strong>on</strong>formal<br />

mappings to relate the complex task of finding the field energy <strong>on</strong> an arbitrarily de-<br />

formed target surface to an equivalent problem <strong>on</strong> a homogeneous reference surface<br />

(see Fig. 4.4). A c<strong>on</strong>formal mapping has two equivalent defining properties: angles<br />

map to equal angles, <str<strong>on</strong>g>and</str<strong>on</strong>g> very small figures map to figures of nearly the same shape.<br />

One can always find a c<strong>on</strong>formal mapping from the target to the reference surface<br />

[40] such that gT = e −2ω(u) gR, where gT <str<strong>on</strong>g>and</str<strong>on</strong>g> gR are the metric tensors <strong>on</strong> the target


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 109<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> reference surfaces respectively. The scaling factor e ω(u) varies with the positi<strong>on</strong> u<br />

<strong>on</strong> the target surface, so that larger figures are inhomogeneously distorted when they<br />

are mapped from the target to the reference surface. We choose the reference surfaces<br />

to be undeformed <str<strong>on</strong>g>and</str<strong>on</strong>g> of the same topology as the target spaces (eg., gR = δαβ for<br />

a corrugated plane). Defects <strong>on</strong> the target surface are mapped <strong>on</strong>to a set of “image<br />

defects” <strong>on</strong> the reference surface.<br />

The crucial property of the scalar free energy Fs is its invariance under the<br />

rescaling of the metric by the c<strong>on</strong>formal factor 7 . However, the c<strong>on</strong>formal symmetry of<br />

Fs is broken up<strong>on</strong> introducing a short distance cutoff a that is necessary to prevent the<br />

energy from diverging in the core of the defect. Because of the varying scaling factor,<br />

the c<strong>on</strong>stant physical core radius a is stretched or c<strong>on</strong>tracted when projected <strong>on</strong> the<br />

reference space by an amount dependent <strong>on</strong> the positi<strong>on</strong> of the defect (see Fig. 4.4).<br />

The radius of the image of the i th core is ai = e ω(ui) a. It is this c<strong>on</strong>formal anomaly that<br />

is resp<strong>on</strong>sible for generating the geometric interacti<strong>on</strong> in Eq.(4.4). In fact, the energy<br />

of the defects in the target space ET is equal to the energy of a c<strong>on</strong>figurati<strong>on</strong> of defects<br />

(<strong>on</strong> the reference surface) whose core radii are positi<strong>on</strong>-dependent. This problem can<br />

be further transformed into the simpler task of finding the energy ER for a set of<br />

interacting defects with c<strong>on</strong>stant core radius a plus an effective geometric potential<br />

that accounts for the variati<strong>on</strong> of the core size with positi<strong>on</strong>. This geometric potential<br />

can be derived with the aid of Fig. 4.4. If ai is smaller (larger) than a, the energies<br />

stored in the annular regi<strong>on</strong>s indicated in Fig. 4.4 need to be added (subtracted)<br />

7 When gT is substituted in Eq.(4.2) the c<strong>on</strong>formal factor e −2ω(u) cancels out <str<strong>on</strong>g>and</str<strong>on</strong>g> √ gT g αβ<br />

T = √ gR g αβ<br />

R<br />

[74].


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 110<br />

Figure 4.4: C<strong>on</strong>formal mapping of the target surface T <strong>on</strong>to the reference space<br />

R. The c<strong>on</strong>tinuous disks <strong>on</strong> both surfaces represent the “physical” cores of c<strong>on</strong>stant<br />

radius a. The dashed lines represent the positi<strong>on</strong> dependent images <strong>on</strong> R of the defect<br />

cores <strong>on</strong> T with variable radii ai. Note that the energy stored in the annuli comprised<br />

by the dashed <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>tinuous lines in R must be added or subtracted to ER to obtain<br />

ET .<br />

from ER to obtain ET . To calculate this extra energy, we introduce a set of polar<br />

coordinates (r,φ) centered <strong>on</strong> the i th defect. Near the defect of charge qi, the phase<br />

is given by θ ≈ qi<br />

2π φ <str<strong>on</strong>g>and</str<strong>on</strong>g> the energy density is Kq2 i<br />

8π 2 r 2 . Up<strong>on</strong> integrating it over the<br />

annulus comprised between a <str<strong>on</strong>g>and</str<strong>on</strong>g> ai = e ω(ui) a (see Fig. 4.2), we obtain<br />

ET − ER = −K<br />

Nd �<br />

i=1<br />

q 2 i<br />

4π ω(ui) , (4.7)<br />

where Nd is the number of defects. The energy ER accounts for defect-defect inter-<br />

acti<strong>on</strong>s since any potential felt by a single defect would have to be c<strong>on</strong>stant because<br />

all points are equivalent <strong>on</strong> the reference surface (undeformed sphere or plane). Re-<br />

calling that ω(u) = V (u), we recover the result of Eq.(4.4) with no dependence <strong>on</strong>


Chapter 4: <str<strong>on</strong>g>Superfluid</str<strong>on</strong>g> films <strong>on</strong> a curved surface. 111<br />

the microscopic physics because the core size a drops out in Eq.(4.7) 8 . In the case<br />

of liquid crystal order, the c<strong>on</strong>tributi<strong>on</strong> of the anomaly is simply added to the term<br />

linear in q as indicated in Eq.(4.6) [74].<br />

Experiments that test our predicti<strong>on</strong>s can be realized by coating a bump<br />

with a thin layer of superfluid helium <str<strong>on</strong>g>and</str<strong>on</strong>g> rotating it around its axis of symmetry so<br />

that a single vortex forms [76]. The competiti<strong>on</strong> between the (repulsive) geometric<br />

interacti<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> the c<strong>on</strong>fining parabolic potential (generated by the rotati<strong>on</strong>) would<br />

cause the equilibrium positi<strong>on</strong> of the vortex to shift from the center of the bump if<br />

its height exceeds a critical value. The vortex line could be detected by trapping of<br />

electr<strong>on</strong>s <strong>on</strong> its core [77]. Other experiments may detect an inhomogeneous distribu-<br />

ti<strong>on</strong> of thermally induced defects resulting from the combined effect of the anomalous<br />

coupling <str<strong>on</strong>g>and</str<strong>on</strong>g> the dependence of their Coulomb-like interacti<strong>on</strong> <strong>on</strong> the varying curva-<br />

ture.<br />

We have dem<strong>on</strong>strated that the interacti<strong>on</strong> between defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature in<br />

2D XY-like models depends crucially <strong>on</strong> the nature of the underlying order parameter<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> we have shown how to explicitly derive the resulting geometric force from the<br />

shape of the surface.<br />

8 Additi<strong>on</strong>al couplings resulting from other terms in a L<str<strong>on</strong>g>and</str<strong>on</strong>g>au expansi<strong>on</strong> introduce negligible correcti<strong>on</strong>s<br />

proporti<strong>on</strong>al to the inverse of the area of the surface.


Chapter 5<br />

Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong><br />

curved surfaces.<br />

Based <strong>on</strong> V. Vitelli, J. B. Lucks <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>, c<strong>on</strong>d-mat/0604203<br />

112


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 113<br />

5.1 Introducti<strong>on</strong><br />

The physics of two dimensi<strong>on</strong>al crystals <strong>on</strong> curved substrates is emerging as<br />

an intriguing route to the engineering of self assembled systems such as the ”colloi-<br />

dosome”, a colloidal armor used for drug delivery [3], or devices based <strong>on</strong> ordered<br />

arrays of block copolymers which are a promising tool for “soft lithography” [54, 78].<br />

<strong>Curved</strong> crystalline order also affects the mechanical properties of biological structures<br />

like clathrin-coated pits [5, 6] or HIV viral capsids [8, 9] whose irregular shapes appear<br />

to induce a n<strong>on</strong>-uniform distributi<strong>on</strong> of disclinati<strong>on</strong>s in their shell [79].<br />

In this chapter, we present a theoretical <str<strong>on</strong>g>and</str<strong>on</strong>g> numerical study of point-like<br />

defects in a“soft” crystalline m<strong>on</strong>olayer grown <strong>on</strong> a rigid substrate of varying Gaus-<br />

sian curvature with lattice c<strong>on</strong>stant of order, say, 10 nm or more. The substrate can<br />

then be assumed smooth <strong>on</strong> the scale of the m<strong>on</strong>olayer lattice c<strong>on</strong>stant, as would be<br />

the case for di-block copolymers [54, 78]. Disclinati<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> dislocati<strong>on</strong>s are important<br />

topological defects that induce l<strong>on</strong>g range disrupti<strong>on</strong>s of orientati<strong>on</strong>al or translati<strong>on</strong>al<br />

order respectively [2]. Disclinati<strong>on</strong>s are points of local 5- <str<strong>on</strong>g>and</str<strong>on</strong>g> 7-fold symmetry in a<br />

triangular lattice (labeled by topological charges q = ± 2π<br />

6<br />

respectively), while dislo-<br />

cati<strong>on</strong>s are disclinati<strong>on</strong> dipoles characterized by a Burgers vector, � b, defined as the<br />

amount by which a circuit drawn around the dislocati<strong>on</strong> fails to close (see Figure<br />

5.1 inset). Other point defects such as vacancies, interstitials or impurity atoms cre-<br />

ate shorter range disturbances that introduce <strong>on</strong>ly local stretching or compressi<strong>on</strong><br />

in the lattice (see Figure 5.4). Such defects are important for particle diffusi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

relaxati<strong>on</strong> of c<strong>on</strong>centrati<strong>on</strong> fluctuati<strong>on</strong>s.


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 114<br />

These particle-like objects interact not <strong>on</strong>ly with each other, but also with<br />

the curvature of the substrate via a <strong>on</strong>e-body geometric potential that depends <strong>on</strong> the<br />

particular type of defect [15, 4]. These geometric potentials are in general n<strong>on</strong> − local<br />

functi<strong>on</strong>s of the Gaussian curvature that we determine explicitly here for a model<br />

surface shaped as a ”Gaussian bump”. An isolated bump of this kind models l<strong>on</strong>g<br />

wavelength undulati<strong>on</strong>s of a lithographic substrate, has regi<strong>on</strong>s of both positive <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

negative curvature, <str<strong>on</strong>g>and</str<strong>on</strong>g> yet is simple enough to allow straightforward analytic <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

numerical calculati<strong>on</strong>s. The presence of these geometric potentials triggers defect<br />

unbinding instabilities in the ground state of the curved space crystal, even if no<br />

topological c<strong>on</strong>straints <strong>on</strong> the net number of defects exist. Geometric potentials also<br />

c<strong>on</strong>trol the dynamics of isolated dislocati<strong>on</strong>s whose moti<strong>on</strong> in the glide directi<strong>on</strong><br />

can be suppressed. Similar mechanisms influence the equilibrium distributi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

dynamics of vacancies, interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> impurity atoms.<br />

5.2 Basic Formalism<br />

The in-plane elastic free energy of a crystal embedded in a gently curved<br />

frozen substrate,given in the M<strong>on</strong>ge form by �r(x, y) = (x, y, h(x, y)), can be expressed<br />

in terms of the Lame coefficients µ <str<strong>on</strong>g>and</str<strong>on</strong>g> λ [14]<br />

F = 1<br />

�<br />

2<br />

�<br />

dA σij(�x)uij(�x) = dA<br />

�<br />

µ u 2 ij(�x) + λ<br />

2 u2 �<br />

kk(�x)<br />

, (5.1)<br />

where �x = {x, y} represents a set of st<str<strong>on</strong>g>and</str<strong>on</strong>g>ard cartesian coordinates in the plane<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> dA = dxdy √ g, where √ g = � 1 + (∂xh) 2 + (∂yh) 2 . In Eq.(5.1) the stress ten-<br />

sor σij(�x) = 2µuij(�x) + λδijukk(�x) was written in terms of the strain tensor uij(�x) =


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 115<br />

1<br />

2 [∂iuj(�x) + ∂jui(�x) + Aij(�x)]. The latter c<strong>on</strong>tains an additi<strong>on</strong>al term Aij(�x) = ∂ih(�x)∂jh(�x)<br />

(compared to its flat space counterpart) that couples the gradient of the displacement<br />

field ui(�x) to the geometry of the substrate as embedded in the gradient of the sub-<br />

strate height functi<strong>on</strong> h(�x). We will often illustrate our results for a single Gaussian<br />

r2<br />

− bump whose height functi<strong>on</strong> is given by h(�x) = α x0e 2 where r ≡ |�x|<br />

x0<br />

si<strong>on</strong>less radial coordinate [58].<br />

is the dimen-<br />

In Eq.(5.1) <str<strong>on</strong>g>and</str<strong>on</strong>g> the rest of this paper, we adopt the ordinary flat space metric<br />

(eg. setting dA ≈ dxdy) <str<strong>on</strong>g>and</str<strong>on</strong>g> absorb all the complicati<strong>on</strong>s associated with the curved<br />

substrate in the tensor field Aij(�x) that resembles the more familiar vector potential of<br />

electromagnetism. Like its electromagnetic analog, the curl of the tensor field Aij(�x)<br />

has a clear physical meaning <str<strong>on</strong>g>and</str<strong>on</strong>g> is equal to the Gaussian curvature of the surface<br />

G(�x) = −ɛilɛjk∂l∂kAij(�x) where ɛij is the antisymmetric unit tensor (ɛxy = −ɛyx = 1)<br />

[15]. The c<strong>on</strong>sistency of our perturbative formalism to leading order in α is assessed<br />

in Appendix H.<br />

Minimizati<strong>on</strong> of the free energy in Eq.(5.1) with respect to the displacements<br />

ui(�x), naturally leads to the force-balance equati<strong>on</strong>, ∂iσij(�x) = 0. If we write σij(�x)<br />

in terms of the Airy stress functi<strong>on</strong> χ(�x)<br />

then the force balance equati<strong>on</strong> is automatically satisfied,<br />

σij(�x) = ɛilɛjk∂l∂kχ(�x) (5.2)<br />

∂iσij = ɛjk∂k[∂1, ∂2]χ(�x) = 0 , (5.3)<br />

since the commutator of partial derivatives is zero. Any scalar functi<strong>on</strong> χ(�x) will<br />

generate a stress tensor field σij(�x) that satisfies the force balance equati<strong>on</strong>. To lift


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 116<br />

this redundancy, <strong>on</strong>e choses χ(�x) <str<strong>on</strong>g>and</str<strong>on</strong>g> hence σij(�x) so that the corresp<strong>on</strong>ding strain<br />

tensor, uij(�x), matches boundary c<strong>on</strong>diti<strong>on</strong>s for the problem under investigati<strong>on</strong>.<br />

The free energy in Eq.(5.1) can be recast (up to boundary terms) as a simple<br />

functi<strong>on</strong>al of the scalar field χ(�x)<br />

F = 1<br />

2Y<br />

�<br />

dA (∆χ(�x)) 2 , (5.4)<br />

where ∆ is the flat space Laplacian <str<strong>on</strong>g>and</str<strong>on</strong>g> Y = 4µ(µ+λ)<br />

2µ+λ is the Young modulus [2]. Up<strong>on</strong><br />

minimizing F in the presence of defects, we obtain a bi-harm<strong>on</strong>ic equati<strong>on</strong> for χ(�x)<br />

whose source is c<strong>on</strong>trolled by the distributi<strong>on</strong> of defects <str<strong>on</strong>g>and</str<strong>on</strong>g> by the varying Gaussian<br />

curvature of the surface G(�x) [2, 15]<br />

1<br />

Y ∆2 χ(�x) = S(�x) − G(�x) . (5.5)<br />

The source S(�x) for a distributi<strong>on</strong> of N unbound disclinati<strong>on</strong>s with “topological<br />

charges” {qβ = ± 2π<br />

6 } <str<strong>on</strong>g>and</str<strong>on</strong>g> M dislocati<strong>on</strong>s with Burger’s vectors {� b β } reads [2]<br />

S(�x) =<br />

N�<br />

qαδ(�x, �x α ) +<br />

α=1<br />

M�<br />

β=1<br />

ɛijb β<br />

i ∂jδ(�x, �x β ) , (5.6)<br />

where | � b| is equal to the lattice c<strong>on</strong>stant a for dislocati<strong>on</strong>s with the smallest Burger’s<br />

vector. For N isotropic vacancies, interstitials or impurities, we have [2]<br />

S(�x) = 1<br />

2<br />

N�<br />

Ωα ∆δ(�x, �x α ) , (5.7)<br />

α=1<br />

where Ωα ∼ a 2 is the local area change caused by including the point defect at positi<strong>on</strong><br />

�x α .<br />

It is c<strong>on</strong>venient to introduce an auxiliary functi<strong>on</strong> V (�x) that satisfies the<br />

Poiss<strong>on</strong> equati<strong>on</strong> ∆V (�x) = G(�x) <str<strong>on</strong>g>and</str<strong>on</strong>g> vanishes at infinity where the surface flattens


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 117<br />

out. In order to determine the geometric potential, ζ(�x α ), of a defect at �x α we<br />

integrate by parts twice in Eq.(C.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> use Eq.(5.5) to obtain ∆ 2 χ(�x) <str<strong>on</strong>g>and</str<strong>on</strong>g> χ(�x) in<br />

terms of the Green’s functi<strong>on</strong> of the biharm<strong>on</strong>ic operator. The geometric potential<br />

(<strong>on</strong> a deformed plane flat at infinity) follows from integrating by parts the cross tems<br />

involving the source <str<strong>on</strong>g>and</str<strong>on</strong>g> the Gaussian curvature with the result<br />

ζ(�x α ) = −Y<br />

�<br />

dA ′ S(�x ′ )<br />

�<br />

dA 1<br />

∆�x�x ′<br />

V (�x) . (5.8)<br />

This formula needs to be supplemented with appropriate finite size correcti<strong>on</strong>s to<br />

account for the presence of a boundary, as discussed in Appendix I.<br />

5.3 Geometric Frustrati<strong>on</strong><br />

We start by calculating the energy of a relaxed defect-free two dimensi<strong>on</strong>al<br />

crystal <strong>on</strong> a quenched topography. In analogy with the bending of thin plates we<br />

expect some stretching to arise as an unavoidable c<strong>on</strong>sequence of the geometric c<strong>on</strong>-<br />

straints associated with the Gaussian curvature [14]. The resulting energy of geomet-<br />

ric frustrati<strong>on</strong>, F0, can be estimated with the aid of Eq.(5.5) which, when S(�x) = 0,<br />

reduces to a Poiss<strong>on</strong> equati<strong>on</strong> whose source is given by V (�x)<br />

1<br />

Y ∆χG (�x) = −V (�x) + HR(�x) . (5.9)<br />

where HR(�x) is an harm<strong>on</strong>ic functi<strong>on</strong> of �x parameterized by the radius of the circular<br />

boundary R. As discussed in Appendix I, HR(�x) vanishes in the limit R ≫ x0 if free<br />

boundary c<strong>on</strong>diti<strong>on</strong>s are chosen. We denote the soluti<strong>on</strong> of Eq.(5.9) as χ G (�x) where<br />

the superscript G indicates that the Airy functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> the corresp<strong>on</strong>ding stress tensor


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 118<br />

σ G ij(�x) describe elastic deformati<strong>on</strong>s caused by the Gaussian curvature G(�x) <strong>on</strong>ly,<br />

without c<strong>on</strong>tributi<strong>on</strong> from the defects. Up<strong>on</strong> using the general definiti<strong>on</strong> of the Airy<br />

functi<strong>on</strong> (Eq. 5.2), we obtain<br />

σ G kk(�x) = ∆χ G (�x) = −Y V (�x) . (5.10)<br />

For a surface with azimuthal symmetry, like the bump, Poiss<strong>on</strong>’s equati<strong>on</strong><br />

can be readily solved up<strong>on</strong> applying Gauss theorem with the Gaussian curvature G(r)<br />

as a source [58]:<br />

� ∞<br />

dr<br />

V (r) = −<br />

r<br />

′ � r ′<br />

dr<br />

r 0<br />

′′ r ′′ G(r ′′ ) = − 1<br />

4 α2e −r2<br />

Up<strong>on</strong> substituting Eq.(5.9) in Eq.(C.2), we have<br />

F0 � Y<br />

2<br />

�<br />

dA V (r) 2 = Y πx2 0α 4<br />

The result in Eq.(5.3) is valid to leading order in α, c<strong>on</strong>sistent with the assumpti<strong>on</strong>s of<br />

our formalism after taking the limit R ≫ x0 (see Appendix B for finite size correcti<strong>on</strong>s<br />

in small systems). For a harm<strong>on</strong>ic lattice, Y = 2<br />

√ 3 k, where k is an effective spring<br />

c<strong>on</strong>stant that can be extracted from more realistic inter-particle potentials [80]. For<br />

colloidal particles, k is typically of the order of a few hundred times kBT/a 2 , where<br />

T is room temperature [10]. Our numerical calculati<strong>on</strong>s of F0 in fixed-c<strong>on</strong>nectivity<br />

harm<strong>on</strong>ic solids are in good agreement with the small α expansi<strong>on</strong> in Eq.(5.3) as l<strong>on</strong>g<br />

as the aspect ratio α is around 1/2 or lower (see Appendix H for a numerical plot<br />

of F0 versus α). An immediate implicati<strong>on</strong> of the geometric frustrati<strong>on</strong> embodied in<br />

Eq. (5.3) <str<strong>on</strong>g>and</str<strong>on</strong>g> (5.3) is that nucleati<strong>on</strong> of crystal domains <strong>on</strong> the bump will take place<br />

preferentially away from the top in regi<strong>on</strong>s where the surface flattens out.<br />

64<br />

.<br />

.


