Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences
You have accessResearch articles

Equivalent local force conditions minimizing the Frank free energy for topological defect equilibria in nematic liquid crystals

Hiroyuki Miyoshi

Hiroyuki Miyoshi

Department of Information Physics and Computing, Graduate School of Information Science and Technology, The University of Tokyo, Hongo, 7-3-1, Bunkyo-Ku, Tokyo 113-8656, Japan

[email protected]

Contribution: Conceptualization, Formal analysis, Funding acquisition, Methodology, Writing – original draft

Google Scholar

Find this author on PubMed

,
Darren G. Crowdy

Darren G. Crowdy

Department of Mathematics, Imperial College London, 180 Queen’s Gate, London SW7 2AZ, UK

Contribution: Conceptualization, Formal analysis, Project administration, Supervision, Writing – review and editing

Google Scholar

Find this author on PubMed

,
Hiroki Miyazako

Hiroki Miyazako

Department of Information Physics and Computing, Graduate School of Information Science and Technology, The University of Tokyo, Hongo, 7-3-1, Bunkyo-Ku, Tokyo 113-8656, Japan

Contribution: Conceptualization, Funding acquisition, Project administration, Resources, Writing – review and editing

Google Scholar

Find this author on PubMed

and
Takaaki Nara

Takaaki Nara

Department of Information Physics and Computing, Graduate School of Information Science and Technology, The University of Tokyo, Hongo, 7-3-1, Bunkyo-Ku, Tokyo 113-8656, Japan

Contribution: Conceptualization, Funding acquisition, Methodology, Project administration, Writing – review and editing

Google Scholar

Find this author on PubMed

Published:https://doi.org/10.1098/rspa.2024.0569

    Abstract

    Recently, Miyoshi et al. (Miyoshi et al. 2024 Proc. R. Soc. A 480, 20240405 (doi:10.1098/rspa.2024.0405)) derived analytical formulas for the Frank free energy associated with multiple topological defects in nematic liquid crystals confined to an arbitrary simply connected domain. The energy formulas, derived using an analogy with the so-called Kirchhoff–Routh path function in point vortex dynamics, required the evaluation of contour integrals involving analytical formulas associated with the crystal alignment field. Equilibria for topological defects were then obtained by finding local extrema of this Frank free energy. By contrast, in vortex dynamics, it is rare to find point vortex equilibria by minimizing the Kirchhoff–Routh path function that is a regularized kinetic energy; more commonly, a local ‘non-self-induction rule’ is used that is equivalent to each point vortex being free of net force. It is shown here that an analogous equivalent set of local force conditions can also be used to find equilibria for topological defects in nematic liquid crystals in confined domains, and without any need to evaluate the Frank free energy. This observation is significant theoretically and also because, at a technical level, the new local force conditions are easier to formulate and solve.

    1. Introduction

    There has been much recent interest in finding the equilibrium configurations of defects in liquid crystals. Identifying and controlling equilibria is important in bioengineering because topological defects regulate biological processes [13]. Kawaguchi et al. [3] reported that topological defects drive the aggregation of cultured murine neural progenitor cells. They found that the cells concentrate around defects with a charge of +1/2 but escape from defects with a charge of 1/2. Ienaga et al. [4] studied the existence and transition of topological defect formulations in so-called doublet or triplet regions to control the positions of defects. They showed that the topological defect configuration has significant dependence on the boundary shape (e.g. convexity, connectedness). It is therefore important to find effective mathematical techniques for the investigation of such defect equilibria.

    Based on liquid crystal theory and the one-constant approximation [5], the theory of two-dimensional harmonic functions has been exploited for localizing and finding the equilibria of topological defects in confined geometries [69]. Duclos et al. [6] investigated the positions of topological defects in a circular domain of radius R and confirmed both experimentally and mathematically that two defects with a charge of +1/2 are located at a distance of 51/4R from the centre of the disc. Since the alignment angles are harmonic functions in two dimensions, it is also natural to use a complex analysis formulation. Chandler & Spagnolie [10] used Schwarz–Christoffel mappings to find the optimal positions of topological defects on the boundary of polygonal-shaped immersed bodies with finite anchoring boundary conditions. They later extended their theory to study two-body interactions including consideration of body forces and torques [9]. Tang et al. [8] showed that the Frank free energy can be classified into torque-free energy and orientation energy and found that torque determines the dynamics of topological defects. Miyazako & Nara [7] formulated the notion of a complex function for the cell alignment angles based on complex analysis to represent the orientational order of cells. Using conformal mappings, they derived an analytical formula that describes the orientation of cells with topological defects at arbitrary positions strictly inside arbitrary closed domains (i.e. not on the boundaries). In addition, by utilizing the complex form of Green’s theorem, they evaluated the energy associated with the crystals in the domain as line integrals and avoided the divergence of the energy at the singular points associated with the defects. Miyazako et al. also estimated equilibrium distributions of the defects using a Markov chain Monte Carlo method and compared mousemyoblast (C2C12) cells experimental data [11]. Recently, this approach using line integrals has been extended to doubly connected domains [12] using the prime function machinery developed by Crowdy [13] relevant to solving problems involving multiply connected domains.