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 119<br />

5.4 Geometric potential for dislocati<strong>on</strong>s<br />

The energy of a two dimensi<strong>on</strong>al curved crystal with defects will include the<br />

frustrati<strong>on</strong> energy, the inter-defect interacti<strong>on</strong>s (to leading order these are unchanged<br />

from their flat space form, see Appendix H), possible core energies <str<strong>on</strong>g>and</str<strong>on</strong>g> a character-<br />

istic, <strong>on</strong>e-body potential of purely geometrical origin that describes the coupling of<br />

the defects to the curvature given by Eq.(5.8). The geometric potential of an isolated<br />

dislocati<strong>on</strong>, ζ(�x) ≡ D(�x, θ), is a functi<strong>on</strong> of its positi<strong>on</strong> as well as of the angle θ that<br />

the Burger vector � b forms with respect to the radial directi<strong>on</strong> (in the tangent plane<br />

of the surface). Up<strong>on</strong> setting all qβ = 0 in Eq.(5.6) <str<strong>on</strong>g>and</str<strong>on</strong>g> substituting it into Eq.(5.8),<br />

we obtain, for an isolated dislocati<strong>on</strong>, the resulting functi<strong>on</strong> D(r = |�x|<br />

, θ)<br />

�<br />

D(r, θ) = Y bi ɛij ∂j dA ′<br />

�<br />

1<br />

V (�x<br />

∆�x x �′ ′ ) + α2x2 0<br />

4R2 �<br />

α<br />

≈ Y b x0<br />

2<br />

��<br />

e<br />

sin θ<br />

8 −r2<br />

� �<br />

− 1<br />

� �<br />

x0<br />

2<br />

+ r<br />

r R<br />

x0<br />

. (5.11)<br />

In view of the azimuthal symmetry of the surface, Gauss’ theorem as expressed in<br />

Eq.(5.3), was employed in deriving the sec<strong>on</strong>d equality in Eq.(5.11) which is functi<strong>on</strong><br />

<strong>on</strong>ly of the dimensi<strong>on</strong>less radial coordinate r. The first term in Eq.(5.11) corresp<strong>on</strong>ds<br />

to the infinite plane geometric potential obtained from Eq.(5.8) while the sec<strong>on</strong>d term<br />

is a finite size correcti<strong>on</strong> arising from a circular boundary of radius R (see Appendix<br />

I for a detailed derivati<strong>on</strong>). Eq.(5.11) is valid to leading order in perturbati<strong>on</strong> theory,<br />

c<strong>on</strong>sistent with the small α approximati<strong>on</strong> adopted in this work (see Appendix H).<br />

In Fig. 5.1 we present a detailed comparis<strong>on</strong> between the theoretical pre-<br />

dicti<strong>on</strong>s for the geometric potential D(r,θ)<br />

Y bx0<br />

plotted versus r as c<strong>on</strong>tinuous lines <str<strong>on</strong>g>and</str<strong>on</strong>g>


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 120<br />

numerical data from c<strong>on</strong>strained minimizati<strong>on</strong> of an harm<strong>on</strong>ic solid <strong>on</strong> a bump with<br />

α = 0.5, under c<strong>on</strong>diti<strong>on</strong>s such that R ≫ x0 ≫ a. (See Appendix J for a discussi<strong>on</strong><br />

of our numerical approach). The lower <str<strong>on</strong>g>and</str<strong>on</strong>g> upper branches of the graph are obtained<br />

from Eq. (5.11) by setting θ = ± π<br />

2<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> letting R<br />

x0<br />

equal to 4 (blue curve) or 8 (red<br />

curve). Indeed, the (scaled) data from simulati<strong>on</strong>s with different choices of x0<br />

b (see<br />

capti<strong>on</strong>) collapse <strong>on</strong> the two master-curves according to their ratio of R . Two dif-<br />

x0<br />

ferent curves arise because the dislocati<strong>on</strong> interacts with the curvature directly <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

via its image. The image-mediated interacti<strong>on</strong> is given by the R dependent term in<br />

Eq.(5.11). The simple dependence of D(r, θ) <strong>on</strong> the directi<strong>on</strong> of the Burgers vector is<br />

revealed by Fig. 5.1, since the upper branch of the graph corresp<strong>on</strong>ding to the unsta-<br />

ble equilibrium θ = − π<br />

2<br />

to θ = π<br />

2 .<br />

is approximately symmetric to the lower <strong>on</strong>e corresp<strong>on</strong>ding<br />

The analogy between the geometrical potential of the dislocati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> the<br />

more familiar interacti<strong>on</strong> of an electric dipole in an external field can be elucidated<br />

by regarding the dislocati<strong>on</strong> as a charge neutral pair of disclinati<strong>on</strong>s whose dipole<br />

moment qdi = ɛijbj is a lattice vector perpendicular to � b that c<strong>on</strong>nects the two<br />

points of 5 <str<strong>on</strong>g>and</str<strong>on</strong>g> 7-fold symmetry. The geometric potential, U(r), of a disclinati<strong>on</strong><br />

of topological charge q interacting with the Gaussian curvature satisfies the Poiss<strong>on</strong><br />

equati<strong>on</strong> ∆U(r) = −qV (r) as can be seen by substituting the source in Eq.(5.6) with<br />

all � b α = 0 into Eq.(5.8) . For small r, positive (negative) disclinati<strong>on</strong>s are attracted<br />

(repelled) from the center of the bump by the integrated background source V (r)<br />

which increases for r ≤ 1 like α 2 r 2 <str<strong>on</strong>g>and</str<strong>on</strong>g> is multiplied by the 2D electric field 1<br />

r ,


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 121<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

−0.005<br />

−0.01<br />

−0.015<br />

−0.02<br />

D(r,θ)<br />

Ybr 0<br />

0 0.5 1 1.5 2 2.5 3 3.5<br />

Figure 5.1: The dislocati<strong>on</strong> potential D(x, θ = ± π)<br />

in Eq.(5.11) (including finite size<br />

2<br />

correcti<strong>on</strong>s) is plotted as a c<strong>on</strong>tinuous line for a Gaussian bump parameterized by α =<br />

0.5 in the limit R >> x0 >> a. Open symbols represent the numerical minimizati<strong>on</strong><br />

of a fixed c<strong>on</strong>nectivity harm<strong>on</strong>ic model for which the separati<strong>on</strong> of the +/- disclinati<strong>on</strong><br />

pair representing the dislocati<strong>on</strong> is fixed while allowing the dislocati<strong>on</strong> as a whole to<br />

move radially with respect to the bump. Lower branch: θ = π/2, R/x0 =4 (blue),<br />

8 (red); <str<strong>on</strong>g>and</str<strong>on</strong>g> x0/a = 10 (○), 20 (△), 40 (✷). Upper branch: θ = −π/2, x0/R = 8,<br />

x0/a = 10. Inset: A schematic view of a dislocati<strong>on</strong> with filled symbols representing<br />

six- (circles), five- (diam<strong>on</strong>d) <str<strong>on</strong>g>and</str<strong>on</strong>g> seven- (square) coordinate particles. Also depicted<br />

are the two rows of extra atoms emanating from the five-coordinated particle. The<br />

Burger’s vector is shown as a red arrow, completing the circuit around the dislocati<strong>on</strong><br />

(dashed line).<br />

r


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 122<br />

resulting in a geometric force that increases linearly in r. If the positive disclinati<strong>on</strong><br />

within the dipole is closer to the top, the force it experiences will be opposite <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

slightly less than the <strong>on</strong>e experienced by the negative disclinati<strong>on</strong> that is further away<br />

from the top. As a result a ”tidal” force will push the dislocati<strong>on</strong> (as a whole) down<br />

hill as shown by the lower branch of the plot of Fig. 5.1. For large r, however, the<br />

source V (r) saturates <str<strong>on</strong>g>and</str<strong>on</strong>g> the attractive force exerted <strong>on</strong> the positive disclinati<strong>on</strong><br />

wins <str<strong>on</strong>g>and</str<strong>on</strong>g> drags the dislocati<strong>on</strong> towards the bump. The minimum of the geometric<br />

potential occurs at |�xmin| ≈ 1.1 x0, close to the circle of zero Gaussian curvature<br />

|�x| = x0, as a result of the competing interacti<strong>on</strong>s of the two disclinati<strong>on</strong>s comprising<br />

the dislocati<strong>on</strong>. If the orientati<strong>on</strong> of the disclinati<strong>on</strong> dipole is flipped, the geometric<br />

potential flips sign. Similarly, if the sign of the Gaussian curvature is reversed, that<br />

is to say the bump turns into a saddle, the sign of the geometric interacti<strong>on</strong> flips. As<br />

a result, a dislocati<strong>on</strong> close to a saddle will have its Burger vector oriented so that<br />

the closest disclinati<strong>on</strong> to the saddle is 7-fold coordinated.<br />

The physical origin of the dislocati<strong>on</strong> potential can be understood heuris-<br />

tically without explicit recourse to our source formalism. According to st<str<strong>on</strong>g>and</str<strong>on</strong>g>ard<br />

elasticity theory, a dislocati<strong>on</strong> in an external stress field σij(�x) experiences a Peach<br />

Kohler force, � f(�x), given by fk(�x) = ɛkjbiσij(�x) [47]. Similarly, a dislocati<strong>on</strong> intro-<br />

duced into the curved 2D crystal will experience a Peach Kohler force as a result of<br />

the pre-existing stress field of geometric frustrati<strong>on</strong> σ G ij( �x α ) whose n<strong>on</strong>-diag<strong>on</strong>al com-<br />

p<strong>on</strong>ents vanish. This interpretati<strong>on</strong> is c<strong>on</strong>sistent with the geometric potential derived<br />

in Eq.(5.11), provided we use Eq. (5.9) to write D(�x) = biɛij∂jχ G . With � b al<strong>on</strong>g its


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 123<br />

minimum orientati<strong>on</strong> (azimuthal counter-clockwise), we obtain a radial Peach Kohler<br />

force of magnitude f(r) = −b σG ∂D(r)<br />

φφ (r) that matches −<br />

∂r = −b ∂2χG (r)<br />

∂r2 .<br />

5.5 Dislocati<strong>on</strong> unbinding <str<strong>on</strong>g>and</str<strong>on</strong>g> Grain Boundaries<br />

If the 2D crystal is grown <strong>on</strong> a substrate which is sufficiently deformed, the<br />

resulting elastic strain can be partially relaxed by introducing unbound dislocati<strong>on</strong>s<br />

into the ground state [15, 58]. Here we present a simple estimate of the threshold<br />

aspect ratio, αc, necessary to trigger this instability. Boundary effects will be ignored<br />

in what follows by letting R → ∞ in Eq.(5.11).<br />

C<strong>on</strong>sider two dislocati<strong>on</strong>s located at �x1 <str<strong>on</strong>g>and</str<strong>on</strong>g> �x2 a distance of approximately<br />

x0 from the center of the bump <strong>on</strong> opposite sides (see the inset of Fig. 5.2). Their<br />

disclinati<strong>on</strong>-dipole moments are opposite to each other <str<strong>on</strong>g>and</str<strong>on</strong>g> aligned in the radial direc-<br />

ti<strong>on</strong> so that the two (antiparallel) Burger vectors are perpendicular to the separati<strong>on</strong><br />

vector in the plane �x12 ≡ �x1 − �x2. In this case, the interacti<strong>on</strong> between the disloca-<br />

�<br />

[2]. The instability occurs when the energy gain<br />

ti<strong>on</strong>s reduces to V12 ≈ Y<br />

4π b2 �<br />

|�x12|<br />

ln a<br />

from placing each dislocati<strong>on</strong> in the minima of the potential D(�x) given by Eq.(5.11)<br />

outweights the sum of the work needed to tear them apart plus the core energies 2Ec.<br />

The critical aspect ratio at the threshold which results is given by<br />

α 2 c ≈ c b<br />

x0<br />

ln<br />

�<br />

x0<br />

b ′<br />

�<br />

, (5.12)<br />

where b ′ ≡ b<br />

8πEc<br />

e− Y b<br />

2 2 is the magnitude of the rescaled Burger vector <str<strong>on</strong>g>and</str<strong>on</strong>g> c ≈ 1/2.<br />

In Fig. 5.2a we present a comparis<strong>on</strong> between Eq. (5.12) <str<strong>on</strong>g>and</str<strong>on</strong>g> numerical re-<br />

sults. For each value of x0<br />

b , the corresp<strong>on</strong>ding αc is obtained numerically by comparing


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 124<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

α C<br />

r 0 /b<br />

0<br />

10 12 14 16 18 20 22 24<br />

(a)<br />

Figure 5.2: (a) Dislocati<strong>on</strong> unbinding critical aspect ratio, αc, as a functi<strong>on</strong> of r0<br />

b .<br />

The theoretical estimate (5.12) is plotted vs. x0 for core energies Ec = 0 (dashed)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> Ec = 0.1Y b 2 (solid). Numerical values (circles) were obtained as described in<br />

Secti<strong>on</strong> 5.5 <str<strong>on</strong>g>and</str<strong>on</strong>g> Appendix J. Inset: particle c<strong>on</strong>figurati<strong>on</strong> projected in the plane<br />

for a dislocati<strong>on</strong> pair straddling the bump. The extra atoms associated with the<br />

dislocati<strong>on</strong>s are highlighted with black lines. (b) Log of the numerical strain energy<br />

density <strong>on</strong> a bump for (b) the c<strong>on</strong>figurati<strong>on</strong> shown in the inset. (c) The same quantity<br />

for a defect-free bump. Both plots were c<strong>on</strong>structed numerically with x0 = 10, α =<br />

0.7 > αc. Red represents high strain.<br />

the energy of a lattice without defects to the c<strong>on</strong>figurati<strong>on</strong> with the two dislocati<strong>on</strong>s<br />

in their equilibrium positi<strong>on</strong>s. This interpretati<strong>on</strong> for the origin of the instability is<br />

corroborated by the (numerical) strain energy density plots of Fig. 5.2b <str<strong>on</strong>g>and</str<strong>on</strong>g> 5.2c ,<br />

where it is shown that introducing the pair of dislocati<strong>on</strong>s reduces the strain energy<br />

density <strong>on</strong> top of the bump at the price of creating some large, but localized strains<br />

around the dislocati<strong>on</strong> cores where uij diverges. In the c<strong>on</strong>tinuum limit b ≪ x0,<br />

very small deformati<strong>on</strong>s are enough to trigger the instability. This is the regime in<br />

which our perturbati<strong>on</strong> treatment applies. As α is increased even further a cascade of<br />

dislocati<strong>on</strong> unbinding transiti<strong>on</strong>s occurs involving larger numbers of dislocati<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

more complicated equilibrium arrangements of zero net Burger vector. For sufficiently<br />

(b)<br />

(c)


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 125<br />

large aspect ratios, we expect that the dislocati<strong>on</strong>s tend to line up in grain boundary<br />

scars similar to the <strong>on</strong>es observed in spherical crystals [3]. This scenario is c<strong>on</strong>sistent<br />

with preliminary results from M<strong>on</strong>te Carlo simulati<strong>on</strong>s in which the fixed-c<strong>on</strong>nectivity<br />

c<strong>on</strong>straint is lifted <str<strong>on</strong>g>and</str<strong>on</strong>g> more complicated surface morphologies are c<strong>on</strong>sidered [18].<br />

5.6 Glide suppressi<strong>on</strong><br />

The dynamics of dislocati<strong>on</strong>s proceeds by means of two distinct processes:<br />

glide <str<strong>on</strong>g>and</str<strong>on</strong>g> climb. Glide describes moti<strong>on</strong> al<strong>on</strong>g the directi<strong>on</strong> defined by the Burgers<br />

vector; in flat space glide requires a very low activati<strong>on</strong> energy <str<strong>on</strong>g>and</str<strong>on</strong>g> is the dominant<br />

form of moti<strong>on</strong> at low temperature, see Fig. 5.3a inset. Climb, or moti<strong>on</strong> perpendicu-<br />

lar to the Burger vector requires diffusi<strong>on</strong> of vacancies <str<strong>on</strong>g>and</str<strong>on</strong>g> interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> is usually<br />

frozen out relative to glide which involves <strong>on</strong>ly local rearrangements of atoms. On a<br />

curved surface, the geometric potential D(r, θ) imposes c<strong>on</strong>straints <strong>on</strong> the glide dy-<br />

namics of isolated dislocati<strong>on</strong>s, in sharp c<strong>on</strong>trast to flat space where the <strong>on</strong>ly energy<br />

barriers present are due to the periodic Peierls potential [47].<br />

As the dislocati<strong>on</strong> represented in the inset of Fig. 5.3a moves in the glide di-<br />

recti<strong>on</strong>, it experiences a restoring potential generated by the variati<strong>on</strong> of the (scaled)<br />

radial distance <str<strong>on</strong>g>and</str<strong>on</strong>g> the deviati<strong>on</strong> from the radial alignment of the dislocati<strong>on</strong> dipole.<br />

For a small transverse displacement, y, the harm<strong>on</strong>ic potential U(y) = 1<br />

2 kdy 2 is c<strong>on</strong>-<br />

trolled by a radial, positi<strong>on</strong>-dependent effective spring c<strong>on</strong>stant, kd which depends <strong>on</strong><br />

the radial coordinate r. Up<strong>on</strong> exp<str<strong>on</strong>g>and</str<strong>on</strong>g>ing Eq.(5.11) to leading order in y <str<strong>on</strong>g>and</str<strong>on</strong>g> α, we


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 126<br />

20<br />

15<br />

10<br />

5<br />

0<br />

E(y)<br />

Ybr 0<br />

−4 −3 −2 −1 0 1 2 3 4<br />

(a)<br />

y<br />

2<br />

1.5<br />

1<br />

0.5<br />

k d (r)r 0<br />

Yb<br />

0<br />

0 0.5 1 1.5 2 2.5<br />

Figure 5.3: Dislocati<strong>on</strong> glide. (a) Filled circles represent numerical glide energies<br />

vs. y (units of a) for x0/a = 10, R/x0 = 8, r = 0.5 (at y = 0): α = 0.1 (dark<br />

blue), 0.3 (blue), 0.5 (orange) <str<strong>on</strong>g>and</str<strong>on</strong>g> 0.7 (red). Solid lines represent E(y) = 1<br />

2 kdy 2 , with<br />

kd determined from Eq. (5.13). The energy is scaled by 10 −4 . Inset: schematic of<br />

a dislocati<strong>on</strong> (red) <strong>on</strong> a Gaussian bump. The glide path is highlighted in orange.<br />

(b) The curvature-induced glide-suppressi<strong>on</strong> spring c<strong>on</strong>stant kd(r) (5.13) (solid line)<br />

is plotted versus scaled in-plane distance from the top of the bump r = x/x0 for<br />

x0 = 10a, α = 0.5. Open symbols represent numerical results found by fitting similar<br />

data as in (a) to parabolic curves in the glide coordinate. The energy is scaled by<br />

10 −2 .<br />

obtain<br />

kd(r) ≈ α2 Y b<br />

4 x0<br />

�<br />

1 − (1 + r2 )e−r2 �<br />

. (5.13)<br />

The harm<strong>on</strong>ic potential 1<br />

2 kdy 2 is shown in Fig. 5.3a where the data obtained from<br />

numerical minimizati<strong>on</strong> of the harm<strong>on</strong>ic lattice is explicitly compared to the predic-<br />

ti<strong>on</strong> of Eq.(5.13) for different values of the aspect ratio. Note that the effective spring<br />

c<strong>on</strong>stant plotted in Fig. 5.3b vanishes in the limit b<br />

x0<br />

r 3<br />

(b)<br />

→ 0 but can still be important<br />

for small systems since Y b 2 is of the order of hundreds of kBT ≡ β −1 (see secti<strong>on</strong><br />

5.3). The c<strong>on</strong>fining potential plotted in Fig. 5.3a is similar to the <strong>on</strong>e experienced<br />

by a dislocati<strong>on</strong> bound to a disclinati<strong>on</strong> [81, 10].<br />

r


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 127<br />

The resulting thermal moti<strong>on</strong> in the glide directi<strong>on</strong> of dislocati<strong>on</strong>s in this<br />

binding harm<strong>on</strong>ic potential is modeled by an over-damped Langevin equati<strong>on</strong> for the<br />

glide coordinate, y. We have 〈∆y 2 〉 = 1−e−2µk d t<br />

βkd<br />

[81, 10]. In the case of a bump<br />

geometry, the effective spring c<strong>on</strong>stant kd can be evaluated using Eq.(5.13). We<br />

emphasize, however, that the glide suppressi<strong>on</strong> mechanism c<strong>on</strong>sidered here is not<br />

caused by the interacti<strong>on</strong> of the dislocati<strong>on</strong> with other defects but purely by the<br />

geometric interacti<strong>on</strong> with the curvature of the substrate.<br />

5.7 Vacancies, Interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> Impurities<br />

We now turn to a derivati<strong>on</strong> of the geometric potential, I(�x α ), for inter-<br />

stitials, isotropic vacancies <str<strong>on</strong>g>and</str<strong>on</strong>g> impurities. Inspecti<strong>on</strong> of Fig. 5.4 reveals that an<br />

interstitial (vacancy) can be viewed either as the product of locally adding (remov-<br />

ing) an atom to the lattice or as a composite object made up of three disclinati<strong>on</strong><br />

dipoles. In order to derive I(�x α ), we substitute the source term of Eq.(5.7) into<br />

Eq.(5.8) <str<strong>on</strong>g>and</str<strong>on</strong>g> integrate by parts twice. The result reads<br />

I(�x α ) ≈ Y<br />

2 Ωα V (�x α ) , (5.14)<br />

where boundary terms <str<strong>on</strong>g>and</str<strong>on</strong>g> a positi<strong>on</strong> independent nucleati<strong>on</strong> energy have been dropped.<br />

The c<strong>on</strong>stant Ω α represents the area excess or deficit associated with the defect. In<br />

Fig. 5.4 a comparis<strong>on</strong> of Eq.(5.14) with the results obtained from mapping out I(r α )<br />

numerically is presented. The area changes Ωi <str<strong>on</strong>g>and</str<strong>on</strong>g> Ωv for interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> vacancies<br />

respectively were fit to the numerical data. We find that Ωv is negative <str<strong>on</strong>g>and</str<strong>on</strong>g> greater in<br />

magnitude than Ωi. The large r behavior of I(r) indicates that the core energy of a


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 128<br />

4.5<br />

4.45<br />

4.4<br />

4.35<br />

4.3<br />

4.25<br />

4.2<br />

I(r)<br />

Ya 2<br />

4.15<br />

0 0.5 1 1.5 2 2.5<br />

Figure 5.4: The scaled potential energy I(x)/Ya 2 of interstitials (bottom) <str<strong>on</strong>g>and</str<strong>on</strong>g> vacancies<br />

(top) is plotted versus the scaled distance r for the same bump as Fig. 5.1.<br />

Solid lines are obtained from Eq. (5.14), where the area changes of interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

vacancies Ω α i <str<strong>on</strong>g>and</str<strong>on</strong>g> Ω α v were fit to the data. Sample c<strong>on</strong>figurati<strong>on</strong>s are plotted in the<br />

insets, where diam<strong>on</strong>ds, circles <str<strong>on</strong>g>and</str<strong>on</strong>g> squares represent five, six <str<strong>on</strong>g>and</str<strong>on</strong>g> seven fold coordinated<br />

particles respectively. Filled <str<strong>on</strong>g>and</str<strong>on</strong>g> unfilled circles in the interstitial plot represent<br />

two distinct but energetically equivalent orientati<strong>on</strong>s of the disclinati<strong>on</strong> dipoles comprising<br />

the interstitial. They differ by swapping the 5 <str<strong>on</strong>g>and</str<strong>on</strong>g> 7-fold disclinati<strong>on</strong>s, which<br />

rotates the orientati<strong>on</strong> of the filled, triangular plaquette by π<br />

3 .<br />

vacancy in flat space is greater than the <strong>on</strong>e of an interstitial for the harm<strong>on</strong>ic lattice.<br />

Interstitials tend to seek the top of the bump whereas vacancies are pushed in the flat<br />

space regi<strong>on</strong>s. From the definiti<strong>on</strong> of V (r) in Eq.(5.3), we deduce that an interstitial<br />

(vacancy) is attracted (repelled) by regi<strong>on</strong>s of positive (negative) Gaussian curva-<br />

ture similar to the behavior of an electrostatic charge interacting with a background<br />

charge distributi<strong>on</strong> given by −G(r). The functi<strong>on</strong> V (r) c<strong>on</strong>trols the curvature-defect<br />

r


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 129<br />

interacti<strong>on</strong> for other types of defects that can be modeled as a Coulomb gas in curved<br />

space such as disclinati<strong>on</strong>s in liquid crystals <str<strong>on</strong>g>and</str<strong>on</strong>g> vortices in 4 He films [58, 43]. The<br />

expressi<strong>on</strong> for V (r) in Eq.(5.3) reveals that I(r) is indeed a n<strong>on</strong> local functi<strong>on</strong> of<br />

the Gaussian curvature determined as in electrostatics by the applicati<strong>on</strong> of Gauss<br />

law. Thus, the vacancy potential <strong>on</strong> a bump does not reach a minimum at the point<br />

where the Gaussian curvature is maximally negative but rather at infinity where the<br />

integrated Gaussian curvature vanishes.<br />

We now argue heuristically how a localized point defect can couple n<strong>on</strong>-<br />

locally to the curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> in the process we provide an alternative justificati<strong>on</strong> of<br />

Eq. (5.14) analogous to the informal derivati<strong>on</strong> of the dislocati<strong>on</strong> potential. The<br />

energy cost, I(�x α ), of a local compressi<strong>on</strong> or stretching in the presence of an ar-<br />

bitrary elastic stress tensor σij(�x) is given by I(�x α ) = p(�x α )δV where δV is the<br />

local volume change <str<strong>on</strong>g>and</str<strong>on</strong>g> p( �x α ) is the local pressure related to the stress tensor via<br />

σij(�x α ) = −p(�x α )δij [14]. In two dimensi<strong>on</strong>s, we have I(�x α ) = −σkk(�x α ) Ωα . We re-<br />

2<br />

cover the result in Eq.(5.14), by assuming that the local deformati<strong>on</strong> Ω α (induced by<br />

the nucleati<strong>on</strong> of a point defect) couples to the preexisting stress of geometric frus-<br />

trati<strong>on</strong> σ G kk<br />

= Y V (�x) (see Eq.(5.10)), which is a n<strong>on</strong>-local functi<strong>on</strong> of the Gaussian<br />

curvature. Elastic deformati<strong>on</strong>s created by the geometric c<strong>on</strong>straint throughout the<br />

curved two dimensi<strong>on</strong>al solid are propagated to the positi<strong>on</strong> of the point defect by<br />

force chains spanning the entire system. The point defect can then be viewed as a<br />

local probe of the stress field that does not measure the additi<strong>on</strong>al stresses induced<br />

by its own presence.


Chapter 5: Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> crystalline order <strong>on</strong> curved surfaces. 130<br />

Note that the geometrical potential of an isotropic point defect is unchanged<br />

if we swap the 5 <str<strong>on</strong>g>and</str<strong>on</strong>g> 7-fold disclinati<strong>on</strong>s comprising it (corresp<strong>on</strong>ding to a rotati<strong>on</strong> of<br />

the point defect by π<br />

3<br />

around its center), as dem<strong>on</strong>strated for interstitials in the lower<br />

branch of Fig. 5.4. C<strong>on</strong>trast this situati<strong>on</strong> with the clear dipolar character of the<br />

dislocati<strong>on</strong> potential plotted in Fig. 5.1. The more complicated case of n<strong>on</strong>-isotropic<br />

point defects appears to still be captured qualitatively by Eq.(5.14); this was explicitly<br />

checked by plotting the geometric potential of ”crushed vacancies” [82], which have<br />

both a lower symmetry <str<strong>on</strong>g>and</str<strong>on</strong>g> a lower energy than their isotropic counterparts.<br />

An arbitrary c<strong>on</strong>figurati<strong>on</strong> of weakly interacting point defects will relax to<br />

its equilibrium distributi<strong>on</strong> by diffusive moti<strong>on</strong> in a force field f(r)= −∇I(r). This<br />

geometric force leads to a biased diffusi<strong>on</strong> dynamics with drift velocity v ∼ β|f|a D<br />

a ∼<br />

DY βΩ<br />

x0<br />

, where β = 1<br />

kBT<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> D is the defect diffusivity [47]. Eventually, a dilute gas of<br />

point defects equilibrates to a n<strong>on</strong>-unform spatial density proporti<strong>on</strong>al to e −βI(r) .


Appendix A<br />

Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Isothermal<br />

Coordinates<br />

131


Appendix A: Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Isothermal Coordinates 132<br />

C<strong>on</strong>formal plane<br />

N<br />

S<br />

Figure A.1: Graphic c<strong>on</strong>structi<strong>on</strong> of the stereographic projecti<strong>on</strong>. Regi<strong>on</strong>s close to<br />

the north pole have larger images in the c<strong>on</strong>formal plane than regi<strong>on</strong>s of equal areas<br />

close to the south pole. The stereographic projecti<strong>on</strong> preserves the topology of the<br />

surface provided all points at infinity are identified with the north pole.<br />

The analysis of ordered phases <strong>on</strong> curved substrates can be simplified by<br />

rewriting the original metric of the surface gab(u) in terms of a c<strong>on</strong>venient set of<br />

coordinates x(u) = (x(u), y(u)) such that the new metric ˜gab(x) reads<br />

˜gab(x) = e ρ(x,y) δab . (A.1)<br />

The metric ˜gab differs from the flat space <strong>on</strong>e δab <strong>on</strong>ly by a c<strong>on</strong>formal factor e ρ(x,y)<br />

that embodies informati<strong>on</strong> <strong>on</strong> the curvature of the surface [44]. These isothermal<br />

coordinates can be used to map arbitrary corrugated surfaces <strong>on</strong>to the plane [83]. The<br />

mapping is c<strong>on</strong>formal so angles are left unchanged but areas are stretched according to<br />

the positi<strong>on</strong>-dependent c<strong>on</strong>formal factor e ρ(x,y) . A familiar example is provided by the<br />

stereographic projecti<strong>on</strong> that maps a sphere <strong>on</strong>to the c<strong>on</strong>formal plane as illustrated<br />

in Fig. A.1. The Green’s functi<strong>on</strong> assumes a very simple form after a c<strong>on</strong>formal<br />

transformati<strong>on</strong>, because the Laplace operator reduces to the familiar flat space result<br />

when expressed in terms of isothermal coordinates. In what follows, we dem<strong>on</strong>strate<br />

that this transformati<strong>on</strong> provides the basis for an efficient strategy to determine the


Appendix A: Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Isothermal Coordinates 133<br />

Green’s functi<strong>on</strong> <strong>on</strong> a bumpy substrate. We start by deriving the radial change of<br />

coordinates ℜ(r) that transforms the original metric of the Gaussian bump, i.e.,<br />

ds 2 �<br />

= 1 + α2r2 r2 e<br />

o<br />

r2<br />

−<br />

r2 o<br />

into the locally flat metric (in polar coordinates),<br />

�<br />

dr 2 + r 2 dφ 2 , (A.2)<br />

ds 2 = e ρ(r) (dℜ 2 + ℜ 2 dφ 2 ) , (A.3)<br />

where ρ(r) <str<strong>on</strong>g>and</str<strong>on</strong>g> ℜ(r) are independent of the azimuthal coordinate φ because of cylin-<br />

drical symmetry. This metric is equivalent to ˜gab(x) up<strong>on</strong> switching from cartesian<br />

(x,y) to polar coordinates (ℜ(r), φ). To simplify the notati<strong>on</strong> we introduce the α-<br />

dependent functi<strong>on</strong> l(r) defined by<br />

l(r) ≡ 1 + α2 r 2<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> plotted in Fig. 2.2 for different choices of α.<br />

r 2 0<br />

�<br />

exp − r2<br />

r2 �<br />

0<br />

, (A.4)<br />

The equivalence of the metrics in Equati<strong>on</strong>s (A.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> (A.3) requires that<br />

ℜ(r) satisfies the differential equati<strong>on</strong><br />

dℜ<br />

ℜ =<br />

The c<strong>on</strong>formal factor is thus given by<br />

The soluti<strong>on</strong> of Eq.(A.5) is<br />

ℜ(r) = A r e − R c<br />

r<br />

�<br />

l(r)<br />

dr . (A.5)<br />

r<br />

e ρ(r) = ( r<br />

ℜ )2 . (A.6)<br />

dr ′<br />

r ′ ( √ l(r ′ ) −1) , (A.7)


Appendix A: Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Isothermal Coordinates 134<br />

where it is c<strong>on</strong>venient to set the arbitrary c<strong>on</strong>stants A <str<strong>on</strong>g>and</str<strong>on</strong>g> c to unity <str<strong>on</strong>g>and</str<strong>on</strong>g> infinity<br />

respectively. This n<strong>on</strong>-linear stretch of the radial coordinate leaves the origin <str<strong>on</strong>g>and</str<strong>on</strong>g> the<br />

point at infinity invariant <str<strong>on</strong>g>and</str<strong>on</strong>g> can be c<strong>on</strong>cisely written as<br />

where the functi<strong>on</strong> V (r) defined by<br />

ℜ(r) = r e V (r) , (A.8)<br />

� ∞<br />

V (r) ≡ − dr ′<br />

r<br />

� l(r ′ ) − 1<br />

r ′ , (A.9)<br />

plays an important role in our formalism <str<strong>on</strong>g>and</str<strong>on</strong>g> its interpretati<strong>on</strong> as a sort of geometric<br />

potential is explored in detail in Appendix B.<br />

The Poiss<strong>on</strong> equati<strong>on</strong> for the Green’s functi<strong>on</strong> Γ(u, u ′ ) <strong>on</strong> a surface with<br />

metric tensor gαβ <str<strong>on</strong>g>and</str<strong>on</strong>g> point source δ(u, u ′ ) reads [40]:<br />

D α DαΓ(u, u ′ ) = − δ(u, u′ )<br />

√ g<br />

where the covariant Laplacian is given for general coordinates by:<br />

, (A.10)<br />

D α Dα ≡ (1/ √ g)∂α[ √ gg αβ ∂β] . (A.11)<br />

The c<strong>on</strong>formal change of coordinates transforms � g(r, φ) into e ρ(r)� g(ℜ, φ) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

g αβ (r, φ) into e −ρ(r) g αβ (ℜ, φ). The factors of e ρ(r) inside the square brackets in<br />

Eq.(A.11) then cancel <str<strong>on</strong>g>and</str<strong>on</strong>g> we are left with the flat space Laplacian in the polar<br />

coordinates (ℜ(r), φ). We c<strong>on</strong>clude that Γ(u, u ′ ) is simply the Green’s functi<strong>on</strong> of an<br />

undeformed plane expressed in terms of the polar coordinates (ℜ(r), φ):<br />

Γ(u, u ′ ) = − 1<br />

4π ln[ℜ(r)2 + ℜ(r ′ ) 2 − 2ℜ(r)ℜ(r ′ ) cos(φ − φ ′ )] , (A.12)


Appendix A: Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Isothermal Coordinates 135<br />

where an arbitrary additive c<strong>on</strong>stant C (which can be used to satisfy boundary c<strong>on</strong>-<br />

diti<strong>on</strong>s at infinity) has been dropped. We note that Γ(u, u ′ ) differs from the flat<br />

space Green’s functi<strong>on</strong> by a n<strong>on</strong> linear stretch of the radial coordinate. In Appendix<br />

C, we will use Γ(u, u ′ ) to solve Poiss<strong>on</strong>’s equati<strong>on</strong> <strong>on</strong> an infinite bumpy domain <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

calculate the energy stored in the field. As in flat space, the Green’s functi<strong>on</strong> Γ(u, u ′ )<br />

will be suitably modified in a finite system in a way that depends <strong>on</strong> the boundary<br />

c<strong>on</strong>diti<strong>on</strong>s chosen at the edge of the sample (see Appendix D).<br />

We c<strong>on</strong>clude this appendix by evaluating Γ(u, u ′ ) when the two points u <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

u ′ are assumed to be separated by a fixed distance a small enough so that the surface<br />

can be approximated by the local tangent plane to the Gaussian bump. This short<br />

distance behavior will be useful when evaluating the effect of a c<strong>on</strong>stant core radius<br />

<strong>on</strong> the energetics of a disclinati<strong>on</strong> at an arbitrary positi<strong>on</strong>. The fixed microscopic<br />

length a <strong>on</strong> the bump is stretched when projected in the c<strong>on</strong>formal plane (see, e.g.,<br />

Fig. A.1) <str<strong>on</strong>g>and</str<strong>on</strong>g> assumes the positi<strong>on</strong> dependent value λ(x, y) given by<br />

ρ(x,y)<br />

−<br />

λ(x, y) = ae 2 . (A.13)<br />

For a Gaussian bump cylindrical symmetry requires that λ(r, φ) is dependent <strong>on</strong>ly <strong>on</strong><br />

r <str<strong>on</strong>g>and</str<strong>on</strong>g> can be explicitly written up<strong>on</strong> using Equati<strong>on</strong>s (A.6) <str<strong>on</strong>g>and</str<strong>on</strong>g> (A.8) as<br />

λ(r, φ) = a ℜ(r)<br />

r = aeV (r) . (A.14)<br />

We can now evaluate Γ(u, u ′ ) in the limit u ′ → u + a, where this c<strong>on</strong>cise notati<strong>on</strong><br />

means that the two points <strong>on</strong> the surface with coordinates u <str<strong>on</strong>g>and</str<strong>on</strong>g> u ′ are separated<br />

by an infinitesimal distance a measured <strong>on</strong> the bump. It does not matter in what


Appendix A: Green’s functi<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Isothermal Coordinates 136<br />

directi<strong>on</strong> the two points approach each other as l<strong>on</strong>g as a is small compared to the<br />

local radii of curvature. Up<strong>on</strong> using Equati<strong>on</strong>s (A.12) <str<strong>on</strong>g>and</str<strong>on</strong>g> (A.14), we obtain<br />

Γ(u, u + a) = − 1<br />

4π ln(a2 V (r)<br />

) −<br />

2π<br />

. (A.15)<br />

We note that Γ(u, u + a) for fixed a is not a c<strong>on</strong>stant like in flat space but varies<br />

with positi<strong>on</strong> as the functi<strong>on</strong> V (r), reflecting the lack of translati<strong>on</strong>al invariance <strong>on</strong><br />

an inhomogeneous surface, where properties such as the Gaussian curvature also vary<br />

with positi<strong>on</strong>.


Appendix B<br />

Geometric Potential<br />

137


Appendix B: Geometric Potential 138<br />

In this appendix we present two equivalent ways of determining the explicit<br />

form of the geometrical potential V (u) valid for azimuthally symmetric surfaces like<br />

the Gaussian bump. The starting point is the general definiti<strong>on</strong> introduced in Secti<strong>on</strong><br />

2.2.3:<br />

�<br />

V (u) ≡ −<br />

dA ′ G(u ′ ) Γ(u, u ′ ) , (B.1)<br />

with the Green’s functi<strong>on</strong> Γ as defined in Eq.(A.10). In the electrostatic analogy, V (u)<br />

is thus the potential induced by a c<strong>on</strong>tinuous distributi<strong>on</strong> of ”charge” represented by<br />

the Gaussian curvature (with sign reversed). We shall derive an analogue of Gauss<br />

law for corrugated surfaces where the curvature of the surface (with sign reversed)<br />

plays the role of a c<strong>on</strong>tinuous density of electrostatic charge.<br />

The first derivati<strong>on</strong> makes use of the fact that V (u) is a scalar under c<strong>on</strong>for-<br />

mal transformati<strong>on</strong>s. This symmetry can be checked explicitly by applying the same<br />

reas<strong>on</strong>ing adopted for the equati<strong>on</strong> satisfied by the Green’s functi<strong>on</strong> in Appendix A<br />

to Eq.(B.1). In fact, up<strong>on</strong> operating <strong>on</strong> both sides of Eq.(B.1) with the covariant<br />

Laplacian <str<strong>on</strong>g>and</str<strong>on</strong>g> using Eq.(A.10), the defining equati<strong>on</strong> for the geometric potential can<br />

be cast into the differential form:<br />

D α DαV (u) = G(u) . (B.2)<br />

The Gaussian curvature in Eq.(B.2) can be written in c<strong>on</strong>formal coordinates [44] as<br />

G(x, y) = −e −ρ(x,y) (∂ 2 x + ∂ 2 ρ(x, y)<br />

y)<br />

2<br />

, (B.3)<br />

where ρ(x, y) is the c<strong>on</strong>formal factor introduced in Appendix A. Similarly the left


Appendix B: Geometric Potential 139<br />

h<str<strong>on</strong>g>and</str<strong>on</strong>g> side of Eq.(B.2) can be expressed in c<strong>on</strong>formal coordinates [44] as<br />

D α DαV (x) = +e −ρ(x,y) (∂ 2 x + ∂ 2 y)V (x, y) . (B.4)<br />

Up<strong>on</strong> substituting Equati<strong>on</strong>s (B.3) <str<strong>on</strong>g>and</str<strong>on</strong>g> (B.4) in Eq.(B.2), we c<strong>on</strong>clude immediately<br />

that the geometric potential in c<strong>on</strong>formal coordinates V (x) is given by<br />

ρ(x, y)<br />

V (x, y) = −<br />

2<br />

. (B.5)<br />

Up<strong>on</strong> using Equati<strong>on</strong>s (A.6), (A.8) <str<strong>on</strong>g>and</str<strong>on</strong>g> (A.9) to substitute in Eq.(B.5), <strong>on</strong>e obtains<br />

the explicit form of the geometric potential for the bump parameterized by the coor-<br />

dinates (r, φ):<br />

� ∞<br />

V (r) = − dr ′<br />

r<br />

� l(r ′ ) − 1<br />

r ′ , (B.6)<br />

where the α-dependent functi<strong>on</strong> l(r) was defined in Eq.(A.4) <str<strong>on</strong>g>and</str<strong>on</strong>g> plotted in Fig. 2.2.<br />

The result of the integrati<strong>on</strong> in Eq.(B.6) is independent of φ because of azimuthal<br />

symmetry. The upper limit of integrati<strong>on</strong> is chosen c<strong>on</strong>sistently with Eq.(B.2) as<br />

usually d<strong>on</strong>e in electrostatics.<br />

A sec<strong>on</strong>d derivati<strong>on</strong> of this result is obtained by making explicit use of the<br />

azimuthal symmetry of the bump <str<strong>on</strong>g>and</str<strong>on</strong>g> deriving a covariant form of Gauss’ law which<br />

allows an intuitive underst<str<strong>on</strong>g>and</str<strong>on</strong>g>ing of the interacti<strong>on</strong> between defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature.<br />

This curved space versi<strong>on</strong> of Gauss’ law illuminates the electrostatic analogy used<br />

throughout the text. The gradient of the geometric potential defines an ”electric<br />

field” Eα given by<br />

�<br />

Eα ≡ −DαV (u) =<br />

dA ′ G(u ′ ) DαΓ(u, u ′ ) , (B.7)


Appendix B: Geometric Potential 140<br />

where we have used Eq.(B.1). One might expect that the flux of the vector E α<br />

through a closed loop is proporti<strong>on</strong>al to the enclosed Gaussian curvature in analogy<br />

with Gauss’ law that relates the flux of the electric field to the electrostatic charge<br />

enclosed. To prove this asserti<strong>on</strong>, we invoke the generalized Stokes’ formula [40] that<br />

relates the surface integral over A of the gradient of a field to its flux through the<br />

c<strong>on</strong>tour loop C:<br />

�<br />

A<br />

dA DαE α �<br />

= −<br />

The covariant antisymmetric tensor γαβ is given by<br />

du<br />

C<br />

α γα β Eβ . (B.8)<br />

γαβ = √ gɛαβ , (B.9)<br />

where ɛαβ is the anti-symmetric tensor with ɛrφ = −ɛφr = 1. Similarly γ αβ equals ɛαβ<br />

√g<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the following identity holds [35]<br />

γ ασ γσβ = −δ α β . (B.10)<br />

The tensor γα β = γασg σβ performs anti-clockwise rotati<strong>on</strong>s of π<br />

2<br />

when acting <strong>on</strong> an<br />

arbitrary tangent vector Vβ, as can be checked by evaluating V α γα β Vβ = γαβV α V β =<br />

0, where we have used the antisymmetry of γαβ. Thus, the vector du α γα β in Eq.(B.8)<br />

represents an infinitesimal c<strong>on</strong>tour length times the inward unit vector perpendicular<br />

to it. The dot product with the field Eβ then generates the flux. To calculate the flux<br />

piercing a circular circuit centered <strong>on</strong> a Gaussian bump, we will need to explicitly<br />

evaluate γφ r ,<br />

γφ r = − r<br />

� l(r) . (B.11)


Appendix B: Geometric Potential 141<br />

Up<strong>on</strong> using Eq.(B.7) we can rewrite the left h<str<strong>on</strong>g>and</str<strong>on</strong>g> side of Eq.(B.8) as<br />

�<br />

dA DαE α �<br />

=<br />

�<br />

dA<br />

dA ′ G(u ′ ) DβD β Γ(u, u ′ ) . (B.12)<br />

If we now recall the defining Equati<strong>on</strong> (A.10) of the Green’s functi<strong>on</strong> Γ(u, u ′ ) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

keep in mind that the Laplacian in Eq.(B.12) operates <strong>on</strong> the variables labelled by u ′<br />

(not u), we obtain using Eq.(B.8) a general result for the flux piercing a closed loop<br />

<strong>on</strong> the surface, namely<br />

�<br />

du<br />

C<br />

α γα β Eβ =<br />

�<br />

A<br />

d 2 u √ g G(u) . (B.13)<br />

We can explicitely evaluate the right h<str<strong>on</strong>g>and</str<strong>on</strong>g> side of Eq.(B.13) with the aid of Eq.(2.12).<br />

By appealing to the cylindrical symmetry, in the special case of the Gaussian bump<br />

<strong>on</strong>e can apply this covariant form of Gauss’ theorem to find the radial field E r (r) in<br />

terms of the integrated Gaussian curvature divided by the length of a boundary circle<br />

of radius r, with the result<br />

E r = 1 − � l(r)<br />

r l(r)<br />

, (B.14)<br />

where we used Equati<strong>on</strong>s (A.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> (B.11). The angular comp<strong>on</strong>ent E φ is zero<br />

everywhere. Note that E r (r) vanishes linearly with r for small r <str<strong>on</strong>g>and</str<strong>on</strong>g> decays like<br />

r2<br />

−<br />

r re 2 0 for r ≫ r0. From Eq.(B.14) <strong>on</strong>e can obtain the geometric potential V (r) by<br />

performing a line integral,<br />

� ∞<br />

V (r) =<br />

r<br />

dr ′ grr E r = −<br />

� ∞<br />

dr ′<br />

which matches the result previously obtained in Eq.(B.6).<br />

r<br />

� l(r ′ ) − 1<br />

r ′ , (B.15)


Appendix C<br />

Free energy <strong>on</strong> a corrugated plane<br />

142


Appendix C: Free energy <strong>on</strong> a corrugated plane 143<br />

In this Appendix, we derive the effective free energy for a charge neutral<br />

c<strong>on</strong>figurati<strong>on</strong> of defects c<strong>on</strong>fined <strong>on</strong> an infinite surface of varying Gaussian curvature<br />

with the topology of the plane. A general method was introduced in Ref.[43] that<br />

allows treatments of the more complicated case of deformed spheres. A detailed<br />

treatment of boundary-effects is developed in Appendix D. Here we simply assume<br />

that the size of the system is much larger than the size of the bump <str<strong>on</strong>g>and</str<strong>on</strong>g> that the<br />

boundary does not impose any topological c<strong>on</strong>straint to the director of the liquid<br />

crystal. The results presented here match those obtained in Appendix D for free<br />

boundary c<strong>on</strong>diti<strong>on</strong>s, as l<strong>on</strong>g as the defects are far from the boundary. Suppose that<br />

all the Nd defects have the same circular core radius a which does not depend <strong>on</strong> where<br />

they are located <strong>on</strong> the surface. This assumpti<strong>on</strong>s is justified if the radius of curvature<br />

r0<br />

α<br />

is much greater than a. In this limit, the microscopic physics that determines a<br />

is insensitive to the presence of the curvature <str<strong>on</strong>g>and</str<strong>on</strong>g> since the bump is locally flat a<br />

is approximately c<strong>on</strong>stant everywhere. The starting point of our analysis is the free<br />

energy expressed in terms of the singular part of the b<strong>on</strong>d angle θs(u)<br />

F = 1<br />

2 KA<br />

�<br />

dA g<br />

S<br />

αβ (∂αθs − Aα)(∂βθs − Aβ) . (C.1)<br />

The cores of the defects are excluded from the area integral in Eq.(C.1), hence S<br />

is a disc<strong>on</strong>nected domain corresp<strong>on</strong>ding to the corrugated surface punctured at the<br />

positi<strong>on</strong>s ui = (ri, φi) of the defects. In the c<strong>on</strong>formal plane parameterized by the<br />

coordinate ℜ(r) defined in Eq.(A.8), the defect cores are circles whose positi<strong>on</strong> de-<br />

pendent radius is given by ae V (ri) . The boundaries of the ”core-disks” are labelled by<br />

Ci while the circular edge of the sample of radius R is denoted by B, see Fig. C.1.


Appendix C: Free energy <strong>on</strong> a corrugated plane 144<br />

(a)<br />

Figure C.1: (a) Defects with fixed core size a <strong>on</strong> a Gaussian bump encircled by a<br />

circular boundary of radius R denoted by B. (b) The size of the vortex cores varies<br />

with positi<strong>on</strong> when plotted in the c<strong>on</strong>formal plane. One can avoid the singularities<br />

associated with the defects cores by puncturing the c<strong>on</strong>formal plane. This introduces<br />

circular boundaries Ci of varying radius at the positi<strong>on</strong> of each defect in the c<strong>on</strong>formal<br />

plane, reflecting corresp<strong>on</strong>ding c<strong>on</strong>stant core radii <strong>on</strong> the Gaussian bump.<br />

Up<strong>on</strong> introducing a ”Cauchy c<strong>on</strong>jugate” functi<strong>on</strong> χ(u) defined by<br />

the free energy in Eq.(C.1) can be cast in the form:<br />

In deriving Eq.(C.3) we used the identity<br />

C3<br />

B<br />

C1<br />

C2<br />

(b)<br />

∂αθs − Aα = γα β ∂βχ , (C.2)<br />

F = 1<br />

2 KA<br />

�<br />

dA g<br />

S<br />

αβ (∂αχ)(∂βχ) . (C.3)<br />

g µν γµ α γν β = g αβ , (C.4)<br />

which can be proved with the aid of Eq.(B.10) <str<strong>on</strong>g>and</str<strong>on</strong>g> the discussi<strong>on</strong> following it.<br />

Eq.(C.4) implies that the (covariant) dot product between two vectors after rotating<br />

each of them by π<br />

2<br />

is equivalent to taking the dot product between the two initial<br />

vectors. The integral in Eq.(C.3) can be rewritten as<br />

F<br />

KA<br />

= 1<br />

�<br />

dA Dα(χD<br />

2 S<br />

α χ) − 1<br />

�<br />

dA χDαD<br />

2 S<br />

α χ , (C.5)


Appendix C: Free energy <strong>on</strong> a corrugated plane 145<br />

where<br />

Dα(χD α χ) ≡ (1/ √ g)∂α( √ gg αβ χ∂βχ) , (C.6)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> D α Dα is defined in Eq.(A.11). Up<strong>on</strong> taking an additi<strong>on</strong>al covariant derivative,<br />

we can recast Eq.(C.2) in the form of a Poiss<strong>on</strong> equati<strong>on</strong> for the electrostatic-like<br />

potential χ(u),<br />

DαD α χ(u) = −ρ(u) , (C.7)<br />

where the analogue of the electrostatic charge density ρ(u) is given by<br />

Nd � δ(u, ui)<br />

ρ(u) ≡ qi √ − G(u) . (C.8)<br />

g<br />

i=1<br />

It is useful to compare Eq.(C.7) to Eq.(B.2) used to define the geometric potential<br />

in Appendix B. Both expressi<strong>on</strong>s are Poiss<strong>on</strong> equati<strong>on</strong>s, the <strong>on</strong>ly difference being<br />

that the source term of Eq.(C.8) includes both the point-like charges of the defects<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the Gaussian curvature with its sign reversed. Hence the Gauss law discussed in<br />

Appendix B for the geometric field Eα applies also to ∂αχ, provided that Eq.(B.13)<br />

is suitably modified to include the c<strong>on</strong>tributi<strong>on</strong> from the topological charges of the<br />

defects:<br />

�<br />

du<br />

C<br />

α γα β ∂βχ =<br />

�<br />

A<br />

d 2 u √ g<br />

�<br />

Nd �<br />

G(u) −<br />

i=1<br />

qi<br />

�<br />

δ(u, ui)<br />

√<br />

g<br />

, (C.9)<br />

where C is the c<strong>on</strong>tour enclosing an arbitrary surface A. This relati<strong>on</strong> will be useful<br />

later.<br />

We can formally solve for χ(u) in Eq.(C.7) in terms of the Green’s functi<strong>on</strong><br />

Γ(u, u ′ ) found in Appendix A:<br />

�<br />

χ(u) =<br />

dA<br />

A<br />

′ ρ(u ′ ) Γ(u, u ′ ) , (C.10)