    Recently, Miyoshi et al. [14] have derived explicit formulas for the Frank free energy associated with nematic liquid crystals with multiple topological defects confined in simply connected two-dimensional domains. This result can be seen as an extension of Čopar & Kos [15], who derived explicit expressions for the Frank free energy of defects in an infinite domain without any objects. Miyoshi et al. evaluated the line integrals in the free energy, up to O(ϵlogϵ), where ϵ is a cut-off parameter that approximates the size of a defect core [14], by pointing out and exploiting a mathematical analogy with the Hamiltonians describing the dynamics of point vortex singularities; these Hamiltonians are regularized kinetic energy functionals that have become known in the field of vortex dynamics as Kirchhoff–Routh path functions [1619]. Miyoshi et al. [14] derived formulas for the Frank free energy making use of conformal mappings from canonical pre-image domains to the physical domain of interest and involving the evaluation of line integrals around circles that enclose the defects and the boundary of the crystal domain. Equilibrium configurations for topological defects were then obtained by finding the extrema of this free energy by setting to zero derivatives of the free energy with respect to the pre-images of the topological defects in the conformal pre-image plane. In the literature on nematic liquid crystals to date, this direct minimization of the Frank free energy has been the tool used to obtain the positions of the equilibria of topological defects [710,20].

    The aim of the present paper is to offer a new perspective. As will be shown here, the aforementioned analogy with point vortex dynamics and the Kirchhoff–Routh path function turns out to have even more ramifications for the theory of two-dimensional nematic liquid crystals. In particular, in the field of vortex dynamics, it is rare to calculate a point vortex equilibrium by calculating and then minimizing the Kirchhoff–Routh path function; in other words, it is rare in vortex dynamics to do what is always done for liquid crystals. Rather, the more usual approach in point vortex dynamics is to write down a complex potential for the fluid flow associated with a given point vortex distribution and, in terms of it, isolate a set of local conditions at each point vortex that are equivalent to that vortex being free of net force: it is therefore natural to refer to these as local force conditions. Saffman [16] gives a careful derivation of this local force condition at each vortex. Some of the vortex dynamics literature refers to this condition as the ‘non-self-induction rule’ since it involves subtracting off a singular term in the complex velocity (the very term causing the singular flow at the point vortex itself) and allowing that point vortex to move with the local velocity induced by anything surrounding it (e.g. other point vortices, or sources of flow not associated with the point vortex itself). Routh [21] is generally agreed to be the first to carefully establish this rule although it is implicit in many other works. Llewellyn Smith [22] has given a careful account of the long history of this local force condition in the field of vortex dynamics. The key contribution of the present paper is to demonstrate that there is an analogous set of equivalent local force conditions that can be used to find the equilibria of topological defects in nematic liquid crystals without having to compute the Frank free energy at all. This important observation is significant both theoretically and practically because, at a technical level, it is easier to write down and solve these equivalent local force conditions than to calculate and minimize the Frank free energy.

    This paper is organized as follows. Section 2 introduces the mathematical formulation of the alignment angle field with topological defects and the Frank free energy. Section 3 reviews the recent work of Miyoshi et al. [14] who derived explicit formulas for the Frank free energy in terms of the defect locations. Sections 4 and 5 contain the main new results of this paper and derive the new local force conditions, in simply and doubly connected regions respectively, that are equivalent to finding extrema of the Frank free energy. These sections also include calculations confirming that previous results obtained using Frank energy minimization are retrieved by the new local force conditions. The paper closes with some general remarks in §6.

    Figure 1 summarizes the contents of this paper, with the new contributions highlighted in blue for clarity. Using the analytical formulas for the energy, the equivalent local force conditions (4.4), (4.6) for simply connected domains and (5.17) and (5.18) for doubly connected domains that minimize the Frank free energy are derived. In §6, the local force on a defect is calculated using the residue of the squares of the derivatives of the complex functions associated with alignment angles.

    Summary of the paper, with the new contributions highlighted in blue.

    Figure 1. Summary of the paper, with the new contributions highlighted in blue. Using the analytical formulas for the energy, we derive the equivalent local force conditions (4.4), (4.6) for simply connected domains and (5.17) and (5.18) for doubly connected domains that minimize the Frank free energy. In §6, the local force on a defect is calculated by evaluating the residue of the squares of the derivatives of the complex functions associated with alignment angles.

    2. Topological defects and the Frank free energy

    Nematic liquid crystals are assumed to be spindle-shaped and infinitely small relative to a characteristic length scale associated with the containing domain. They can thus be treated as a continuum of rod-like particles with that located at (x,y) assumed to be aligned in the direction of unit vector nd as follows:

    nd=(cosϕ(x,y),sinϕ(x,y)),(2.1)

    where ϕ(x,y) is the alignment angle field for (x,y). Since nd and nd are indistinguishable, the arbitrary angle ϕ is equivalent to ϕ[π/2,π/2) [7]; that is, the angles ϕ1 and ϕ2 are equivalent if there exists an integer n such that ϕ1=ϕ2+nπ.