Appendix C: Free energy <strong>on</strong> a corrugated plane 146<br />

where boundary terms have been dropped using charge neutrality <str<strong>on</strong>g>and</str<strong>on</strong>g> the fact that<br />

the edge of the sample is assumed to be far away from the defects. The integral in<br />

Eq.(C.10) can be evaluated, with the result<br />

Nd �<br />

�<br />

χ(u) = qiΓ(u, ui) −<br />

i=1<br />

Up<strong>on</strong> using Eq.(B.1), we obtain:<br />

dA<br />

A<br />

′ G(u ′ ) Γ(u, u ′ ) . (C.11)<br />

Nd �<br />

χ(u) = qiΓ(u, ui) + V (u) . (C.12)<br />

i=1<br />

We note that the first term is singular at positi<strong>on</strong>s ui but when χ is substituted in<br />

Eq.(C.3) the resulting energy is finite because the core of the defects are excluded<br />

from the domain of integrati<strong>on</strong> S. Up<strong>on</strong> substituting Eq.(C.7) in the sec<strong>on</strong>d term of<br />

Eq.(C.5) we obtain:<br />

�<br />

S<br />

dA χDαD α �<br />

χ = − dA χ ρ(u)<br />

� S<br />

= dA χ G(u) , (C.13)<br />

where we dropped terms involving the delta functi<strong>on</strong>s in Eq.(C.8) because they vanish<br />

everywhere except at the coordinates of the defects which are excluded from the<br />

domain of integrati<strong>on</strong> S. Up<strong>on</strong> substituting Eq.(C.12) in Eq.(C.13) we obtain:<br />

where we used Eq.(B.1).<br />

−<br />

+<br />

�<br />

�<br />

S<br />

dA χDαD α Nd �<br />

χ = qiV (ui)<br />

�<br />

dA<br />

S<br />

i=1<br />

dA ′ G(u) Γ(u, u ′ ) G(u ′ ) , (C.14)<br />

To evaluate the first term in Eq.(C.5), we apply the generalized Stokes for-<br />

mula of Eq.(B.8) <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>vert the surface integrals into line integrals over the bound-


Appendix C: Free energy <strong>on</strong> a corrugated plane 147<br />

aries:<br />

�<br />

S<br />

dA Dα(χD α χ) =<br />

−<br />

Nd �<br />

i=1<br />

�<br />

�<br />

Ci<br />

du α χγα β Dβχ<br />

du<br />

B<br />

α χγα β χDβχ , (C.15)<br />

where the difference in sign between the two boundary integrals in Eq.(C.15) is due<br />

to the fact that the outward normals for the paths Ci are oriented opposite to the<br />

normal for B, the outermost boundary of the system.<br />

To evaluate the last term in Eq.(C.15), we note that the flux through the<br />

distant boundary B due to a charge neutral distributi<strong>on</strong> of defects is approximately<br />

zero, provided that the integrated Gaussian curvature enclosed by the boundary is<br />

vanishingly small (see Eq.(C.9)). Hence<br />

�<br />

χ du<br />

B<br />

α γα β Dβχ � 0 , (C.16)<br />

where we used the fact that χ, defined in Eq.(C.2), is approximately c<strong>on</strong>stant <strong>on</strong> B<br />

since ∂αθ−Aα � 0. By c<strong>on</strong>trast, the flux of ∂rχ piercing the boundary Ci in Eq.(C.15)<br />

is approximately equal to the charge qi of the inclosed defect 1 . In evaluating the<br />

integrals around the infinitesimal boundaries Ci, we used the fact that the functi<strong>on</strong><br />

χ(ui + a) evaluated <strong>on</strong> the ”rim” of the defect core centered at ui <str<strong>on</strong>g>and</str<strong>on</strong>g> of radius<br />

ae V (ui) is dominated by a logarithmically diverging c<strong>on</strong>tributi<strong>on</strong> due to the i th defect.<br />

This leading c<strong>on</strong>tributi<strong>on</strong> is approximately c<strong>on</strong>stant <strong>on</strong> Ci. On the other h<str<strong>on</strong>g>and</str<strong>on</strong>g>, the<br />

n<strong>on</strong> diverging part of χ(ui + a) is multiplied by the perimeter of Ci <str<strong>on</strong>g>and</str<strong>on</strong>g> hence its<br />

c<strong>on</strong>tributi<strong>on</strong> is of the order of a. The result of the integrati<strong>on</strong> will be insensitive to<br />

1 The integrated Gaussian curvature in the microscopic disk is vanishingly small.


Appendix C: Free energy <strong>on</strong> a corrugated plane 148<br />

the orientati<strong>on</strong> of the vectors al<strong>on</strong>g the boundary Ci, provided the defect core is small<br />

2 . In this way, we find<br />

�<br />

Ci<br />

du α γα β �<br />

χDβχ � χ(ui + a)<br />

Ci<br />

du φ γφ r ∂rχ<br />

� qiχ(ui + a) . (C.17)<br />

Up<strong>on</strong> substituting Equati<strong>on</strong>s (C.17) <str<strong>on</strong>g>and</str<strong>on</strong>g> (C.16) in Eq.(C.15), we obtain<br />

�<br />

S<br />

dA Dα(χD α χ) =<br />

Nd �<br />

i=1<br />

qiV (ui) +<br />

+<br />

Nd �<br />

i=1<br />

Nd Nd �<br />

i=1<br />

qiχ(ui + a) =<br />

�<br />

qiqjΓ(ui, uj)<br />

j�=i<br />

Nd �<br />

q 2 i Γ(ui, ui + a) , (C.18)<br />

where we used Eq.(C.12) to substitute for χ. Substituti<strong>on</strong> of Eq.(C.18) <str<strong>on</strong>g>and</str<strong>on</strong>g> Eq.(C.14)<br />

in Eq.(C.5) then yields:<br />

F<br />

KA<br />

i=1<br />

Nd �<br />

= F0 + qi(1 − qi<br />

4π )V (ri) −<br />

+ 1<br />

2<br />

i=1<br />

Nd Nd �<br />

i=1<br />

Nd �<br />

i=1<br />

qi 2<br />

8π ln(a2 )<br />

�<br />

qiqj Γ(ui, uj) , (C.19)<br />

j�=i<br />

where we have used Eq.(A.15) to evaluate Γ(ui, ui + a). The first term in Eq.(C.19)<br />

is the free energy of the smooth defect-free texture (see Eq.(2.17) <str<strong>on</strong>g>and</str<strong>on</strong>g> preceding dis-<br />

cussi<strong>on</strong>). The free energy difference △F<br />

KA<br />

between a charge neutral defect c<strong>on</strong>figurati<strong>on</strong><br />

2 If the defect core is very large, it may be necessary to place images within the defect core itself to impose<br />

a desired boundary c<strong>on</strong>diti<strong>on</strong> <strong>on</strong> its rim. This is not the regime c<strong>on</strong>sidered in the present work (see Ref.<br />

[84] for similar calculati<strong>on</strong>s performed in flat space).


Appendix C: Free energy <strong>on</strong> a corrugated plane 149<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> a smooth texture thus reads<br />

∆F (α)<br />

KA<br />

= 1<br />

2<br />

+<br />

Nd Nd �<br />

i=1<br />

Nd �<br />

qi<br />

i=1<br />

�<br />

qiqjΓa(ri, φi, rj, φj) + Ec<br />

where the subscript a in the Green’s functi<strong>on</strong> indicates<br />

j�=i<br />

Nd �<br />

q<br />

KA<br />

i=1<br />

2 i<br />

�<br />

1 − qi<br />

�<br />

V (ri) , (C.20)<br />

4π<br />

Γa(ri, φi, rj, φj) = − 1<br />

4π ln[ℜ(r)2<br />

a2 − 2 ℜ(r)<br />

a<br />

ℜ(r ′ )<br />

a<br />

+ ℜ(r′ ) 2<br />

a 2<br />

cos(φ − φ ′ )] . (C.21)<br />

In order to absorb the core radius a in the inter-defect interacti<strong>on</strong>, we used the<br />

elementary algebraic identity<br />

Nd �<br />

i=1<br />

q 2 Nd Nd � �<br />

i = − qiqj , (C.22)<br />

i=1<br />

valid for charge neutral c<strong>on</strong>figurati<strong>on</strong>s. The core energy Ec in Eq.(C.20) was added<br />

by h<str<strong>on</strong>g>and</str<strong>on</strong>g> <str<strong>on</strong>g>and</str<strong>on</strong>g> represents short distance physics <strong>on</strong> scales less than or equal to the core<br />

radius. Although the sec<strong>on</strong>d part of the last term in Eq.(C.20) arises as a positi<strong>on</strong><br />

dependent self energy, it has the same functi<strong>on</strong>al form as the geometrical potential<br />

discussed in Appendix B, <str<strong>on</strong>g>and</str<strong>on</strong>g> hence depends <strong>on</strong> the global shape of the surface.<br />

j�=i


Appendix D<br />

Bump with a boundary<br />

150


Appendix D: Bump with a boundary 151<br />

(a)<br />

R R<br />

Figure D.1: (a) Schematic illustrati<strong>on</strong> of the boundary director texture corresp<strong>on</strong>ding<br />

to free boundary c<strong>on</strong>diti<strong>on</strong>s. The vector order parameter orientati<strong>on</strong> close to the<br />

edge of the sample does not vary appreciably as <strong>on</strong>e moves al<strong>on</strong>g the radial directi<strong>on</strong><br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> it is parallel to itself at every point <strong>on</strong> the boundary (b) Tangential boundary<br />

c<strong>on</strong>diti<strong>on</strong>s. The vector order parameter is locally aligned to a wall located at the edge<br />

of the sample.<br />

(b)<br />

The aim of this appendix is to study the energetics of a singular vector field<br />

<strong>on</strong> a bumpy surface of circular shape <str<strong>on</strong>g>and</str<strong>on</strong>g> finite size R. To evaluate the ground state<br />

energy in Eq.(C.1), we first need to solve the covariant Laplace equati<strong>on</strong> for the b<strong>on</strong>d<br />

angle field θ(u) in the presence of Nd defects at positi<strong>on</strong>s ui. This is more easily<br />

d<strong>on</strong>e by switching from θ(u) to the c<strong>on</strong>jugate field χ(u), as shown in Eq.(C.3), <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

solving the Poiss<strong>on</strong> Equati<strong>on</strong> (C.7) satisfied by χ for both free <str<strong>on</strong>g>and</str<strong>on</strong>g> fixed boundary<br />

c<strong>on</strong>diti<strong>on</strong>s (see Fig. D.1). The b<strong>on</strong>d angle field satisfies free boundary c<strong>on</strong>diti<strong>on</strong>s if<br />

the following relati<strong>on</strong> holds <strong>on</strong> the circular edge B:<br />

while for fixed tangential boundary c<strong>on</strong>diti<strong>on</strong>s we have:<br />

∂rθ|r=R = 0 , (D.1)<br />

∂φθ|r=R = 0 . (D.2)<br />

To underst<str<strong>on</strong>g>and</str<strong>on</strong>g> Eq.(D.2), recall that we measure the b<strong>on</strong>d angle θ with respect to a<br />

rotating basis vector Er in the radial directi<strong>on</strong>. With this c<strong>on</strong>venti<strong>on</strong>, θ is equal to a


Appendix D: Bump with a boundary 152<br />

c<strong>on</strong>stant when the vector order parameter is aligned with the circular boundary B.<br />

We can c<strong>on</strong>vert Equati<strong>on</strong>s (D.1) <str<strong>on</strong>g>and</str<strong>on</strong>g> (D.2) into boundary c<strong>on</strong>diti<strong>on</strong>s to be satisfied<br />

by the c<strong>on</strong>jugate field χ <strong>on</strong> B. Up<strong>on</strong> substituting Eq.(D.1) in Eq.(C.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> using the<br />

fact that Ar is equal to zero, we obtain the c<strong>on</strong>straint that χ(u) satisfies <strong>on</strong> B in the<br />

case of free boundary c<strong>on</strong>diti<strong>on</strong>s:<br />

∂φχ|r=R = 0 . (D.3)<br />

This corresp<strong>on</strong>ds to a Dirichlet problem where χ D (u) evaluated <strong>on</strong> the boundary B<br />

assumes an arbitrary c<strong>on</strong>stant value, c:<br />

since γ φ<br />

φ<br />

Up<strong>on</strong> substituting Eq.(D.2) in Eq.(C.2), we obtain<br />

χ D (B) = c . (D.4)<br />

∂rχ = − Aφ<br />

, (D.5)<br />

r<br />

γφ<br />

= 0. Up<strong>on</strong> substituting Equati<strong>on</strong>s (B.11) <str<strong>on</strong>g>and</str<strong>on</strong>g> (2.8) in Eq.(D.5) we obtain<br />

the boundary c<strong>on</strong>diti<strong>on</strong> <strong>on</strong> χ that corresp<strong>on</strong>ds to Eq.(D.2)<br />

∂rχ N |r=R = − 1<br />

R<br />

, (D.6)<br />

where the superscript indicates that this is a Neumann boundary problem with the<br />

normal derivative assuming a c<strong>on</strong>stant value.<br />

To solve the Poiss<strong>on</strong> Eq.(C.7) with Neumann or Dirichlet’s boundary c<strong>on</strong>di-<br />

ti<strong>on</strong>s in terms of suitable Green’s functi<strong>on</strong>s we exploit a covariant versi<strong>on</strong> of Green’s


Appendix D: Bump with a boundary 153<br />

theorem expressed in terms of two invariant functi<strong>on</strong>s of positi<strong>on</strong> ψ(u) <str<strong>on</strong>g>and</str<strong>on</strong>g> ϕ(u) [85]:<br />

�<br />

dA (ϕ(u) DαD<br />

S<br />

α ψ(u) − ψ(u) DαD α ϕ(u)) =<br />

�<br />

−<br />

du<br />

B<br />

α γα β (ϕ(u) ∂β ψ(u) − ψ(u) ∂β ϕ(u)) . (D.7)<br />

By applying Eq.(D.7) to ϕ(u) = χ(u) <str<strong>on</strong>g>and</str<strong>on</strong>g> ψ(u) = Γ(u, u ′ ) <str<strong>on</strong>g>and</str<strong>on</strong>g> using Equati<strong>on</strong>s<br />

(A.10) <str<strong>on</strong>g>and</str<strong>on</strong>g> (C.7) we obtain<br />

χ(u) =<br />

+<br />

−<br />

�<br />

�<br />

�<br />

dA<br />

S<br />

′ Γ(u ′ , u) ρ(u ′ )<br />

du<br />

B<br />

′α γα β χ(u ′ ) ∂ ′ βΓ(u ′ , u)<br />

du<br />

B<br />

′α γα β Γ(u ′ , u) ∂ ′ βχ(u ′ ) , (D.8)<br />

where u <str<strong>on</strong>g>and</str<strong>on</strong>g> u ′ have been exchanged 1 . The boundary c<strong>on</strong>diti<strong>on</strong>s for the Green’s<br />

functi<strong>on</strong> can be c<strong>on</strong>veniently chosen to eliminate unknown quantities in Eq.(D.7) as<br />

in flat space [86].<br />

For the Dirichlet’s problem, we choose the Green’s functi<strong>on</strong> Γ D so that it<br />

vanishes when u ′ is <strong>on</strong> the boundary B<br />

Γ D (B, u) = 0 . (D.9)<br />

Up<strong>on</strong> substituting Eq.(D.9) in Eq.(D.8) <str<strong>on</strong>g>and</str<strong>on</strong>g> noting that χ(u ′ ) is c<strong>on</strong>stant <strong>on</strong> the<br />

boundary B (see Eq.(D.4)), we obtain:<br />

χ D �<br />

(u) =<br />

+ χ D �<br />

(B)<br />

dA<br />

S<br />

′ Γ D (u ′ , u) ρ(u ′ )<br />

du<br />

B<br />

′α γα β ∂ ′ βΓ D (u ′ , u) , (D.10)<br />

1 These manipulati<strong>on</strong>s are comm<strong>on</strong> in electromagnetism, see for example Ref. [86].


Appendix D: Bump with a boundary 154<br />

The c<strong>on</strong>tour integral in the sec<strong>on</strong>d term of Eq.(D.10) corresp<strong>on</strong>ds to the flux piercing<br />

B which is, in turn, equal to the (unit) charge of the singularity located at u 2 . The<br />

final result reads<br />

χ D �<br />

(u) =<br />

dA<br />

S<br />

′ ρ(u ′ ) Γ D (u ′ , u) + χ D (B) , (D.11)<br />

where χ D (B) can be set to zero, since the energy in Eq.(C.3) is <strong>on</strong>ly defined in terms<br />

of derivatives of χ. One can check that χ D (u) in Eq.(D.11) satisfies both the Poiss<strong>on</strong>’s<br />

Equati<strong>on</strong> (C.7) <str<strong>on</strong>g>and</str<strong>on</strong>g> the required boundary c<strong>on</strong>diti<strong>on</strong> in Eq.(D.3). This can be more<br />

easily proved by noting that Γ D (u ′ , u) is symmetric under exchange of its arguments<br />

u ′ <str<strong>on</strong>g>and</str<strong>on</strong>g> u 3 .<br />

A similar reas<strong>on</strong>ing applies to χ N (u). However, we cannot choose the<br />

Green’s functi<strong>on</strong> so that the sec<strong>on</strong>d term of Eq.(D.8) (that c<strong>on</strong>tains the unknown<br />

quantity χ(u ′ )) vanishes. In fact, by invoking Stokes theorem (see Eq.(B.8)), we note<br />

that<br />

�<br />

du<br />

B<br />

′α γα β ∂ ′ βΓ(u ′ , u) = 1 , (D.12)<br />

where γα β (u ′ = B) is c<strong>on</strong>stant <strong>on</strong> the circular boundary B <str<strong>on</strong>g>and</str<strong>on</strong>g> can be brought out<br />

of the integral. An appropriate choice of boundary c<strong>on</strong>diti<strong>on</strong> <strong>on</strong> Γ N that satisfies the<br />

c<strong>on</strong>straint in Eq.(D.12) is<br />

∂r ′ ΓN (r ′ , φ ′ ; r, φ)|r ′ =R =<br />

1<br />

2πγα β (R)<br />

�<br />

l(R)<br />

= − , (D.13)<br />

2πR<br />

2 This asserti<strong>on</strong> can be proved by applying Stokes Theorem (as stated in Eq.(B.8) with Eβ replaced by<br />

∂ ′ βΓ(u ′ , u)) <str<strong>on</strong>g>and</str<strong>on</strong>g> using Eq.(A.10)) to evaluate the surface integral.<br />

3 This asserti<strong>on</strong> can be proved by applying Green’s theorem in Eq.D.7 to ψ(u) = Γ(u, u ′ ) <str<strong>on</strong>g>and</str<strong>on</strong>g> ϕ(u) =<br />

Γ(u, u ′′ ) <str<strong>on</strong>g>and</str<strong>on</strong>g> noticing that the right h<str<strong>on</strong>g>and</str<strong>on</strong>g> side vanishes if the boundary c<strong>on</strong>diti<strong>on</strong> in Eq.(D.9) is assumed.<br />

We can then c<strong>on</strong>clude that Γ D (u ′ , u ′′ ) = Γ D (u ′′ , u ′ ) in analogy with familiar results in 3D electrostatics<br />

[86].


Appendix D: Bump with a boundary 155<br />

where we used Eq.(B.11). The α-dependent functi<strong>on</strong> l(r ′ ) was defined in Eq.(A.4).<br />

Note that l(R) � 1, for R ≫ r0 (see Fig. 2.2). By substituting Equati<strong>on</strong>s (D.13) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

(D.6) in Eq.(D.8) we obtain<br />

χ N (u) =<br />

�<br />

dA<br />

S<br />

′ Γ N (u ′ , u) ρ(u ′ )<br />

� 2π<br />

dφ<br />

0<br />

′ χ(R, φ ′ )<br />

� 2π<br />

dφ<br />

0<br />

′ Γ N (R, φ ′ ; r, φ) , (D.14)<br />

+ 1<br />

2π<br />

1<br />

− �<br />

l(R)<br />

The last two integral are c<strong>on</strong>stant <str<strong>on</strong>g>and</str<strong>on</strong>g> hence can be dropped. We can check explicitly<br />

that χ N (u) satisfies Eq.(D.6) by evaluating the radial derivative of χ N (u) in Eq.(D.14)<br />

∂rχ N �<br />

(r, φ)|r=R =<br />

dA<br />

S<br />

′ ρ(u ′ ) ∂rΓ N (R, φ ′ ; r, φ)|r=R , (D.15)<br />

where the radial derivative of the Green’s functi<strong>on</strong> assumes the c<strong>on</strong>stant value derived<br />

in Eq.(D.13), provided that Γ N (u ′ , u) is c<strong>on</strong>structed so that it is symmetric under<br />

exchange of its arguments u ′ <str<strong>on</strong>g>and</str<strong>on</strong>g> u (see Eq.(D.19)). With the aid of Equati<strong>on</strong>s (C.8)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> (2.12), we obtain<br />

�<br />

dA<br />

S<br />

′ ρ(u ′ Nd<br />

) =<br />

i<br />

� �<br />

�<br />

1<br />

qi + 2π � − 1 . (D.16)<br />

l(R)<br />

Up<strong>on</strong> substituting Equati<strong>on</strong>s (D.16) <str<strong>on</strong>g>and</str<strong>on</strong>g> (D.13) in Eq.(D.15), we c<strong>on</strong>clude that the<br />

the Neumann boundary c<strong>on</strong>diti<strong>on</strong> in Eq.(D.6) is satisfied provided that<br />

Nd �<br />

i<br />

qi = 2π . (D.17)<br />

It is reassuring that the topological c<strong>on</strong>straint <strong>on</strong> the vorticity of the field imposed<br />

by the presence of the wall arises as a natural requirement within this formalism.<br />

Similarly, the Poiss<strong>on</strong>’s equati<strong>on</strong> for χ N (u) is automatically satisfied.


Appendix D: Bump with a boundary 156<br />

B<br />

R<br />

(a)<br />

+ - + -<br />

+ +<br />

Figure D.2: Schematic illustrati<strong>on</strong> of the method of images. The image defect is<br />

of the same sign for free boundary c<strong>on</strong>diti<strong>on</strong>s (a) <str<strong>on</strong>g>and</str<strong>on</strong>g> opposite for fixed boundary<br />

c<strong>on</strong>diti<strong>on</strong>s (b). Defects closer to the center of the circle have images further away<br />

from it.<br />

We are now left with the task of guessing the Green’s functi<strong>on</strong>s for the<br />

Dirichlet’s <str<strong>on</strong>g>and</str<strong>on</strong>g> Neumann problems satisfying the boundary c<strong>on</strong>diti<strong>on</strong>s in Equati<strong>on</strong>s<br />

(D.9) <str<strong>on</strong>g>and</str<strong>on</strong>g> (D.13) respectively. In both cases the Green’s functi<strong>on</strong> can be determined<br />

by the method of images applied in the c<strong>on</strong>formal plane. For every defect with<br />

radial coordinate ri we need an image defect of opposite (equal) topological charge<br />

at positi<strong>on</strong> r ′ i to ensure that Dirichlet’s (Neumann) boundary c<strong>on</strong>diti<strong>on</strong>s are enforced<br />

(see Fig. D.2). The radial coordinate of the image defect r ′ i is determined by the<br />

relati<strong>on</strong>:<br />

ℜ(r ′ i) = ℜ 2 (R)<br />

ℜ(ri)<br />

B<br />

R<br />

(b)<br />

. (D.18)<br />

Except for the coordinate change r → ℜ(r), a similar relati<strong>on</strong> arises in elementary<br />

electrostatic problems in flat space [48]. A geometric argument that justifies this<br />

choice of images in flat space is illustrated in Fig. D.3.<br />

Once the positi<strong>on</strong> of the source is chosen according to Eq.(D.18), we can


Appendix D: Bump with a boundary 157<br />

express the two Green’s functi<strong>on</strong>s with the c<strong>on</strong>cise notati<strong>on</strong> Γ D/N as follows:<br />

Γ D/N (u, u ′ ) = − 1<br />

4π ln[ℜ(r)2 + ℜ(r ′ ) 2 − 2ℜ(r)ℜ(r ′ ) cos(φ − φ ′ )]<br />

± 1<br />

4π ln[ℜ(r)2 + ℜ(R)4<br />

ℜ(r ′ − 2ℜ(r)ℜ(R)2<br />

) 2<br />

ℜ(r ′ ) cos(φ − φ′ )] ± f(r ′ ) ,<br />

(D.19)<br />

where we have introduced a functi<strong>on</strong> f(r ′ ) to make Γ D/N (u, u ′ ) symmetric under<br />

exchange of its arguments <str<strong>on</strong>g>and</str<strong>on</strong>g> to remove a singularity at r ′ = 0. Note that we<br />

can add f(r ′ ) since the defining equati<strong>on</strong> of the Green’s functi<strong>on</strong> does not c<strong>on</strong>tain<br />

derivatives of r ′ , <strong>on</strong>ly of r.<br />

f(r ′ ) = 1<br />

2π ln<br />

� � ′ ℜ (r )<br />

ℜ(R)<br />

. (D.20)<br />

The plus <str<strong>on</strong>g>and</str<strong>on</strong>g> minus signs in Eq.(D.19) insure that the Dirichlet <str<strong>on</strong>g>and</str<strong>on</strong>g> Neumann bound-<br />

ary c<strong>on</strong>diti<strong>on</strong>s respectively are obeyed. In what follows the sign placed above in the<br />

symbols ± or ∓ always indicates the choice suitable for the Dirichlet’s problem while<br />

the <strong>on</strong>e below refers to Neumann’s boundary c<strong>on</strong>diti<strong>on</strong>s. One can explicitly check by<br />

substituti<strong>on</strong> that the symmetrized Green’s functi<strong>on</strong>s Γ D/N (u, u ′ ) satisfy the correct<br />

boundary c<strong>on</strong>diti<strong>on</strong>s, as l<strong>on</strong>g as the plus sign is chosen when Γ D is substituted in<br />

Eq.(D.9) <str<strong>on</strong>g>and</str<strong>on</strong>g> the minus sign when Γ N is substituted in Eq.(D.13). Note that, with-<br />

out the extra term f(r ′ ) in the expressi<strong>on</strong>s for both Green’s functi<strong>on</strong>s, Γ D would not<br />

be equal to zero <strong>on</strong> the boundary B <str<strong>on</strong>g>and</str<strong>on</strong>g> the last term in Eq.(D.14) would not be<br />

c<strong>on</strong>stant when Γ N is substituted in.<br />

Once the Green’s functi<strong>on</strong> is obtained, <strong>on</strong>e can readily write down χ D/N (u)


Appendix D: Bump with a boundary 158<br />

by dropping the c<strong>on</strong>stant terms in Equati<strong>on</strong>s (D.11) <str<strong>on</strong>g>and</str<strong>on</strong>g> (D.14)<br />

χ D/N (u) =<br />

−<br />

Nd �<br />

i=1<br />

�<br />

qiΓ D/N (u, ui)<br />

dA<br />

S<br />

′ G(u ′ ) Γ D/N (u, u ′ ) . (D.21)<br />

The Gaussian curvature is given by the covariant Laplacian of the geometric potential<br />

introduced in Eq.(B.2). Up<strong>on</strong> integrating by parts twice the sec<strong>on</strong>d term in Eq.(D.21)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> applying Stokes theorem repeatedly, we find<br />

χ D/N Nd �<br />

(u) = qiΓ D/N (u, ui) + V (u) , (D.22)<br />

i=1<br />

where we assume that R ≫ r0 so that we can neglect boundary terms. The geometric<br />

potential V (u) has the same functi<strong>on</strong>al form previously discussed in Appendix B,<br />

despite the change in the Green’s functi<strong>on</strong>.<br />

The evaluati<strong>on</strong> of the energy stored in the field now proceeds al<strong>on</strong>g the lines<br />

sketched in Appendix C with the <strong>on</strong>ly caveat that <strong>on</strong>e needs to choose the appropriate<br />

Green’s functi<strong>on</strong> in Eq.(D.19). In the case of Dirichlet’s boundary c<strong>on</strong>diti<strong>on</strong>s <strong>on</strong>e can<br />

prove that the boundary integral in Eq.(C.16) still vanishes by virtue of the fact that<br />

χ is c<strong>on</strong>stant <strong>on</strong> the boundary B <str<strong>on</strong>g>and</str<strong>on</strong>g> the defects c<strong>on</strong>figurati<strong>on</strong> is charge neutral. For<br />

the Neumann’s problem we have:<br />

χ N<br />

�<br />

du<br />

B<br />

α γα β Dβχ N ≈ 4π ln (ℜ(R)) , (D.23)<br />

where we assumed that R ≫ r0. In this limit ℜ(R) is approximately equal to R, as<br />

can be checked with the aid of Equati<strong>on</strong>s (A.8) <str<strong>on</strong>g>and</str<strong>on</strong>g> (A.9).<br />

All the remaining intermediate steps to derive the free energy follow as in<br />

Appendix C without further assumpti<strong>on</strong>s. We can readily generalize Eq.(C.20) to


Appendix D: Bump with a boundary 159<br />

Figure D.3: A topological defect located at positi<strong>on</strong> P in a circular domain of radius<br />

OQ in flat space. Fixed boundary c<strong>on</strong>diti<strong>on</strong>s are obtained by placing an image defect<br />

of the same sign at a distance OP ′ from the center such that OP OP ′ = OQ 2 . The<br />

two triangles △OQP <str<strong>on</strong>g>and</str<strong>on</strong>g> △OQP ′ are similar <str<strong>on</strong>g>and</str<strong>on</strong>g> ∠OQP = ∠OP ′ Q = π −θ ′ . By the<br />

theorem of the external angle, we c<strong>on</strong>clude that θ ′ +θ = φ+π as l<strong>on</strong>g as Q lies <strong>on</strong> the<br />

circumference B. This is equivalent to the boundary c<strong>on</strong>diti<strong>on</strong> in Eq.(D.2) if a n<strong>on</strong><br />

rotating vector basis is used. Similarly, for free boundary c<strong>on</strong>diti<strong>on</strong>s the symmetric<br />

Green’ functi<strong>on</strong> evaluated is c<strong>on</strong>stant <strong>on</strong> the boundary if the image defect is negative.<br />

Since<br />

P Q<br />

P ′ Q<br />

= OP<br />

OQ , the potential ln(P Q) − ln(P ′ Q) (generated by the defect at distance<br />

OP <str<strong>on</strong>g>and</str<strong>on</strong>g> its image) is c<strong>on</strong>stant <strong>on</strong> the circumference of radius OQ.


Appendix D: Bump with a boundary 160<br />

evaluate the energy stored in the singular field in the presence of a boundary in the<br />

case of both free <str<strong>on</strong>g>and</str<strong>on</strong>g> fixed boundary c<strong>on</strong>diti<strong>on</strong>s. We assume R ≫ r0, but the defects<br />

do not need to be far away from the boundary. The result is<br />

F D/N<br />

KA<br />

= 1<br />

2<br />

+<br />

±<br />

Nd �<br />

j�=i<br />

Nd �<br />

i=1<br />

Nd �<br />

i=1<br />

qiqj Γ D/N (xi; xj) + F0<br />

qi(1 − qi<br />

4π )V (ri) +<br />

Nd �<br />

i=1<br />

qi 2<br />

4π ln<br />

� �<br />

ℜ(R)<br />

qi 2<br />

4π ln � 1 − x 2� i , (D.24)<br />

where F0 is defined in Eq.(2.17). The Green’s functi<strong>on</strong> expressed in scaled coordinates<br />

reads<br />

Γ D/N (xi; xj) = − 1<br />

4π ln � x 2 i + x 2 j − 2xixj cos (φi − φj) �<br />

± 1<br />

4π ln � x 2 i x 2 j + 1 − 2xixj cos (φi − φj) � .<br />

a<br />

(D.25)<br />

In the case of Neumann’s boundary c<strong>on</strong>diti<strong>on</strong>s, we have suppressed a term diverging<br />

like ln(ℜ(R)) associated with the boundary c<strong>on</strong>tributi<strong>on</strong> in Eq.(D.23). Eq.(D.25) is<br />

expressed in terms of a dimensi<strong>on</strong>less defect positi<strong>on</strong> xi<br />

xi ≡ ℜ(ri)<br />

ℜ(R)<br />

. (D.26)<br />

The plus sign in Equati<strong>on</strong>s (D.25) <str<strong>on</strong>g>and</str<strong>on</strong>g> (D.24) is to be chosen for Dirichlet’s boundary<br />

c<strong>on</strong>diti<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> the minus sign for Neumann’s. The last term in Eq.(D.24) represents<br />

the interacti<strong>on</strong> , U D/N<br />

b (xi), between a single defect located at xi <str<strong>on</strong>g>and</str<strong>on</strong>g> the boundary<br />

U D/N<br />

b<br />

Nd � qi<br />

(xi) = ±KA<br />

2<br />

4π ln � 1 − x 2� i . (D.27)<br />

i=1


Appendix D: Bump with a boundary 161<br />

Note that the q-dependent prefactors of U D/N<br />

b (xi) <str<strong>on</strong>g>and</str<strong>on</strong>g> the quadratic correcti<strong>on</strong> to the<br />

curvature interacti<strong>on</strong> (III term in Eq.(D.24) have the same magnitude. This is not a<br />

coincidence but a clue to their comm<strong>on</strong> origin. As the geometry of a plane is modified,<br />

either by creating a varying curvature or imposing boundaries, the defects feel an<br />

additi<strong>on</strong>al interacti<strong>on</strong> caused by the c<strong>on</strong>formal transformati<strong>on</strong> of the underlying space.<br />

This line of reas<strong>on</strong>ing is powerful <str<strong>on</strong>g>and</str<strong>on</strong>g> it has been pursued in Ref. [43] to explain<br />

some basic features of the interacti<strong>on</strong> between defects <str<strong>on</strong>g>and</str<strong>on</strong>g> curvature without explicit<br />

recourse to the Green’s functi<strong>on</strong> techniques adopted in this work.


Appendix E<br />

Free energy of a vector field <strong>on</strong> a<br />

sphere<br />

162


Appendix E: Free energy of a vector field <strong>on</strong> a sphere 163<br />

We start our analysis with the Frank free energy with splay <str<strong>on</strong>g>and</str<strong>on</strong>g> bend terms<br />

proporti<strong>on</strong>al to K1 <str<strong>on</strong>g>and</str<strong>on</strong>g> K3 <str<strong>on</strong>g>and</str<strong>on</strong>g> expressed in terms of the covariant derivative Di n j<br />

where<br />

F = 1<br />

�<br />

2<br />

The Christoffel c<strong>on</strong>necti<strong>on</strong> Γ j<br />

i k<br />

d 2 x √ �<br />

g [K1 Di n i�2 + K3 (Di nj − Dj ni) � D i n j − D j n i� ] , (E.1)<br />

Di n j = ∂i n j + Γ j<br />

i k nk . (E.2)<br />

[60] is unchanged if the lower indices i <str<strong>on</strong>g>and</str<strong>on</strong>g> k are<br />

interchanged. As a result, the covariant derivative Di can be replaced by ∂i in the<br />

sec<strong>on</strong>d term of Eq.(E.1) because the covariant curl is antisymmetric. It follows that<br />

The covariant form of the divergence is given by [60]<br />

�D × �n = ∇ × �n . (E.3)<br />

Di n i ≡ 1 �√ i<br />

√ ∂i g n<br />

g � . (E.4)<br />

For a rigid sphere of radius R with polar coordinates {θ, φ}, we have √ g = R 2 sin θ,<br />

Γ θ φ φ<br />

φ φ<br />

j<br />

= − sin θ cos θ, Γ φ θ = Γ θ φ = − cot θ, <str<strong>on</strong>g>and</str<strong>on</strong>g> all other Γ i k = 0.<br />

Up<strong>on</strong> adding <str<strong>on</strong>g>and</str<strong>on</strong>g> subtracting the expressi<strong>on</strong> for the curl of n multiplied by<br />

K1 from the first <str<strong>on</strong>g>and</str<strong>on</strong>g> sec<strong>on</strong>d term in Eq.(E.1) we obtain<br />

F = 1<br />

�<br />

2<br />

d 2 x √ �<br />

g [K1 Di n j ) ( D i �<br />

nj<br />

+(K3 − K1)(∇ × n) 2 ] . (E.5)<br />

Similarly, up<strong>on</strong> adding <str<strong>on</strong>g>and</str<strong>on</strong>g> subtracting the covariant divergence of n multi-


Appendix E: Free energy of a vector field <strong>on</strong> a sphere 164<br />

plied by K3 from the sec<strong>on</strong>d <str<strong>on</strong>g>and</str<strong>on</strong>g> first term in Eq.(E.1) we obtain<br />

F = 1<br />

�<br />

2<br />

d 2 x √ �<br />

g [K3 Di n j ) ( D i �<br />

nj<br />

+ (K1 − K3) (∇ · n) 2 ] . (E.6)<br />

Up<strong>on</strong> adding the two equivalent expressi<strong>on</strong>s for F in Equati<strong>on</strong>s (E.5) <str<strong>on</strong>g>and</str<strong>on</strong>g> (E.6) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

dividing by two we can express the free energy in terms of the c<strong>on</strong>stants<br />

namely<br />

F = K<br />

2<br />

�<br />

ɛ ≡ K3 − K1<br />

K3 + K1<br />

K ≡ K3 + K1<br />

2<br />

d 2 x √ g [ � Di n j ) ( D i �<br />

nj<br />

, (E.7)<br />

, (E.8)<br />

+ɛ � (∇ × n) 2 − (∇ · n) 2� ] . (E.9)<br />

We now parameterize the orientati<strong>on</strong> of the unit vector n in terms of the<br />

b<strong>on</strong>d angle field α(θ, φ) that it forms with respect to the l<strong>on</strong>gitudinal directi<strong>on</strong> �eθ<br />

which in polar coordinates (θ, φ) is given by<br />

eθ = R(cos θ sin φ, cos θ cos φ, − sin θ) , (E.10)<br />

while the orthog<strong>on</strong>al unit vector eφ is given by<br />

eφ = R(− sin θ sin φ, sin θ cos φ, 0) . (E.11)<br />

The comp<strong>on</strong>ents of the vector n with respect to �eφ <str<strong>on</strong>g>and</str<strong>on</strong>g> �eθ are given by<br />

n θ =<br />

n φ =<br />

cos α<br />

,<br />

R<br />

(E.12)<br />

sin α<br />

.<br />

R sin θ<br />

(E.13)


Appendix E: Free energy of a vector field <strong>on</strong> a sphere 165<br />

Up<strong>on</strong> substituting the relevant expressi<strong>on</strong>s for the n<strong>on</strong> vanishing comp<strong>on</strong>ents of the<br />

c<strong>on</strong>necti<strong>on</strong> in the covariant derivative (see Eq.(E.2)) <str<strong>on</strong>g>and</str<strong>on</strong>g> using Eq.(E.13), the free<br />

energy density f(θ, φ) is given by<br />

4R2 K1 + K3<br />

f =<br />

� �2 ∂α<br />

+ Υα<br />

∂θ<br />

2 +<br />

(θ, φ)<br />

��∂α �2 ɛ cos 2α − Υα 2 �<br />

(θ, φ)<br />

+ 2ɛ sin 2α<br />

∂θ<br />

� ∂α<br />

∂θ<br />

�<br />

Υα(θ, φ) , (E.14)<br />

where the α-dependent functi<strong>on</strong> Υα(θ, φ) is<br />

Υα(θ, φ) ≡ 1<br />

� �<br />

∂α<br />

+ cos θ . (E.15)<br />

sin θ ∂φ<br />

The energies of the latitudinal (α = π/2) <str<strong>on</strong>g>and</str<strong>on</strong>g> l<strong>on</strong>gitudinal (α = 0) tilted-molecules<br />

textures favored for ɛ less or greater than zero respectively (see secti<strong>on</strong> 3.2.1) are easily<br />

determined by substituting the appropriate α <str<strong>on</strong>g>and</str<strong>on</strong>g> ɛ in Eq.(E.14). After integrating<br />

the free energy density f in Eq.(E.14) we obtain<br />

� π<br />

2 cos<br />

F = 2πK(1 − |ɛ|)<br />

a<br />

R<br />

2 (θ)<br />

sin(θ) dθ<br />

� � � �<br />

2R<br />

= 2πK(1 − |ɛ|) ln − 1 + 2Ec , (E.16)<br />

a<br />

where we have introduced a core radius, a, <str<strong>on</strong>g>and</str<strong>on</strong>g> corresp<strong>on</strong>ding core energy Ec for each<br />

defect.<br />

In the zero anisotropy limit (ɛ = 0), <strong>on</strong>ly the first line in Eq.(E.15) survives.<br />

The resulting free energy F = � dSf(θ, φ) then matches the <strong>on</strong>e obtained using the<br />

spin c<strong>on</strong>necti<strong>on</strong> in the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong> up<strong>on</strong> setting K1 = K3 = K,<br />

F = 1<br />

2 K<br />

�<br />

dS g ij (∂iα − Ai)(∂jα − Aj) , (E.17)


Appendix E: Free energy of a vector field <strong>on</strong> a sphere 166<br />

where dS = dθdφ R2 sin θ <str<strong>on</strong>g>and</str<strong>on</strong>g> the metric tensor gφφ = diag � 1<br />

R2 1 , R2 sin2 �<br />

. The curl of<br />

θ<br />

the spin-c<strong>on</strong>necti<strong>on</strong> Ai is the Gaussian curvature 1<br />

R 2 [40, 42] <str<strong>on</strong>g>and</str<strong>on</strong>g> its <strong>on</strong>ly n<strong>on</strong>-vanishing<br />

comp<strong>on</strong>ent is Aφ = − cos θ.<br />

We now adopt a Coulomb gas representati<strong>on</strong> of the liquid crystal free energy<br />

(in the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>) obtained by exploiting in Eq.(E.17) the<br />

relati<strong>on</strong><br />

γ ij ∂i(∂jα − Aj) = s(u) − G(u) ≡ n(u) , (E.18)<br />

where γ ij is the covariant antisymmetric tensor [40], G(u) is the Gaussian curvature<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> s(u) ≡ 1 �Nd √<br />

g i=1 qiδ(u − ui) is the disclinati<strong>on</strong> density with Nd defects of charge<br />

qi at positi<strong>on</strong>s ui. The final result is an effective free energy whose basic degrees of<br />

freedom are the defect positi<strong>on</strong>s themselves [35, 4]:<br />

F = K<br />

2<br />

�<br />

�<br />

dA<br />

dA ′ n(u) Γ(u, u ′ ) n(u ′ ) , (E.19)<br />

where n(u), the defect density relative to the Gaussian curvature, was defined in<br />

Eq.(E.18). The equilibrium positi<strong>on</strong>s of the defects are determined <strong>on</strong>ly by defect-<br />

defect interacti<strong>on</strong>s because the Gaussian curvature is c<strong>on</strong>stant G = 1<br />

R 2 <strong>on</strong> an un-<br />

deformed sphere. To calculate the Green’s functi<strong>on</strong> Γ(u, u ′ ) we need to invert the<br />

covariant Laplacian defined <strong>on</strong> the sphere<br />

Γ(u, u ′ � �<br />

1<br />

) ≡ −<br />

∆ uu ′<br />

, (E.20)<br />

As shown below, this inversi<strong>on</strong> can be accomplished by performing a weighted sum<br />

over eigenmodes of the covariant Laplacian, [4].


Appendix E: Free energy of a vector field <strong>on</strong> a sphere 167<br />

We first recall that the (generalized) Green functi<strong>on</strong> Γ(u, u ′ ) is defined by<br />

∆u Γ(u, u ′ ) = δ(u, u′ )<br />

√ g − 1<br />

S<br />

, (E.21)<br />

where S = 4πR 2 denotes the area of the surface <str<strong>on</strong>g>and</str<strong>on</strong>g> ∆ = 1<br />

√ g ∂i( √ gg ij ∂j). The<br />

presence of the sec<strong>on</strong>d term <strong>on</strong> the left h<str<strong>on</strong>g>and</str<strong>on</strong>g> side of Eq.(E.21) can be understood<br />

as follows. The Green functi<strong>on</strong> of the Laplacian (according to the usual definiti<strong>on</strong><br />

without the area correcti<strong>on</strong> in Eq.(E.21)) can be interpreted physically as the steady<br />

temperature resp<strong>on</strong>se of the system to a point-like unit source of heat. However, for<br />

a closed system such as the surface of the sphere heat cannot escape. Hence, it is<br />

impossible to impose a point source, that would inject heat at a c<strong>on</strong>stant rate <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

have the system resp<strong>on</strong>d with a time-independent distributi<strong>on</strong>. To prevent energy<br />

from building up in such a system, we put the spherical surface of area S in c<strong>on</strong>tact<br />

with a reservoir that uniformly absorbs heat at the same rate it is pumped in. The<br />

need for subtracting the ”neutralizing background heat” 1<br />

S<br />

in Eq.(E.21) will become<br />

transparent mathematically <strong>on</strong>ce we proceed to determine Γ(u, u ′ ) explicitly.<br />

The first step c<strong>on</strong>sists in writing the delta functi<strong>on</strong> as a sum over spherical<br />

harm<strong>on</strong>ics Y m<br />

l (θ, φ),<br />

δ(u, u ′ )<br />

√ g<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> recall the eigenvalue equati<strong>on</strong><br />

≡ δ(θ − θ′ )δ(φ − φ ′ )<br />

R2 sin θ ′<br />

= 1<br />

R2 ∞� m=+l �<br />

Y m<br />

l=0 m=−l<br />

∆ Y m l(l + 1)<br />

l (θ, φ) = −<br />

R2 l (θ, φ)Y m<br />

l<br />

∗ (θ, φ) , (E.22)<br />

Y m<br />

l (θ, φ) . (E.23)


Appendix E: Free energy of a vector field <strong>on</strong> a sphere 168<br />

Up<strong>on</strong> substituting Eq.(E.22) in Eq.(E.21) <str<strong>on</strong>g>and</str<strong>on</strong>g> using the eigenvalue Equati<strong>on</strong> (E.23),<br />

we can write down the Green functi<strong>on</strong> as<br />

Γ(u, u ′ ) = −R 2<br />

∞�<br />

m=+l �<br />

l=1 m=−l<br />

Y m<br />

l<br />

(θ, φ)Y m<br />

l<br />

l(l + 1)<br />

∗ (θ ′ , φ ′ )<br />

. (E.24)<br />

We have used the fact that Y 0<br />

�<br />

1<br />

0 = , <str<strong>on</strong>g>and</str<strong>on</strong>g> used the neutralizing background charge<br />

4π<br />

1<br />

S<br />

in Eq.(E.21) to cancel out the l = 0 diverging mode.<br />

To simplify the series in Eq.(E.24), we exploit the familiar identity [86]<br />

m=+l �<br />

m=−l<br />

l (θ, φ)Y m∗<br />

′ ′ 2l + 1<br />

l (θ , φ ) =<br />

Y m<br />

4π Pl(cos β) , (E.25)<br />

where β is the angle (relative to the center of the sphere) between the directi<strong>on</strong>s {θ, φ}<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> {θ ′ , φ ′ } (see also Eq.(3.36)). Up<strong>on</strong> substituting Eq.(E.25) in Eq.(E.24), we find<br />

Γ(u, u ′ m=+l �<br />

) = −<br />

m=−l<br />

� 1<br />

l<br />

�<br />

1 Pl(cos β)<br />

+ . (E.26)<br />

l + 1 4π<br />

The first term of the sum in Eq.(E.24) can be simplified using the following identity<br />

[87]<br />

l=∞ �<br />

l=l<br />

Pl(cos β)<br />

l<br />

while for the sec<strong>on</strong>d term we substitute<br />

with the result<br />

l=∞ �<br />

l=l<br />

Pl(cos β)<br />

l + 1<br />

Γ(u, u ′ ) = 1<br />

4π ln<br />

� �<br />

1 − cos β<br />

2<br />

� � ��<br />

β<br />

= − ln 1 + sin<br />

2<br />

� � ��<br />

β<br />

− ln sin , (E.27)<br />

2<br />

� �<br />

= ln 1 + sin β<br />

��<br />

2<br />

� � ��<br />

β<br />

− ln sin − 1 , (E.28)<br />

2<br />

+ 1<br />

. (E.29)<br />


Appendix E: Free energy of a vector field <strong>on</strong> a sphere 169<br />

Up<strong>on</strong> dropping additive c<strong>on</strong>stants that do not c<strong>on</strong>tribute to the energy <str<strong>on</strong>g>and</str<strong>on</strong>g> substi-<br />

tuting in Eq.(E.19) we obtain<br />

F = − πK<br />

2p 2<br />

�<br />

i�=j<br />

Nd �<br />

qiqj ln [1 − cos βij] +<br />

i=1<br />

q 2 i Ec , (E.30)<br />

where the phenomenologically determined core energy Ec has been added by h<str<strong>on</strong>g>and</str<strong>on</strong>g><br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> reflect the microscopic physics not captured by our l<strong>on</strong>g-wavelength theory.


Appendix F<br />

<str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

c<strong>on</strong>formal mappings<br />

170


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 171<br />

This Appendix collects a number of results from the theory of complex<br />

variables relevant to the study of liquid crystal textures 1 . The perspective adopted<br />

is to link the liquid crystal elasticity to the intrinsic geometry of the texture by the use<br />

of c<strong>on</strong>formal transformati<strong>on</strong>s. The same method provides an elegant route to finding<br />

the flow lines of simple incompressible fluids in 2D [88] <str<strong>on</strong>g>and</str<strong>on</strong>g> to the exact soluti<strong>on</strong> of<br />

analogous problems in electromagnetism <str<strong>on</strong>g>and</str<strong>on</strong>g> elasticity [45].<br />

Nematic textures in the plane in the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong><br />

can be obtained by solving Laplace equati<strong>on</strong>, which is c<strong>on</strong>formally invariant, <str<strong>on</strong>g>and</str<strong>on</strong>g> in<br />

complex coordinates {z = x + iy, z = x − iy} reads<br />

∂z ∂z α = 0 . (F.1)<br />

Here α(x, y) ≡ α(z, z) is the b<strong>on</strong>d angle that the director n = (cos α, sin α) forms<br />

with respect to a fixed directi<strong>on</strong>, say the real axis �x. The flat space laplacian equati<strong>on</strong><br />

(F.1) is also obtained by minimizing the free energy of a vector field <strong>on</strong> a sphere (see<br />

Appendix A) provided that a ”stereographic projecti<strong>on</strong> gauge” is chosen to carry out<br />

the calculati<strong>on</strong> [12, 89]. The stereographic projecti<strong>on</strong> maps an arbitrary point <strong>on</strong> the<br />

sphere R(θ, φ) to the corresp<strong>on</strong>ding point z = 2R tan � �<br />

θ iφ e in the complex plane.<br />

2<br />

The metric reads [12]<br />

gij =<br />

1<br />

2 � 1 + z¯z<br />

4R2 �<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the comp<strong>on</strong>ents of the gauge field in Eq.(4.3) are given by<br />

Az = Ā¯z = − 1<br />

2iz<br />

⎛ ⎞<br />

⎜0<br />

⎝<br />

1⎟<br />

⎠ , (F.2)<br />

1 0<br />

� z¯z 1 − 4R2 1 + z¯z<br />

4R2 �<br />

. (F.3)<br />

1 An inspiring coverage of complex analysis that emphasizes the geometric viewpoint is given in Ref. [63].


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 172<br />

In this representati<strong>on</strong> of liquid crystal order <strong>on</strong> a sphere, the Frank free energy is<br />

F = K<br />

4<br />

�<br />

d 2 z |∂zα − Az| 2 , (F.4)<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> the corresp<strong>on</strong>ding Euler Lagrange equati<strong>on</strong> is indeed Eq.(F.1), since the diver-<br />

gence of the gauge field is zero (∂zA¯z + ∂¯zAz = 0) [12]. The stereographic projecti<strong>on</strong><br />

provides an example of how c<strong>on</strong>formal transformati<strong>on</strong>s can be used to map physics<br />

<strong>on</strong> an arbitrary curved surface <strong>on</strong>to simpler planar problems. This technique can<br />

also be employed to analyze two dimensi<strong>on</strong>al order <strong>on</strong> surfaces of varying Gaussian<br />

curvature (see Ref.[58, 59]).<br />

A sec<strong>on</strong>d use of c<strong>on</strong>formal mappings is as generators of 2D liquid crystal<br />

textures in bounded geometries or in the presence of defects. C<strong>on</strong>sider an analytic<br />

functi<strong>on</strong> w = f(z) that maps a grid of horiz<strong>on</strong>tal <str<strong>on</strong>g>and</str<strong>on</strong>g> vertical lines in the complex<br />

plane z = x + iy <strong>on</strong>to a family of orthog<strong>on</strong>al curves in the w = u + iv plane that<br />

are respectively streamlines <str<strong>on</strong>g>and</str<strong>on</strong>g> equipotential lines of the corresp<strong>on</strong>ding flow (see<br />

Fig. F.1). Similarly, we can define the inverse functi<strong>on</strong> Ω(z) = f −1 (z) <str<strong>on</strong>g>and</str<strong>on</strong>g> note<br />

that Ω(z) = Φ(x, y) + iΨ(x, y) maps equipotential lines <str<strong>on</strong>g>and</str<strong>on</strong>g> streamlines, given by<br />

the c<strong>on</strong>tour lines of Φ(x, y) <str<strong>on</strong>g>and</str<strong>on</strong>g> Ψ(x, y), into a grid of vertical <str<strong>on</strong>g>and</str<strong>on</strong>g> horiz<strong>on</strong>tal lines<br />

respectively.<br />

The c<strong>on</strong>necti<strong>on</strong> between liquid crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings<br />

rests <strong>on</strong> the following observati<strong>on</strong>: if the director n(z) forms a c<strong>on</strong>stant angle with<br />

respect to the streamlines (or the equipotential lines) of Ω(z), then α(z) automatically<br />

satisfies Eq.(F.1) [63]. In the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>, the complex nematic


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 173<br />

director �n(z) = nx(x, y) + iny(x, y) is given (up to an arbitrary global rotati<strong>on</strong>) by 2<br />

�n(z) = Ω′ (z)<br />

|Ω ′ (z)|<br />

, (F.5)<br />

The b<strong>on</strong>d angle is readily expressed in terms of Ω(z) via the relati<strong>on</strong><br />

α(z) = −ℑm log [Ω ′ (z)] , (F.6)<br />

where ℑm denotes the imaginary part of a complex number.<br />

As an illustrati<strong>on</strong> c<strong>on</strong>sider the simple case of two disclinati<strong>on</strong>s <strong>on</strong> the real<br />

axis at positi<strong>on</strong>s x = ±1 respectively. The nematic director rotates by π <strong>on</strong> a path<br />

encircling <strong>on</strong>ly <strong>on</strong>e defect <str<strong>on</strong>g>and</str<strong>on</strong>g> by 2π <strong>on</strong> a path enclosing both (see Fig. F.1). These<br />

requirements are met by choosing the complex functi<strong>on</strong> Ω(z) = arccos(z) that is<br />

analytic everywhere except for a branch cut <strong>on</strong> the real axis from x = −1 to x =<br />

1. The streamlines <str<strong>on</strong>g>and</str<strong>on</strong>g> equipotential lines are a family of hyperbolas <str<strong>on</strong>g>and</str<strong>on</strong>g> ellipses<br />

with coinciding foci at x = ±1; they corresp<strong>on</strong>d to two distinct nematic textures<br />

dominated by splay <str<strong>on</strong>g>and</str<strong>on</strong>g> bend respectively. The b<strong>on</strong>d angle of the director oriented<br />

al<strong>on</strong>g the streamlines is easily extracted from the argument of the complex vector<br />

field in Eq.(F.5), with the result<br />

�<br />

y<br />

α(x, y) = arctan<br />

2 + 1 − x2 − � (y2 + x2 + 1) 2 − 4x2 �<br />

. (F.7)<br />

2xy<br />

In this simple example, <strong>on</strong>e can explicitly check that the result in Eq.(F.7) is recovered<br />

through the more familiar route of superposing the soluti<strong>on</strong>s corresp<strong>on</strong>ding to the two<br />

isolated defects<br />

α(x, y) = 1<br />

2 arctan<br />

� �<br />

y<br />

+<br />

x − 1<br />

1<br />

2 arctan<br />

� �<br />

y<br />

. (F.8)<br />

x + 1<br />

2 The complex functi<strong>on</strong> Ω ′ (z) denotes the derivative with respect to z of the complex functi<strong>on</strong> Ω =<br />

Φ(x, y) + iΨ(x, y) <str<strong>on</strong>g>and</str<strong>on</strong>g> we define Ω ≡= Φ(x, y) − iΨ(x, y).


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 174<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-2 -1 0 1 2<br />

Figure F.1: (Color <strong>on</strong>line) Splay rich (hyperbolic) <str<strong>on</strong>g>and</str<strong>on</strong>g> bend-rich (ellipsoidal) families<br />

of nematic ”flow” lines generated by two s = 1 disclinati<strong>on</strong>s. The two families of flow<br />

2<br />

lines are the equipotential <str<strong>on</strong>g>and</str<strong>on</strong>g> field lines of a complex functi<strong>on</strong> Ω(z) whose branch<br />

cut is a horiz<strong>on</strong>tal line c<strong>on</strong>necting the two defects.<br />

The applicability of the method of c<strong>on</strong>formal mappings to finding liquid<br />

crystal textures can be justified by means of simple geometric reas<strong>on</strong>ing. We start by<br />

noting that the curl <str<strong>on</strong>g>and</str<strong>on</strong>g> divergence of a two dimensi<strong>on</strong>al vector field v, whose stream-<br />

lines <str<strong>on</strong>g>and</str<strong>on</strong>g> orthog<strong>on</strong>al trajectories are labelled by the subscripts s <str<strong>on</strong>g>and</str<strong>on</strong>g> p respectively,<br />

can be expressed geometrically via the relati<strong>on</strong>s [63]<br />

∂x vx + ∂y vy ≡ ∇ · v = ∂s |v| + κp |v| . (F.9)<br />

∂x vy − ∂y vx ≡ ∇ × v = −∂p |v| + κs |v| , (F.10)<br />

where κs <str<strong>on</strong>g>and</str<strong>on</strong>g> κp are the respective curvatures while ∂s <str<strong>on</strong>g>and</str<strong>on</strong>g> ∂p are the directi<strong>on</strong>al


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 175<br />

derivatives al<strong>on</strong>g the two orthog<strong>on</strong>al families of level curves 3 . For example the black<br />

lines in Fig. F.1 trace the electric field v generated by a uniformly charged plate or<br />

the flow lines of an ideal fluid exiting a slit of width given by the branch cut. Unlike<br />

the liquid crystal director in Eq.(F.5), the magnitude of v is allowed to vary with<br />

positi<strong>on</strong>. By c<strong>on</strong>structi<strong>on</strong>, such a vector field is divergence free <str<strong>on</strong>g>and</str<strong>on</strong>g> curl free, hence<br />

κs = ∂p log |v| , (F.11)<br />

κp = − ∂s log |v| . (F.12)<br />

By combining Equati<strong>on</strong>s (F.11) <str<strong>on</strong>g>and</str<strong>on</strong>g> (F.12), we obtain the geometric c<strong>on</strong>diti<strong>on</strong> that a<br />

family of equipotential lines (or streamlines) needs to satisfy in order to be identified<br />

as level curves of an harm<strong>on</strong>ic potential, namely<br />

∂κp<br />

∂p<br />

+ ∂κs<br />

∂s<br />

= 0 . (F.13)<br />

This c<strong>on</strong>diti<strong>on</strong> is entirely cast in terms of the curvatures of the equipotential lines<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> streamlines without explicit reference to either the potential to be assigned or to<br />

the magnitude of the vector field v(z) [90, 91]. This is a natural language to discuss<br />

orientati<strong>on</strong>al order in liquid crystals since the director �n(z) is a vector field of unit<br />

magnitude.<br />

If we take the liquid crystal director to form a c<strong>on</strong>stant angle with respect<br />

to v(z), the curvatures κs <str<strong>on</strong>g>and</str<strong>on</strong>g> κp in Eq.(F.13) can be simply cast as the directi<strong>on</strong>al<br />

derivatives of α al<strong>on</strong>g the streamlines <str<strong>on</strong>g>and</str<strong>on</strong>g> equipotential lines respectively. In fact,<br />

the curvature of these c<strong>on</strong>tour lines is the rate of change of their directi<strong>on</strong>s which is<br />

3 The directi<strong>on</strong> of increasing p is chosen to make a counterclockwise π<br />

2<br />

angle with v.


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 176<br />

naturally parameterized by α. Hence Eq.(F.13) reduces to<br />

∂2α ∂p2 + ∂2α ∂2s = 0 . (F.14)<br />

The left h<str<strong>on</strong>g>and</str<strong>on</strong>g> side of Eq.(F.17) is the Laplacian of α expressed in terms of orthog<strong>on</strong>al<br />

coordinates al<strong>on</strong>g s <str<strong>on</strong>g>and</str<strong>on</strong>g> p. Since the Laplacian is coordinate independent, Eq.(F.14) is<br />

equivalent to Eq.(F.1) <str<strong>on</strong>g>and</str<strong>on</strong>g> α(x, y) represents (apart from an arbitrary global rotati<strong>on</strong>)<br />

the desired texture that minimizes the Frank free energy when K1 = K3.<br />

As a byproduct, Equati<strong>on</strong>s (F.11) <str<strong>on</strong>g>and</str<strong>on</strong>g> (F.12) can help to visualize how the<br />

elastic energy stored in every porti<strong>on</strong> of the texture of Fig. F.1 is distributed between<br />

bend <str<strong>on</strong>g>and</str<strong>on</strong>g> splay deformati<strong>on</strong>s. For most liquid crystals K3 > K1, so the texture with<br />

director tangent to the streamlines will be energetically favored (black lines in Fig.<br />

F.1). In this case, the full Frank energy density can be rewritten in terms of the local<br />

curvatures of streamlines <str<strong>on</strong>g>and</str<strong>on</strong>g> equipotential lines via the simple relati<strong>on</strong><br />

K3 (∇ × n) 2 + K1 (∇ · n) 2 = K3 κs 2 + K1 κp 2 . (F.15)<br />

The energetically costly deformati<strong>on</strong> involving bend takes place <strong>on</strong>ly around the de-<br />

fects in a regi<strong>on</strong> of radius of the order of their separati<strong>on</strong>. Elsewhere κs is vanishingly<br />

small. In c<strong>on</strong>trast, κp drops off more slowly at large distances like the inverse of<br />

the radius of a circle centered <strong>on</strong> the midpoint between the two disclinati<strong>on</strong>s. Splay<br />

deformati<strong>on</strong>s are present throughout the system but they have a smaller energy cost<br />

K1. The c<strong>on</strong>verse situati<strong>on</strong> occurs if K3 < K1 so that the texture represented by the<br />

red equipotential lines in Fig. F.1 becomes energetically favored.<br />

The curvatures κs <str<strong>on</strong>g>and</str<strong>on</strong>g> κp are respectively the real <str<strong>on</strong>g>and</str<strong>on</strong>g> imaginary parts of<br />

the complex curvature, K(z) = κs +iκp, of the mapping. This quantity can be readily


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 177<br />

derived from the complex potential Ω(z) 4<br />

K(z) ≡ κs + iκp = −i |Ω′ | Ω ′′<br />

Ω ′ 2<br />

. (F.16)<br />

The reader is referred to the mathematical literature [90, 63] for a proof of Eq.(F.16).<br />

The intuitive significance of the complex curvature can be grasped by c<strong>on</strong>sidering<br />

how a c<strong>on</strong>formal mapping f = Ω −1 acts <strong>on</strong> a curve with local curvature κ at a<br />

point in the z = x + iy plane. The curve is mapped <strong>on</strong>to an image curve in the<br />

w = u + iv plane whose curvature κ ′ at the corresp<strong>on</strong>ding point differs from κ.<br />

The curvature κ ′ of the image curve is determined by two mechanisms. Firstly,<br />

the mapping f(z) locally stretches distances by a factor |f ′ (z)|, hence the radius of<br />

curvature of the image curve will be naturally multiplied by this amplifying factor.<br />

The sec<strong>on</strong>d mechanism arises because a c<strong>on</strong>formal mapping can introduce curvature<br />

even if n<strong>on</strong>e was originally present (in the isothermal net) simply by locally twisting<br />

the directi<strong>on</strong> of the isothermal net by an angle equal to arg [f ′ (z)] 5 . The n<strong>on</strong>-analytic<br />

functi<strong>on</strong> K(z) c<strong>on</strong>trols the amount of curvature generated ex-novo by the mapping.<br />

For example, the mapping f(z) = Cos(z) transforms a grid of horiz<strong>on</strong>tal lines (think<br />

of them as a possible directi<strong>on</strong> for the nematic molecules in the defect-free ground<br />

state) into the family of hyperbolae in Fig. F.1 corresp<strong>on</strong>ding to a defected texture<br />

with two +1/2 disclinati<strong>on</strong>s. It is not surprising that the free energy density stored in<br />

the defected texture is simply proporti<strong>on</strong>al (in the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>)<br />

4 For calculati<strong>on</strong>al purposes, it is more c<strong>on</strong>venient to recast Eq.(F.16) in the form κp + iκs = − |Ω ′ | Ω ′′<br />

Ω ′2 .<br />

5 ′<br />

For this reas<strong>on</strong>, the complex derivative f (z) is sometimes called an amplitwist <str<strong>on</strong>g>and</str<strong>on</strong>g> encodes informati<strong>on</strong><br />

<strong>on</strong> the local effect of the mapping.


Appendix F: <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> crystal textures <str<strong>on</strong>g>and</str<strong>on</strong>g> c<strong>on</strong>formal mappings 178<br />

to<br />

� �2 ∂α<br />

+<br />

∂p<br />

� �2 ∂α<br />

= κ<br />

∂s<br />

2 s + κ 2 p = |K(z)| 2 , (F.17)<br />

where the two elastic c<strong>on</strong>stants were set to be equal in Eq.(F.15) <str<strong>on</strong>g>and</str<strong>on</strong>g> the director<br />

n(z) was parameterized in terms of the b<strong>on</strong>d angle α(z). The Frank free energy<br />

is thus proporti<strong>on</strong>al to the complex curvature modulus-squared in analogy with the<br />

Helfrich free energy of a membrane whose derivati<strong>on</strong> rests <strong>on</strong> an higher dimensi<strong>on</strong>al<br />

generalizati<strong>on</strong> of Eq.(F.15).


Appendix G<br />

Vibrati<strong>on</strong>al spectrum of colloidal<br />

molecules<br />

179


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 180<br />

a b<br />

Figure G.1: (a) The ground state of a tetratic phase exhibits eight short disclinati<strong>on</strong><br />

lines located at the vertices of a square antiprism inscribed in the sphere. (b) The<br />

twelve disclinati<strong>on</strong> that characterize hexatic order s sphere lie at the vertices of an<br />

icosahedr<strong>on</strong>.<br />

In this appendix we provide an introducti<strong>on</strong> to the group theoretical treat-<br />

ment of the vibrati<strong>on</strong>al spectrum of colloidal ”molecules”. The more complicated<br />

cases of hexatic (p=6) <str<strong>on</strong>g>and</str<strong>on</strong>g> tetratic (p=4) order are analyzed in some detail (see Fig.<br />

G.1).<br />

The starting point of the group theoretical treatment of the vibrati<strong>on</strong>al<br />

spectrum of colloidal ”molecules” is the observati<strong>on</strong> that the defects displacements<br />

from equilibrium, q (see Eq.(3.39), form the basis of a reducible representati<strong>on</strong> of the<br />

point group of the molecule. If a molecule is acted up<strong>on</strong> by a symmetry operati<strong>on</strong>,<br />

a new c<strong>on</strong>figurati<strong>on</strong> will result in which the displacements of each defect will be<br />

permuted <str<strong>on</strong>g>and</str<strong>on</strong>g> transformed 1 , but inter-defect distances <str<strong>on</strong>g>and</str<strong>on</strong>g> angles will be preserved.<br />

The liquid crystal free energy (in the <strong>on</strong>e Frank c<strong>on</strong>stant approximati<strong>on</strong>) is therefore<br />

invariant under all the operati<strong>on</strong>s of the point group of the colloidal defect array.<br />

The acti<strong>on</strong> of each operati<strong>on</strong> of the group is naturally represented by a dis-<br />

1 Here we take the point of view that the defects themselves are not permuted <strong>on</strong>ly their displacements,<br />

for example defect i may exchange its displacement coordinates qi with defect j.


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 181<br />

Table G.1: Character for the irreducible representati<strong>on</strong>s of the icosahedral point<br />

group together with the character of the twenty-four-dimensi<strong>on</strong>al representati<strong>on</strong> Υ<br />

generated by the defect displacements.<br />

Y E 12C5 12C 2 5 20C3 15C2<br />

A 1 1 1 1 1<br />

F1 3 τ 2 − τ −1 0 −1<br />

F2 3 − τ −1 τ 0 −1<br />

G 4 −1 −1 1 0<br />

H 5 0 0 −1 1<br />

Y 24 2τ −1 −2τ 0 0<br />

tinct 2N × 2N matrix (N is the number of defects) that relates the new <str<strong>on</strong>g>and</str<strong>on</strong>g> old<br />

defect positi<strong>on</strong>s. This representati<strong>on</strong> can be completely reduced by choosing a set<br />

of symmetry-related normal coordinates that are obtained from the original <strong>on</strong>es by<br />

means of a linear transformati<strong>on</strong>. When normal coordinates are used, the matrixes<br />

representing the acti<strong>on</strong> of the symmetry group can be brought in block diag<strong>on</strong>al form<br />

simultaneously. Energetically degenerate linear combinati<strong>on</strong> of the original coordi-<br />

nates form the smallest sets invariant under applicati<strong>on</strong> of any symmetry operati<strong>on</strong> of<br />

the group. The members of any <strong>on</strong>e set generate an irreducible representati<strong>on</strong> of the<br />

group. For each point group there is <strong>on</strong>ly a small number of inequivalent irreducible<br />

representati<strong>on</strong>s generally classified by the characters of their transformati<strong>on</strong>s. The<br />

characters of the transformati<strong>on</strong>s are simply defined as the traces of the matrices cor-<br />

resp<strong>on</strong>ding to each symmetry operati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> they are c<strong>on</strong>veniently tabulated in most


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 182<br />

texts of group theory [64, 65] (see Tables 3.1-G.2 for the character tables relevant<br />

to the tetrahedral, icosahedral <str<strong>on</strong>g>and</str<strong>on</strong>g> twisted-cube shaped distributi<strong>on</strong>s of defects <strong>on</strong> a<br />

sphere).<br />

The task of finding the number of degenerate eigenmodes, n (γ) , with a given<br />

symmetry (labelled by γ) reduces to counting how many times the corresp<strong>on</strong>ding irre-<br />

ducible representati<strong>on</strong> appears in a reducible representati<strong>on</strong>. Note that the characters<br />

of the original representati<strong>on</strong> are the same as the <strong>on</strong>es of the completely reduced <strong>on</strong>e<br />

since the two differ <strong>on</strong>ly by a change of coordinates which preserves the trace. Thus,<br />

the character χR of the completely reduced representati<strong>on</strong> will be the sum of the<br />

characters of the various irreducible representati<strong>on</strong>s that it c<strong>on</strong>tains<br />

where χ (γ)<br />

R<br />

χR = �<br />

γ<br />

n (γ) χ (γ)<br />

R , (G.1)<br />

labels the character of the symmetry operati<strong>on</strong> R in the irreducible rep-<br />

resentati<strong>on</strong> γ. By appealing to the orthog<strong>on</strong>ality of the characters <strong>on</strong>e can write an<br />

expressi<strong>on</strong> for n (γ) in analogy with the familiar expressi<strong>on</strong> for the comp<strong>on</strong>ent of a<br />

vector al<strong>on</strong>g a given basis axis [64, 65]<br />

n (γ) = 1<br />

g<br />

�<br />

R<br />

χ (γ)∗<br />

R χR , (G.2)<br />

where g is the number of the symmetry operati<strong>on</strong>s in the group <str<strong>on</strong>g>and</str<strong>on</strong>g> χR is the character<br />

of the completely reduced representati<strong>on</strong>. Eq.(G.2) is equivalent to Eq.(3.43) quoted<br />

in the main text, as l<strong>on</strong>g as the sum over the group elements R is replaced by a<br />

weighted sum over the classes in the group since the characters of group elements in<br />

the same class are equal.


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 183<br />

P<br />

σ<br />

Figure G.2: Schematic illustrati<strong>on</strong> of the inversi<strong>on</strong> through an equatorial plane of<br />

symmetry (shaded circle bisecting a sphere) for a reflecti<strong>on</strong> σ. The l<strong>on</strong>gitudinal<br />

displacement of the defect is unchanged while the directi<strong>on</strong> of the latitudinal <strong>on</strong>e is<br />

inverted.<br />

We now adopt the analysis of Ref.[64, 65] to provide a set of rules that<br />

produce the characters χR of the reducible representati<strong>on</strong> generated by the coordi-<br />

nates without working out the full form of the transformati<strong>on</strong> matrices. There are<br />

two key points to notice. First <strong>on</strong>ly the defects located <strong>on</strong> a symmetry axis or plane<br />

c<strong>on</strong>tribute to χR the trace of the transformati<strong>on</strong> matrix; defects whose displacements<br />

are instead interchanged or permuted by the symmetry operati<strong>on</strong> c<strong>on</strong>tribute <strong>on</strong>ly to<br />

the n<strong>on</strong>-diag<strong>on</strong>al terms of the matrix <str<strong>on</strong>g>and</str<strong>on</strong>g> hence can be ignored in determining the<br />

character χR. Sec<strong>on</strong>d the directi<strong>on</strong>s al<strong>on</strong>g which the displacements from equilibrium<br />

are measured can be chosen freely since the trace is invariant up<strong>on</strong> coordinate trans-<br />

formati<strong>on</strong>s. It is generally c<strong>on</strong>venient to choose them so that <strong>on</strong>ly <strong>on</strong>e of the two<br />

displacements comp<strong>on</strong>ents is affected by the symmetry operati<strong>on</strong>.<br />

P


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 184<br />

As a simple example, c<strong>on</strong>sider a defect lying <strong>on</strong> a reflecti<strong>on</strong> plane (see Fig.<br />

G.2). The displacement vectors before <str<strong>on</strong>g>and</str<strong>on</strong>g> after the symmetry operati<strong>on</strong> is applied are<br />

{δθ, δφ} <str<strong>on</strong>g>and</str<strong>on</strong>g> {δθ, −δφ} respectively. The resulting c<strong>on</strong>tributi<strong>on</strong> to the character from<br />

a single defect is thus 1 − 1 = 0. Inversi<strong>on</strong>s through a center of symmetry (coinciding<br />

with the center of the sphere) have a vanishing c<strong>on</strong>tributi<strong>on</strong> to χR because there are<br />

no defects there that can c<strong>on</strong>tribute to χR.<br />

A rotati<strong>on</strong> C k n by an angle 2πk<br />

n<br />

through an n-fold axis of symmetry <strong>on</strong> which<br />

the defect lies leads to the following transformati<strong>on</strong> laws for the l<strong>on</strong>gitudinal <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

latitudinal displacements<br />

⎛ ⎞<br />

⎛<br />

⎜ δθi<br />

⎜<br />

⎝<br />

′<br />

δφi ′<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎟ = ⎜<br />

⎟ ⎜<br />

⎠ ⎝<br />

cos � 2πk<br />

n<br />

sin � �<br />

2πk<br />

n<br />

� � �<br />

2πk − sin n<br />

cos � �<br />

2πk<br />

n<br />

⎞ ⎛<br />

δφi<br />

⎞<br />

⎟ ⎜ δθi ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ ⎟<br />

⎟ ⎜ ⎟<br />

⎠ ⎝ ⎠<br />

. (G.3)<br />

We measure displacements using polar coordinates with respect to the symmetry<br />

axis; the prime denotes the orthog<strong>on</strong>al displacements after the symmetry operati<strong>on</strong><br />

C k n is applied. Inspecti<strong>on</strong> of Eq.(G.3) shows that the c<strong>on</strong>tributi<strong>on</strong> from C k n to the<br />

character χR is equal to 2 cos � �<br />

2πk times the number of defects lying <strong>on</strong> the axis<br />

n<br />

of rotati<strong>on</strong>. On the other h<str<strong>on</strong>g>and</str<strong>on</strong>g> the c<strong>on</strong>tributi<strong>on</strong> to χR from the improper rotati<strong>on</strong><br />

S k n is zero. To see this note that the symmetry operati<strong>on</strong> S k n is a rotary reflecti<strong>on</strong><br />

achieved by performing a successive rotati<strong>on</strong> through an (alternating) axis followed<br />

by a reflecti<strong>on</strong> in the plane perpendicular to the axis k times. An example of a<br />

molecule possessing the symmetry operati<strong>on</strong> S4 is methane, CH4, with the carb<strong>on</strong><br />

atom lying at the intersecti<strong>on</strong> between an alternating axis <str<strong>on</strong>g>and</str<strong>on</strong>g> the reflecti<strong>on</strong> plane


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 185<br />

3 . The tetrahedral defect c<strong>on</strong>figurati<strong>on</strong> c<strong>on</strong>sidered in this paper does not posses any<br />

defect at the positi<strong>on</strong> occupied by the carb<strong>on</strong> atom of the methane molecule. More<br />

generally, the possibility of having a defect whose equilibrium positi<strong>on</strong> is unchanged<br />

by the rotary reflecti<strong>on</strong> is ruled out because such defect would have to lie off the<br />

spherical surface at the intersecti<strong>on</strong> between the alternating axis <str<strong>on</strong>g>and</str<strong>on</strong>g> the plane of<br />

reflecti<strong>on</strong>.<br />

To sum up, each of the characters, χR, of the completely reduced repre-<br />

sentati<strong>on</strong> formed by the displacement coordinates is given by the number of atoms<br />

whose equilibrium positi<strong>on</strong>s are not changed by the symmetry operati<strong>on</strong> R times its<br />

fundamental character as derived in the previous paragraphs 4 . The resulting charac-<br />

ters for the tetrahedral, icosahedral <str<strong>on</strong>g>and</str<strong>on</strong>g> tilted cube defects c<strong>on</strong>figurati<strong>on</strong>s are listed<br />

in Tables 3.1-G.2.<br />

Up<strong>on</strong> using Eq.(3.43) <str<strong>on</strong>g>and</str<strong>on</strong>g> the character table G.1 we can decompose the<br />

24 dimensi<strong>on</strong>al representati<strong>on</strong>, Y , formed by the displacements from an icosahedral<br />

equilibrium c<strong>on</strong>figurati<strong>on</strong> into irreducible representati<strong>on</strong>s. The result reads<br />

Y = 2H + 2G + 2F1 . (G.4)<br />

The three rigid body rotati<strong>on</strong>s corresp<strong>on</strong>d to <strong>on</strong>e of the two triplets in F1 while the<br />

remaining 21 independent normal coordinates form energetically degenerate multi-<br />

plets with the following degeneracy factors: 2 quintets, 2 quartets <str<strong>on</strong>g>and</str<strong>on</strong>g> 1 triplet. This<br />

analysis is c<strong>on</strong>firmed up<strong>on</strong> direct diag<strong>on</strong>alizati<strong>on</strong> of the representati<strong>on</strong> Y which leads<br />

3 See Fig. 5-2 in Ref. [64]<br />

4 Similar results that apply to unc<strong>on</strong>strained molecules whose atoms have three dimensi<strong>on</strong>al displacements<br />

are listed in Table 6-1 of Ref. [64].


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 186<br />

Table G.2: Character for the irreducible representati<strong>on</strong>s of the tilted cube point group<br />

together with the character of the sixteen-dimensi<strong>on</strong>al representati<strong>on</strong> Ξ generated by<br />

the defect displacements.<br />

D4d E 2S8 2C4 2S 3 8 C2 4C 1 2 4σ2<br />

A1 1 1 1 1 1 1 1<br />

A2 1 1 1 1 1 −1 −1<br />

B1 1 −1 1 −1 1 1 −1<br />

B2 1 −1 1 −1 1 −1 1<br />

E1 2 √ 2 0 − √ 2 −2 0 0<br />

E2 2 0 −2 0 2 0 0<br />

E3 2 − √ 2 0<br />

√ 2 −2 0 0<br />

Ξ 16 0 0 0 0 0 0


Appendix G: Vibrati<strong>on</strong>al spectrum of colloidal molecules 187<br />

the 21 n<strong>on</strong>-vanishing eigenvalues λi, with the multiplicities shown bold in parenthesis<br />

λi = {0.87× (5), 0.09× (5),<br />

0.74× (4), 0.22× (4), 0.96× (3)} . (G.5)<br />

Note that the normal modes in the sec<strong>on</strong>d quintet are much ”softer” than the rest.<br />

A similar analysis applied to the twisted cube c<strong>on</strong>figurati<strong>on</strong> of defects leads<br />

to the following decompositi<strong>on</strong> of the (defects’ displacements) representati<strong>on</strong>, Ξ,<br />

Ξ = 2E1 + 2E2 + 2E3 + A1 + A2 + B1 + B2 . (G.6)<br />

where the three rigid body rotati<strong>on</strong>s are c<strong>on</strong>tained in <strong>on</strong>e of the two doublets E3 <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

in the singlet A2, which leaves five doublets <str<strong>on</strong>g>and</str<strong>on</strong>g> three singlets for the eigenvalues<br />

ζi. Direct diag<strong>on</strong>alizati<strong>on</strong> of Ξ leads four doublets, <strong>on</strong>e triplet <str<strong>on</strong>g>and</str<strong>on</strong>g> two singlets of<br />

n<strong>on</strong>-vanishing eigenvalues,<br />

ζi = {1.37× (3), 1.31× (2), 0.89× (2),<br />

0.47× (2), 0.06× (2), 0.11, 1.26} (G.7)<br />

The discrepancy between the degeneracies found by direct diag<strong>on</strong>alizati<strong>on</strong> <strong>on</strong> <strong>on</strong>e<br />

h<str<strong>on</strong>g>and</str<strong>on</strong>g> <str<strong>on</strong>g>and</str<strong>on</strong>g> group theory <strong>on</strong> the other is caused by an accidental symmetry of the<br />

potential energy of the tilted-cube arrangement of defects. Hence the first triplet is<br />

to be interpreted as the missing doublet <str<strong>on</strong>g>and</str<strong>on</strong>g> singlet that happen to have the same<br />

energy even if there is no symmetry reas<strong>on</strong>s to expect so. The modes in the last<br />

doublet of Eq.(G.7) are the softest.


Appendix H<br />

Perturbati<strong>on</strong> theory of curved<br />

crystals<br />

188


Appendix H: Perturbati<strong>on</strong> theory of curved crystals 189<br />

The starting point of our perturbative analysis of curved crystalline order<br />

is Eq.(5.5) which incorporates Gaussian curvature by adding an extra source to the<br />

defect term but it is n<strong>on</strong>etheless written using a flat space metric. To underscore the<br />

subtleties involved we write a covariant generalizati<strong>on</strong> of the force balance equati<strong>on</strong>,<br />

∂iσij(�x) = 0, [92]<br />

1 ∂<br />

√<br />

g ∂xi (√gσ i j) − Γ k jiσ i k = 0 . (H.1)<br />

where g is the determinant of the metric tensor gij <str<strong>on</strong>g>and</str<strong>on</strong>g> Γ k ji is the Christoffel symbol.<br />

Eq.(H.1) can be c<strong>on</strong>cisely written as σ i ;j;i = 0, where the semicol<strong>on</strong>s indicate the<br />

covariant derivatives Di [92].<br />

Strictly speaking, this equati<strong>on</strong> cannot be solved by using the flat space<br />

trick of writing the stress tensor in terms of the Airy functi<strong>on</strong> because of a distinctive<br />

property of curved space: torsi<strong>on</strong> [93, 40]. An arbitrary vector � T parallel transported<br />

around a closed loop <strong>on</strong> a surface of n<strong>on</strong> vanishing Gaussian curvature is rotated from<br />

its original orientati<strong>on</strong> 1 . In differential form, this c<strong>on</strong>straint reads<br />

[Di, Dj] Tk = G(�x) γij γ m k Tm, (H.2)<br />

where γij = ɛij<br />

√g denotes the antisymmetric tensor [40]. Up<strong>on</strong> substituting Tm =<br />

γm n ∂n χ(�x) into Eq.(H.2), we see that the covariant analogue of Eq.(5.3) does not<br />

hold because the commutator of covariant derivatives (known as the torsi<strong>on</strong> tensor)<br />

does not vanish.<br />

We can n<strong>on</strong>etheless make progress by noting first that the Gaussian cur-<br />

vature of a bump in Eq.(H.2) is proporti<strong>on</strong>al to α2<br />

x2 , so that Eq. (5.3) is in fact<br />

0<br />

1 In fact, the net effect of parallel transport is equivalent to m<strong>on</strong>itoring the change of � T first in <strong>on</strong>e<br />

directi<strong>on</strong> then in the other followed by subtracting changes in the reverse order.


Appendix H: Perturbati<strong>on</strong> theory of curved crystals 190<br />

0.01<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

F (α) 0<br />

2 Yr0 0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

Figure H.1: The geometric frustrati<strong>on</strong> energy F0(α) (Eq. (5.3)) is plotted versus α for<br />

x0/R = 10/80 (○) <str<strong>on</strong>g>and</str<strong>on</strong>g> 10/40 (✷). Solid (x0/R = 10/80) <str<strong>on</strong>g>and</str<strong>on</strong>g> dashed (x0/R = 10/40)<br />

lines indicate plots of Eq. (5.3) using Y = 2 √ k [80], including the finite size correcti<strong>on</strong>s<br />

3<br />

discussed in Appedix I for these two c<strong>on</strong>diti<strong>on</strong>s, respectively.<br />

approximately correct for small α. We first c<strong>on</strong>sider the case S(�x) = 0, for which<br />

Eq.(5.5) reduces to <strong>on</strong>e of the two celebrated Foppl-v<strong>on</strong> Karman equati<strong>on</strong>s describing<br />

slightly deformed thin plates [14]. For a frozen substrate, the sec<strong>on</strong>d Foppl-v<strong>on</strong> Kar-<br />

man equati<strong>on</strong> (arising from the variati<strong>on</strong> of the normal coordinate h(x,y)) determines<br />

the adhesi<strong>on</strong> pressure (normal to the surface) which is needed to c<strong>on</strong>strain the 2D<br />

solid to the curved substrate [14]. The Foppl-v<strong>on</strong> Karman analysis rests <strong>on</strong> a c<strong>on</strong>-<br />

sistent perturbati<strong>on</strong> theory in α <str<strong>on</strong>g>and</str<strong>on</strong>g> predicts that χ(�x) is O(α 2 ), as can be seen by<br />

applying dimensi<strong>on</strong>al analysis to Eq.(5.5) with S(�x) = 0. To this order, the commu-<br />

α


Appendix H: Perturbati<strong>on</strong> theory of curved crystals 191<br />

tator in Eq.(H.2) indeed vanishes (the leading order correcti<strong>on</strong>s are O(α 4 )) <str<strong>on</strong>g>and</str<strong>on</strong>g> <strong>on</strong>e<br />

is justified in expressing σij(�x) in terms of an Airy functi<strong>on</strong> χ(�x).<br />

The situati<strong>on</strong> is more delicate in the presence of defects because χ(�x) no<br />

l<strong>on</strong>ger vanishes as α → 0 but tends instead to the flat space form (see Ref.[2] for<br />

explicit results). Dimensi<strong>on</strong>al analysis reveals that the commutator in Eq.(H.2) is<br />

now proporti<strong>on</strong>al to α2 Y<br />

x0<br />

b<br />

x0 for dislocati<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> to α2 Y<br />

x0<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> impurities where correcti<strong>on</strong>s of order ln � R<br />

a<br />

Ω<br />

x 2 0<br />

for vacancies, interstitials<br />

� are ignored. Hence, the commutator<br />

in Eq.(H.2), set to zero in the derivati<strong>on</strong> of Eq.(5.5), appears to be of the same order<br />

in α as the curvature correcti<strong>on</strong>s to χ(�x) that we wish to calculate. The commutator<br />

is still small, however, in the c<strong>on</strong>tinuum limit x0 ≫ a. The use of Eq.(5.5) to study<br />

isolated disclinati<strong>on</strong>s to leading order in α is more difficult to justify because in this<br />

case the commutator can be as large as α2 Y<br />

x0<br />

R<br />

x0<br />

. However, such an investigati<strong>on</strong> <strong>on</strong><br />

a surface with the topology of the plane is of limited interest, in view of the large<br />

energy cost of isolated disclinati<strong>on</strong>s (quadratic in R). In this work, we c<strong>on</strong>centrate<br />

primarily <strong>on</strong> the physics of dislocati<strong>on</strong>s, vacancies, interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> impurities, which<br />

(as c<strong>on</strong>firmed by our simulati<strong>on</strong>s) is adequately described by the formalism embodied<br />

in Equati<strong>on</strong>s (C.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> (5.5). Our analytical results are compared whenever possible<br />

to numerical minimizati<strong>on</strong>s of an harm<strong>on</strong>ic lattice model; these computati<strong>on</strong>s cor-<br />

roborate our qualitative c<strong>on</strong>clusi<strong>on</strong>s even bey<strong>on</strong>d the limits α ≪ 1 <str<strong>on</strong>g>and</str<strong>on</strong>g> a ≪ x0 for<br />

which Equati<strong>on</strong>s (C.2) <str<strong>on</strong>g>and</str<strong>on</strong>g> (5.5) are strictly valid. For example, Fig. H.1 illustrates<br />

how our theoretical predicti<strong>on</strong>s of the frustrati<strong>on</strong> energy F0, based <strong>on</strong> Eq.(5.3), fit<br />

numerical data for increasingly larger values of α. Similarly, rather good agreement


Appendix H: Perturbati<strong>on</strong> theory of curved crystals 192<br />

0<br />

−0.01<br />

−0.02<br />

−0.03<br />

−0.04<br />

−0.05<br />

−0.06<br />

−0.07<br />

D(r,π/2)<br />

Ybr 0<br />

−0.08<br />

0 0.5 1 1.5 2 2.5 3 3.5<br />

Figure H.2: The dislocati<strong>on</strong> potential D(r, π/2) in Eq.(5.11) is plotted as a c<strong>on</strong>tinuous<br />

line for a Gaussian bump parameterized by α = 1. Open symbols represent the<br />

numerical minimizati<strong>on</strong> of a fixed c<strong>on</strong>nectivity harm<strong>on</strong>ic model for which the separati<strong>on</strong><br />

of the two 5-7 disclinati<strong>on</strong>s is fixed while sliding the dislocati<strong>on</strong> as a whole<br />

radially with respect to the bump: θ = π/2, R/x0 =4 (blue), 8 (red); x0/a = 10 (○),<br />

15 (✷), 20 (△), 30 (♦). The R-dependent finite size correcti<strong>on</strong>s of Appendix I are<br />

included.<br />

between theoretical predicti<strong>on</strong>s <str<strong>on</strong>g>and</str<strong>on</strong>g> numerics is obtained for the dislocati<strong>on</strong> potential<br />

D(r, θ) even if α = 1, as illustrated by Fig. H.2.<br />

r


Appendix I<br />

Curvature induced finite size<br />

effects<br />

193


Appendix I: Curvature induced finite size effects 194<br />

In this appendix, we discuss how the geometric potentials derived above for<br />

an infinite system are modified by the presence of a circular boundary of radius R <strong>on</strong><br />

which free boundary c<strong>on</strong>diti<strong>on</strong>s apply<br />

niσij(R) = 1<br />

R<br />

� �<br />

∂χ<br />

= 0 . (I.1)<br />

∂r r=R<br />

where ni is the unit normal to the circumference of the system. We first determine<br />

the Airy functi<strong>on</strong> χ G (r) that describes elastic deformati<strong>on</strong>s caused by the Gaussian<br />

curvature G(r) <strong>on</strong>ly, without any c<strong>on</strong>tributi<strong>on</strong>s from the defects. As explained in<br />

Secti<strong>on</strong> 5.3, <strong>on</strong>ce χ G (r) is known the energy of geometric frustrati<strong>on</strong> can be easily<br />

calculated. We start by fixing the harm<strong>on</strong>ic functi<strong>on</strong> HR(�x) introduced in Eq.(5.9)<br />

up<strong>on</strong> using the boundary c<strong>on</strong>diti<strong>on</strong> of Eq.(I.1). Note that by azimuthal symmetry,<br />

the Gaussian curvature G(r) <str<strong>on</strong>g>and</str<strong>on</strong>g> hence HR(r) are c<strong>on</strong>stant <strong>on</strong> the circular boundary.<br />

This allows to c<strong>on</strong>clude that the harm<strong>on</strong>ic functi<strong>on</strong> is c<strong>on</strong>stant everywhere; we denote<br />

this c<strong>on</strong>stant by HR. The subscript indicates that the c<strong>on</strong>stant HR is a functi<strong>on</strong> of<br />

the system size R that we can explicitly determine by integrating<br />

over the area of the circular disk, with the result<br />

1<br />

Y ∆χG (r) = −V (r) + HR , (I.2)<br />

�<br />

r ∂χ<br />

�R = Y<br />

∂r 0<br />

� R<br />

0<br />

[HR − V (r)] rdr . (I.3)<br />

Up<strong>on</strong> substituting Equati<strong>on</strong>s (I.1) <str<strong>on</strong>g>and</str<strong>on</strong>g> (5.3) into Eq.(I.3), we obtain by integrati<strong>on</strong><br />

HR = α2 x 2 0<br />

4R 2<br />

�<br />

e −<br />

“ ” �<br />

2<br />

R<br />

x0 − 1<br />

, (I.4)


Appendix I: Curvature induced finite size effects 195<br />

which insures that the forces <strong>on</strong> the boundary vanish. Note that σ G rφ<br />

1 = r2 ∂2χG ∂φ2 0 as a c<strong>on</strong>sequence of azimuthal symmetry, so the sec<strong>on</strong>d boundary c<strong>on</strong>diti<strong>on</strong> is<br />

automatically satisfied. For large system sizes R ≫ x0, HR � − α2 x 2 0<br />

4R 2 .<br />

Up<strong>on</strong> using the general definiti<strong>on</strong> of the Airy functi<strong>on</strong>, we obtain<br />

σ G kk(r) = ∆χ G (r) = −Y V (r) − α2 Y x 2 0<br />

4R 2 . (I.5)<br />

Substituti<strong>on</strong> of Eq.(I.5) in Eq.(C.2) leads an estimate of the stretching energy, F0, of<br />

the defect-free crystal that accounts for finite size effects, namely<br />

F0 � Y<br />

� �<br />

dA V (r) +<br />

2<br />

α2x2 0<br />

4R2 �2 . (I.6)<br />

The graph in Fig. H.1 is a numerical plot of F0 versus α for R<br />

x0<br />

=<br />

equal to 8 <str<strong>on</strong>g>and</str<strong>on</strong>g> 4 that<br />

corroborates the results of Eq.(I.6). This result is obtained from the <strong>on</strong>e quoted in<br />

the main text by simply performing the substituti<strong>on</strong> V (r) → V (r) − HR in Eq.(5.3).<br />

Note that the form of HR was determined by solving for χ G (r) in Eq.(I.2). Strictly<br />

speaking, in the presence of defects <strong>on</strong>e should solve the full Eq.(5.5) that accounts for<br />

the presence of defects by means of an extra source term. The soluti<strong>on</strong> χ(�x) would<br />

not be azimuthally symmetric especially when defects are located well away from<br />

the center of the bump. However, as detailed in Secti<strong>on</strong>s 5.4 <str<strong>on</strong>g>and</str<strong>on</strong>g> 5.7, the geometric<br />

forces experienced by the defects at different locati<strong>on</strong>s <strong>on</strong> the bump can be easily<br />

calculated <strong>on</strong>ce σ G ij(r) is known without solving the full biharm<strong>on</strong>ic equati<strong>on</strong> provided<br />

that x0 ≪ R. The leading finite size correcti<strong>on</strong> to σ G ij(r) was indeed calculated in<br />

Eq.(I.5).<br />

We can thus make progress in the calculati<strong>on</strong> of the defect potentials in the<br />

presence of the boundary by simply letting V (r) → V (r) − HR in Eq.(5.8) that now


Appendix I: Curvature induced finite size effects 196<br />

reads<br />

ζ(�x α ) = −Y<br />

�<br />

dA ′ S(�x ′ )<br />

�<br />

dA 1<br />

∆�x�x ′<br />

[V (�x) − HR] . (I.7)<br />

The result for the geometric potential of dislocati<strong>on</strong>s D(r, θ) with finite size effects<br />

(see Secti<strong>on</strong> 5.4) reads<br />

�<br />

D(r, θ) = Y bi ɛij ∂j dA ′<br />

�<br />

1<br />

V (r<br />

∆�r r �′ ′ ) + α2x2 0<br />

4R2 �<br />

α<br />

≈ Y b x0<br />

2<br />

��<br />

e<br />

sin θ<br />

8 −r2<br />

� �<br />

− 1<br />

� �<br />

x0<br />

2<br />

+ r<br />

r R<br />

. (I.8)<br />

The extra boundary term in D(r, θ) can be viewed as the geometric potential in<br />

an infinite system (let R → ∞ in Eq.(5.11)) evaluated at the positi<strong>on</strong> of an image<br />

dislocati<strong>on</strong> located at r ′ = R2<br />

r <str<strong>on</strong>g>and</str<strong>on</strong>g> with Burger vector � b ′ = � −b. Thus, the disloca-<br />

ti<strong>on</strong> interacts with the curvature directly <str<strong>on</strong>g>and</str<strong>on</strong>g> via its image. This choice of image<br />

insures that the stress normal to the boundary, σrr = 1 ∂χ<br />

, vanishes, similar to the<br />

r ∂r<br />

normal comp<strong>on</strong>ent of the magnetic field generated by a wire parallel to the axis of<br />

a surrounding cylindrical wall of infinite magnetic permeability [94]. The curvature<br />

induced boundary term included in Eq.(I.8) has no analogue in flat space but its<br />

effect is noticed by examining the results of our curved space simulati<strong>on</strong>s presented<br />

in Fig. 5.1 for two different choices of R<br />

x0<br />

equal to 8 <str<strong>on</strong>g>and</str<strong>on</strong>g> 4 <str<strong>on</strong>g>and</str<strong>on</strong>g> aspect ratio α = 0.5.<br />

The comparis<strong>on</strong> between numerics <str<strong>on</strong>g>and</str<strong>on</strong>g> the analytical expressi<strong>on</strong> derived in Eq.(I.8)<br />

appears to c<strong>on</strong>firm the validity of our approximati<strong>on</strong> scheme.<br />

We should stress, the electromagnetic analogy does not guarantee that the<br />

sec<strong>on</strong>d boundary c<strong>on</strong>diti<strong>on</strong> σrφ = 0 is satisfied when the presence of a dislocati<strong>on</strong><br />

breaks the azimuthal symmetry. Moreover, as the dislocati<strong>on</strong> approaches the bound-


Appendix I: Curvature induced finite size effects 197<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

D(r,π/2)<br />

Ybr 0<br />

−4<br />

0 0.5 1 1.5 2 2.5 3 3.5 4<br />

Figure I.1: The dislocati<strong>on</strong> potential D(r, π/2) is plotted versus the scaled coordinate<br />

r = x/x0, x0 = 10a, R = 40a for α = 0 (○), 0.05 (△), 0.1 (∇) <str<strong>on</strong>g>and</str<strong>on</strong>g> 0.2 (✷). The<br />

numerical energy is scaled by 10 −3 x0, <str<strong>on</strong>g>and</str<strong>on</strong>g> shifted so that D(0) ≡ 0. Notice that the<br />

minimum is gradually lost for α > αS ≈ 0.1.<br />

ary it will be str<strong>on</strong>gly attracted to it, an effect that is not captured by our formalism<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> that requires solving the full boundary value problem. A systematic treatment<br />

of boundary effects is rather c<strong>on</strong>voluted even in flat space <str<strong>on</strong>g>and</str<strong>on</strong>g> it is outside the scope<br />

of this work. We refer the interested reader to Ref.[95]. Here we will <strong>on</strong>ly provide<br />

some estimates of how the attracti<strong>on</strong> to the boundary when R is finite wins out over<br />

the geometric force in the flat space limit α → 0. Below a critical aspect ratio αs the<br />

minimum of the geometric potential D(r, θ) is lost, as illustrated numerically in Fig.<br />

I.1. The aspect ratio corresp<strong>on</strong>ding to this spinodal point can be obtained by simple<br />

r


Appendix I: Curvature induced finite size effects 198<br />

dimensi<strong>on</strong>al analysis by a force or equivalently an energy balance. The geometric<br />

potential scales like Y bx0 whereas the attracti<strong>on</strong> to the boundary is a functi<strong>on</strong> of the<br />

dimensi<strong>on</strong>less variable � �<br />

�x 2 2 <str<strong>on</strong>g>and</str<strong>on</strong>g> is multiplied by Y b . This choice insures that the<br />

R<br />

boundary interacti<strong>on</strong> is unchanged if the positi<strong>on</strong> of the defects is flipped �x → −�x<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> suggests that the energy of the boundary interacti<strong>on</strong> evaluated at the minimum<br />

of D(r, θ) is given by Y b2 x 2 0<br />

R 2<br />

(to leading order in a perturbati<strong>on</strong> expansi<strong>on</strong> in the small<br />

parameter x0<br />

R ). This leads to an estimate of the spinoidal aspect ratio αs ∼ √ bx0<br />

R<br />

which agrees with our numerical results.


Appendix J<br />

Numerical Methods<br />

199


Appendix J: Numerical Methods 200<br />

We have complemented our analytical studies with numerical minimizati<strong>on</strong>s<br />

of the in-plane stretching energy of a triangular lattice of points c<strong>on</strong>nected by springs<br />

draped over the Gaussian bump described in the text. The discrete stretching energy<br />

for a set of points at positi<strong>on</strong>s �rµ is defined by<br />

F discrete = 1<br />

4<br />

�<br />

kµ,ν(rµν − a) 2 , (J.1)<br />

µ,ν<br />

where rµ,ν = |�rµ − �rν|, �rµ = (xµ, yµ, h(xµ, yµ)), <str<strong>on</strong>g>and</str<strong>on</strong>g> a is the lattice spacing. The<br />

height functi<strong>on</strong> h(x, y) defines the fixed topography of the system <str<strong>on</strong>g>and</str<strong>on</strong>g> is given by<br />

h(x, y) = αx0e −(x2 +y 2 ) 2 /2x 2 0. The spring c<strong>on</strong>stant matrix is kµ,ν = knµ,ν, where the<br />

c<strong>on</strong>nectivity matrix nµ,ν specifies that the underlying lattice is triangular, with a fully<br />

coordinated particle having 6 nearest neighbors (n.n.)<br />

⎧<br />

⎪⎨ 1 if µ, ν n.n.<br />

nµ,ν =<br />

⎪⎩ 0 otherwise<br />

. (J.2)<br />

Defects are introduced into the lattice by changing the number of nearest neighbors<br />

from 6 to 5/7 for +/- disclinati<strong>on</strong>s, respectively. Dislocati<strong>on</strong>s, interstitials <str<strong>on</strong>g>and</str<strong>on</strong>g> vacan-<br />

cies are composed of specific c<strong>on</strong>figurati<strong>on</strong>s of +/- disclinati<strong>on</strong>s. By coarse-graining<br />

Eq.(J.1) [80], we can make an explicit c<strong>on</strong>necti<strong>on</strong> to the macroscopic energy formula<br />

(5.1) with Young’s modulus, Y = 2<br />

√ 3 k <str<strong>on</strong>g>and</str<strong>on</strong>g> Poiss<strong>on</strong> ratio, σ = 1/3, from which the<br />

Lamé coefficients λ <str<strong>on</strong>g>and</str<strong>on</strong>g> µ can be determined [14]. In this work, we use units of energy<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> length such that a = 1 <str<strong>on</strong>g>and</str<strong>on</strong>g> k = 1.<br />

Numerical minimizati<strong>on</strong>s of Eq.(J.1) are performed for circular patches of<br />

triangular lattice draped over a bump, with frozen-in defect c<strong>on</strong>figurati<strong>on</strong>s. Thus<br />

every point in Figure 5.1 represents a separate energy minimizati<strong>on</strong>. To perform the


Appendix J: Numerical Methods 201<br />

Table J.1: Particles c<strong>on</strong>strained in energy minimizati<strong>on</strong>s.<br />

C<strong>on</strong>figurati<strong>on</strong> Number of Particles Fixed Particles Fixed<br />

Defect Free 1 Particle at (x,y) = (0,0)<br />

Isolated Dislocati<strong>on</strong> 2 +/- Disclinati<strong>on</strong>s<br />

Two Opposing Dislocati<strong>on</strong>s 4 +/- Disclinati<strong>on</strong>s<br />

Interstitials/Vacancies (I/V) 2 Particle at (x,y) = (0,0) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

Particle at Center of I/V<br />

minimizati<strong>on</strong>, we use the st<str<strong>on</strong>g>and</str<strong>on</strong>g>ard Fletcher-Reeves (FR) c<strong>on</strong>jugate-gradient method<br />

[96, 97] in which particles are moved al<strong>on</strong>g successive directi<strong>on</strong>s determined by the<br />

gradient of the energy in Eq.(J.1) with respect to particle coordinates. After taking<br />

into account the fact that the z-coordinate is determined by h(x, y), the gradient with<br />

respect to the x-coordinate of particle η reads<br />

∂F<br />

∂xη<br />

= �<br />

µ,ν<br />

kµ,ν<br />

2<br />

(rµν − a)<br />

rµν<br />

(δµ,η−δν,η)<br />

�<br />

(xµ − xν) + [h(xµ, yµ) − h(xν, yν)] ∂<br />

∂xη<br />

�<br />

h(xη, yη) ,<br />

(J.3)<br />

with δµ,ν the Kr<strong>on</strong>ecker delta. To obtain the gradient with respect to the y-coordinate,<br />

interchange x ↔ y. 1 C<strong>on</strong>vergence is achieved when the magnitude of the gradient<br />

of the energy drops below some defined tolerance, which was set to 10 −5 (k = 1).<br />

In this work, c<strong>on</strong>vergence was accepted <strong>on</strong>ly if the |∆F discrete | between the last two<br />

iterati<strong>on</strong>s of the algorithm was less than 10 −8 (k = 1).<br />

Since there is a n<strong>on</strong>-zero frustrati<strong>on</strong> energy for a defect-free lattice <strong>on</strong> a<br />

1 The FR algorithm requires the use of a <strong>on</strong>e-dimensi<strong>on</strong>al minimizati<strong>on</strong> algorithm, for which we used<br />

the Brent algorithm [96] as implemented by the Gnu Scientific Library [98]. The Brent algorithm requires<br />

bounds to be placed <strong>on</strong> the dimensi<strong>on</strong>less parameter λ, which we typically take to be −5 < λ < 10.


Appendix J: Numerical Methods 202<br />

bump, the particles will collectively slide off of the bump into flatl<str<strong>on</strong>g>and</str<strong>on</strong>g> if allowed. To<br />

prevent this, some particles were always fixed during minimizati<strong>on</strong> as indicated in<br />

Table J.1. These c<strong>on</strong>straints were implemented so that a particle was always located<br />

at the center of the bump.


Bibliography<br />

[1] P. M. Chaikin <str<strong>on</strong>g>and</str<strong>on</strong>g> T. C. Lubensky. Principles of C<strong>on</strong>densed Matter Physics.<br />

Cambridge University Press, Cambridge, 1995.<br />

[2] D. R. Nels<strong>on</strong>. Defects <str<strong>on</strong>g>and</str<strong>on</strong>g> Geometry in C<strong>on</strong>densed Matter Physics. Cambridge<br />

University Press, Cambridge, 2002.<br />

[3] A. R. Bausch, M. J. Bowick, A. Cacciuto, A. D. Dinsmore, M. F. Hsu, D. R.<br />

Nels<strong>on</strong>, M. G. Nikolaides, A. Travesset, <str<strong>on</strong>g>and</str<strong>on</strong>g> D. A. Weitz. Science, 299:1716,<br />

2003.<br />

[4] A. Travesset M. Bowick, D. R. Nels<strong>on</strong>. Phys. Rev. B, 62:8738, 2000.<br />

[5] J. Heuser. J. Cell. Biol., 108:401, 1989.<br />

[6] D. M. Kroll T. Kohyama <str<strong>on</strong>g>and</str<strong>on</strong>g> G. Gompper. Europhys. Lett., 68:061905, 2003.<br />

[7] S. Schneider <str<strong>on</strong>g>and</str<strong>on</strong>g> G. Gompper. Europhys. Lett., 70:136, 2005.<br />

[8] B. K. Ganser, S. Li, V. Y. Klishko, J. T. Finch, <str<strong>on</strong>g>and</str<strong>on</strong>g> W. I. Sundquist. Nature,<br />

407:409, 2000.<br />

[9] S. Li, C. P. Hill, Sundquist W. I., <str<strong>on</strong>g>and</str<strong>on</strong>g> J. T. Finch. Nature, 407:409, 2000.<br />

203


Bibliography 204<br />

[10] P. Lipowsky <str<strong>on</strong>g>and</str<strong>on</strong>g> al. Nat. Mat., 4:407, 2005.<br />

[11] A. Fern<str<strong>on</strong>g>and</str<strong>on</strong>g>ez-Nieves. private communicati<strong>on</strong>.<br />

[12] T. C. Lubensky <str<strong>on</strong>g>and</str<strong>on</strong>g> J. Prost. J. Phys. II (France), 2:371, 1992.<br />

[13] D. R. Nels<strong>on</strong>. Nano Lett., 2:11252, 2002.<br />

[14] L. D. L<str<strong>on</strong>g>and</str<strong>on</strong>g>au <str<strong>on</strong>g>and</str<strong>on</strong>g> E. M. Lifshitz. Theory of Elasticity. Butterworth, Oxford,<br />

1999.<br />

[15] S. Sachdev <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>. J. Phys. C., 17:5473, 1984.<br />

[16] D. R. Nels<strong>on</strong>. Phys. Rev. B, 27:2902, 1983.<br />

[17] D. R. Nels<strong>on</strong> V. Ambegaokar, B. I. Halperin <str<strong>on</strong>g>and</str<strong>on</strong>g> E. D. Siggia. Phys. Rev. B,<br />

21:1806, 1980.<br />

[18] V. Vitelli A. Hexemer, E. Kremer <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>. American Physical Society<br />

March Meeting, poster sessi<strong>on</strong>, Baltimore 2006.<br />

[19] D. R. Nels<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> B. I. Halperin. Phys. Rev. B, 19:2457, 1979.<br />

[20] J. M. Kosterlitz <str<strong>on</strong>g>and</str<strong>on</strong>g> D. J. Thouless. J. Phys. C., 6:1181, 1973.<br />

[21] A. P. Young. Phys. Rev. B, 19:1855, 1979.<br />

[22] C. F. Chou, A. J. Jin, S. W. Hui, C. C. Huang, <str<strong>on</strong>g>and</str<strong>on</strong>g> J. T. Ho. Science, 280:1424,<br />

1998.<br />

[23] C. Knobler <str<strong>on</strong>g>and</str<strong>on</strong>g> R. Desai. Ann. Rev. Phys. Chem., 43:207, 1992.


Bibliography 205<br />

[24] R. Seshadri <str<strong>on</strong>g>and</str<strong>on</strong>g> R. M. Westervelt. Phys. Rev. Lett., 66:2774, 1991.<br />

[25] C. C. Grimes <str<strong>on</strong>g>and</str<strong>on</strong>g> G. Adams. Phys. Rev. Lett., 42:795, 1979.<br />

[26] E. Y. Andrei D. C. Glattli <str<strong>on</strong>g>and</str<strong>on</strong>g> F. I. B. Williams. Phys. Rev. Lett., 60:420, 1998.<br />

[27] G. Deville, A. Valdes, E. Y. Andrei, <str<strong>on</strong>g>and</str<strong>on</strong>g> F. I. B. Williams. Phys. Rev. Lett.,<br />

53:588, 1984.<br />

[28] C. M. Murray. in B<strong>on</strong>d Orientati<strong>on</strong>al Order in C<strong>on</strong>densed Matter Systems, edited<br />

by K. J. Str<str<strong>on</strong>g>and</str<strong>on</strong>g>burg, (Springer, Berlin, 1992).<br />

[29] R. Lenke K. Zahn <str<strong>on</strong>g>and</str<strong>on</strong>g> G. Maret. Phys. Rev. Lett., 82:2721, 1999.<br />

[30] R. A. Segalman, A. Hexemer, R. C. Hayward, <str<strong>on</strong>g>and</str<strong>on</strong>g> E. J. Kramer. Macromolecules,<br />

36:3272, 2003.<br />

[31] D. R. Nels<strong>on</strong>. Phys. Rev. B, 28:5515, 1983.<br />

[32] F. David, E. Guitter, <str<strong>on</strong>g>and</str<strong>on</strong>g> L. Peliti. J. Phys., 48:2059, 1987.<br />

[33] D. R. Nels<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> L. Peliti. J. Phys., 48:1085, 1987.<br />

[34] E. Guitter <str<strong>on</strong>g>and</str<strong>on</strong>g> M. Kardar. Europhys. Lett., 13:441, 1990.<br />

[35] J. M. Park <str<strong>on</strong>g>and</str<strong>on</strong>g> T. C. Lubensky. Phys. Rev. E, 53:2648, 1996.<br />

[36] P. Lenz <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>. Phys. Rev. E, 67:031502, 2003.<br />

[37] E. Kramer. private communicati<strong>on</strong>.<br />

[38] P. G. de Gennes <str<strong>on</strong>g>and</str<strong>on</strong>g> J. Prost. The Physics of <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> <str<strong>on</strong>g>Crystals</str<strong>on</strong>g>. Clarend<strong>on</strong> Press,<br />

Oxford, 1993.


Bibliography 206<br />

[39] M. Bowick, D. R. Nels<strong>on</strong>, <str<strong>on</strong>g>and</str<strong>on</strong>g> A. Travesset. Phys. Rev. E (in press).<br />

[40] F. David. in Statistical Mechanics of Membranes <str<strong>on</strong>g>and</str<strong>on</strong>g> <strong>Surfaces</strong>, edited by D. R.<br />

Nels<strong>on</strong> et al., (World Scientific, Singapore, 2004).<br />

[41] D. J. Struik. Lectures <strong>on</strong> Classical Differential Geometry. Dover, New York,<br />

1961.<br />

[42] R. Kamien. Rev. Mod. Phys., 74:953, 2002.<br />

[43] V. Vitelli <str<strong>on</strong>g>and</str<strong>on</strong>g> A. M. Turner. Phys. Rev. Lett., 93:215301, 2005.<br />

[44] B. A. Dubrovin, A. T. Fomenko, <str<strong>on</strong>g>and</str<strong>on</strong>g> S. P. Novikov. Modern Geometry- Methods<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> Applicati<strong>on</strong>s, volume 1. Springer-Verlag, New York, 1992.<br />

[45] V. I. Smirnov. A Course of Higher Mathematics, volume 3 part 2. Pergam<strong>on</strong><br />

Press, Oxford, 1964.<br />

[46] A. M. Polyakov. 103:307, 1981.<br />

[47] M. Kleman <str<strong>on</strong>g>and</str<strong>on</strong>g> O. D. Lavrentovich. Soft Matter Physics. Springer-Verlag, New<br />

York, 2003.<br />

[48] W. K. H. Panofsky <str<strong>on</strong>g>and</str<strong>on</strong>g> M. Phillips. Classical Electricity <str<strong>on</strong>g>and</str<strong>on</strong>g> Magnetism.<br />

Addis<strong>on</strong>-Wesley, Reading, 1962.<br />

[49] W. F. Vinen. in Superc<strong>on</strong>ductivity, vol. 2, edited by R. D. Parks, (Dekker, New<br />

York, 1969).<br />

[50] G. B. Hess. Phys. Rev., 161:189, 1967.


Bibliography 207<br />

[51] M. Kleman. Points, Lines <str<strong>on</strong>g>and</str<strong>on</strong>g> walls. Wiley, New York, 1983.<br />

[52] R. D. Kamien. Science, 299:1671, 2003.<br />

[53] D. R. Nels<strong>on</strong>. Nano Lett., 2:1125, 2002.<br />

[54] Park <str<strong>on</strong>g>and</str<strong>on</strong>g> al. Science, 276:1401, 1997.<br />

[55] O. D. Lavrentovich. Liq. Cryst., 24:117, 1998.<br />

[56] G. P. Crawford <str<strong>on</strong>g>and</str<strong>on</strong>g> S. Zumer. <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> <str<strong>on</strong>g>Crystals</str<strong>on</strong>g> in Complex Geometries. Francis<br />

<str<strong>on</strong>g>and</str<strong>on</strong>g> Taylor, L<strong>on</strong>d<strong>on</strong>, 1996.<br />

[57] F. C. MacKintosh <str<strong>on</strong>g>and</str<strong>on</strong>g> T. C. Lubensky. Phys. Rev. Lett., 67:1169, 1991.<br />

[58] V. Vitelli <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>. Phys. Rev. E, 70:051105, 2004.<br />

[59] V. Vitelli <str<strong>on</strong>g>and</str<strong>on</strong>g> A. M. Turner. Phys. Rev. Lett., 93:215301, 2005.<br />

[60] S. Weinberg. Gravitati<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> Cosmology. Wiley, New York, 1972.<br />

[61] S. Ch<str<strong>on</strong>g>and</str<strong>on</strong>g>rasekhar. <str<strong>on</strong>g>Liquid</str<strong>on</strong>g> <str<strong>on</strong>g>Crystals</str<strong>on</strong>g>. Cambridge University Press, Cambridge,<br />

1994.<br />

[62] F. Dys<strong>on</strong>. private communicati<strong>on</strong>.<br />

[63] T. Needham. Visual Complex Analysis. Oxford University Press, Oxford, 1997.<br />

[64] E. B. Wils<strong>on</strong>, J. C. Decius, <str<strong>on</strong>g>and</str<strong>on</strong>g> P. C. Cross. Molecular Vibrati<strong>on</strong>s. Dover, New<br />

York, 1955.<br />

[65] E. M Lifshitz <str<strong>on</strong>g>and</str<strong>on</strong>g> L. D. L<str<strong>on</strong>g>and</str<strong>on</strong>g>au. Quantum Mechanics: N<strong>on</strong>-Relativistic Theory.<br />

Butterworth-Heinemann, Oxford, 1977.


Bibliography 208<br />

[66] C. Chiccoli, I. Feruli, O. D. Lavrentovich, P. Pasini, S. V. Shyianovskii, <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

C. Zann<strong>on</strong>i. Phys. Rev. E, 66:030701, 2002.<br />

[67] H. Stark. Phys. Rep., 351:387, 2001.<br />

[68] T. Lubensky, D. Pettey, N. Currier, <str<strong>on</strong>g>and</str<strong>on</strong>g> H. Stark. Phys. Rev. E, 57:610, 1998.<br />

[69] P. Poulin, H. Stark, T. C. Lubensky, <str<strong>on</strong>g>and</str<strong>on</strong>g> D. A. Weitz. Science, 275:1770, 1997.<br />

[70] D. Caspar <str<strong>on</strong>g>and</str<strong>on</strong>g> A. Klug. Principles in the C<strong>on</strong>structi<strong>on</strong> of Regular Viruses, Cold<br />

Spring Harbor Symposium <strong>on</strong> Quantitative Biology, 27, 1, (1962).<br />

[71] D. R. Nels<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> R. A. Pelcovits. Phys. Rev. B, 16:2191, 1977.<br />

[72] D. R. Nels<strong>on</strong> <str<strong>on</strong>g>and</str<strong>on</strong>g> J. M. Kosterlitz. Phys. Rev. Lett., 39:1201, 1977.<br />

[73] M. Bowick, D. R. Nels<strong>on</strong>, <str<strong>on</strong>g>and</str<strong>on</strong>g> A. Travesset. Phys. Rev. B, 62, 8378, (2000) <str<strong>on</strong>g>and</str<strong>on</strong>g><br />

Phys. Rev. E, 69, 041102, (2004).<br />

[74] V. Vitelli <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>. c<strong>on</strong>d-mat/0406328 .<br />

[75] B. I. Halperin. private communicati<strong>on</strong>.<br />

[76] D. R. Nels<strong>on</strong>. private communicati<strong>on</strong>.<br />

[77] E. J. Yarmchuk <str<strong>on</strong>g>and</str<strong>on</strong>g> R. E. Packard. J. Low Temp. Phys., 46:479, 1982.<br />

[78] E. J. Kramer. Nature, 437:824, 2005.<br />

[79] N. Toan, Bruinsma R. F., <str<strong>on</strong>g>and</str<strong>on</strong>g> W. M. Gelbart.<br />

http://arxiv.org/abs/physics/0506127 .


Bibliography 209<br />

[80] S. Seung <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>. 38:1005, 1988.<br />

[81] R. Bruinsma, B. I. Halperin, <str<strong>on</strong>g>and</str<strong>on</strong>g> A. Zippelius. Phys. Rev. B, 25:579, 1982.<br />

[82] S. Jain <str<strong>on</strong>g>and</str<strong>on</strong>g> D. R. Nels<strong>on</strong>. Phys. Rev. E, 61:1599, 2000.<br />

[83] M. Spivak. A Comprehensive Introducti<strong>on</strong> to Differential Geometry. Publish or<br />

Perish,Inc., Berkeley, 1979.<br />

[84] A. Budzin <str<strong>on</strong>g>and</str<strong>on</strong>g> D. Feinberg. Phys. C, 235:2755, 1994.<br />

[85] A. J. McC<strong>on</strong>nell. Applicati<strong>on</strong>s of Tensor Analysis. Dover Publicati<strong>on</strong>s, New<br />

York, 1957. pp. 184-189.<br />

[86] H. W. Wyld. Mahtematical Methods for Physics. Perseus Books, Reading, 1999.<br />

pp. 272-276.<br />

[87] I. S. Gradstein <str<strong>on</strong>g>and</str<strong>on</strong>g> I. M. Ryshik. Tables of series, products <str<strong>on</strong>g>and</str<strong>on</strong>g> integrals. Verlag<br />

Harri Deutsch, Frankfurt, 1981.<br />

[88] R. P. Feynman, R. B. Leight<strong>on</strong>, <str<strong>on</strong>g>and</str<strong>on</strong>g> M. S<str<strong>on</strong>g>and</str<strong>on</strong>g>s. The Feynman Lectures <strong>on</strong> Physics,<br />

volume 2. Addis<strong>on</strong>-Wesley, Reading, 1963.<br />

[89] B. A. Ovrut <str<strong>on</strong>g>and</str<strong>on</strong>g> S. Thomas. Phys. Rev. D, 43:1314, 1991.<br />

[90] I. C. Bivens. Math. Mag., 65:226, 1992.<br />

[91] T. Needham. Math. Mag., 67:92, 1994.<br />

[92] J. H. Heinbockel. Introducti<strong>on</strong> to Tensor Calculus <str<strong>on</strong>g>and</str<strong>on</strong>g> C<strong>on</strong>tinuum Mechanics.<br />

Trafford, 2001.


Bibliography 210<br />

[93] J. L. van Hemmen <str<strong>on</strong>g>and</str<strong>on</strong>g> M. A. Peters<strong>on</strong>. c<strong>on</strong>d-mat/0503158 .<br />

[94] J. V<str<strong>on</strong>g>and</str<strong>on</strong>g>erlinde. Classical Electromagnetic Theory. Butterworth, Oxford, 1993.<br />

[95] J. S.Koehler. Phys. Rev., 60:397, 1941.<br />

[96] W. H. Press, S. A. Teukolsky, W. T. Vetterling, <str<strong>on</strong>g>and</str<strong>on</strong>g> B. P. Flannery. Numerical<br />

Recipes in C++ (Sec<strong>on</strong>d Editi<strong>on</strong>). Cambridge University Press, Cambridge,<br />

2002.<br />

[97] E. Polak. Computati<strong>on</strong>al Methods in Optimizati<strong>on</strong>. Academic Press, New York,<br />

1971.<br />

[98] M. Galassi, J. Davies, J. Theiler, B. Gough, G. Jungman, M. Buth, <str<strong>on</strong>g>and</str<strong>on</strong>g> F. Rossi.<br />

Gnu Scientific Library Reference Manual. Network Theory Ltd., 2003.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!