    Let Ω now denote the two-dimensional liquid crystal domain viewed as a subdomain of a complex z=x+iy-plane and let its boundary be denoted by Ω. There is no need for Ω to be simply connected, indeed it will be allowed to have M internal boundaries viewed as the boundaries of M ‘holes’ in the liquid crystal domain. The alignment angle ϕ satisfies the two-dimensional Laplace’s equation except at the defects based on the one-constant approximation of the Frank energy [6], and has a tangential anchoring boundary condition. Assume also that there exist N topological defects at zk=xk+iyk, k=1,,N, strictly inside Ω, whose topological charges are qk. The charges are half-integer or full integer except 0; that is,

    qk{n/2| nZ, n0}.(2.2)

    The set of defect positions is denoted by Z{zk|k=1,2,,N}. In a domain Ω, for any subdomain DkΩ, k=1,,N containing only the k-th defect, the following holds:

    Dkϕsds=2πqk.(2.3)

    Thus, on encircling the defect z=zk once, the alignment angle changes by 2πqk. The quantity (2.3) can be seen as the rotation number of a topological defect located at z=zk.

    Let ψ define the harmonic conjugate of ϕ. Consider the following surface integral in a domain Ω except subdomain k=1NDk:

    0=Ω\(k=1NDk)2ψ(x,y)dxdy=Ωψndsk=1NDkψnds=Ωϕsds+k=1NDkϕsds=2π(1+M+k=1Nqk),(2.4)

    where we used (2.3) and a tangential boundary condition on ϕ on Ω. This result (2.4) suggests that the sum of charges is equal to 1M, i.e.

    k=1Nqk=1M.(2.5)

    A multiply connected generalization of Riemann’s mapping theorem due to Koebe (see [13]) guarantees the existence of a conformal map z=g(ζ) from a pre-image domain Dζ in a complex ζ=ξ+iη-plane to the multiply connected region Ω in the z-plane. A canonical choice for Dζ is a multiply connected circular domain comprising the unit ζ-disc with M smaller circular discs excised [13]. Let the conformal map z=g(ζ) transplant Dζ to Ω. The pre-images of defect positions zk under g(ζ) are denoted as {ζk|k=1,,N} so that

    zk=g(ζk),k=1,,N.(2.6)

    Using the above formulation for a simply connected Ω for which Dζ is just the unit disc (no holes), Miyazako & Nara [7] derived an explicit formula for the alignment angle field ϕ(x,y). When Ω=Dζ, so that the conformal map is just the identity map, the alignment angle field is given by

    ϕ(ξ,η)=Re[fDζ(ζ)],fDζ(ζ)=k=1NqkG(ζ,ζk)π2,(2.7)

    where

    G(ζ,ζk)i(log(ζζk)+log(1ζkζ)),(2.8)

    where ζk is the complex conjugate of the complex number ζk. For an arbitrarily shaped domain Ω, when z=g(ζ) is no longer trivial, the complex function fΩ(z) in Ω must be modified slightly to

    fΩ(z)=fDζ(ζ)ilogg(ζ)=fDζ(g1(z))ilogg(g1(z)),ddζ.(2.9)

    The sum of topological charges qk should satisfy condition (2.5) with M=0.

    3. Frank free energy formula in simply connected domains

    The positions of static topological defects zk are determined where the energy is a local minimum [6,7]. Miyazako & Nara [7] used the complex form of Green’s theorem to represent the energy as a line integral as follows:

    (2)F=K2Ω(ϵ)|ϕ(x,y)|2dxdy=K4iΩ(ϵ)fΩzfΩ¯dz,z12(xiy),(3.1)

    where Ω(ϵ)Ω\D(ϵ), D(ϵ)k=1NDk(ϵ) and Dk(ϵ) is a small disc around the k-th defect with radius ϵ, defined by

    Dk(ϵ){zk+reiθ| 0rϵ,0θ<2π}, k=1,,N.(3.2)

    The regions Dk(ϵ) are excluded when calculating the energy (3.1) in order that the energy is not unbounded. The integration path in a simply connected domain is shown on the left side of figure 2. Since F is conformally invariant, the energy F can be evaluated using the line integral along the unit circle with the boundaries of the excluded zones D~(ϵ) in the unit disc. Using the line integral formula (3.1), Miyazako et al. [7] numerically confirmed that in the case of the unit disc, the energy is minimized when defects exist at positions z=±51/4. This result was initially obtained by Duclos et al. [6].

    Line integral.

    Figure 2. Line integral along (Ω\D(ϵ)) for the case with N=2 for (i) a simply connected domain and (ii) a doubly connected domain. The branch cuts of logarithmic singularities at ζ=ζk are denoted as lk. The imaginary part of log(ζζk) jumps by 2π on lk. The path of integrals in the z-plane is transformed into the integral in the ζ-plane by a conformal map z=g(ζ). The lines lk+, lk denote the paths of the line integral from C0 to D~k(ϵ) used to avoid the jumps of the branch cuts of log(ζζk). In doubly connected domains, the branch cut between C0 and C1 must be selected such that it does not intersect the cuts between D~k(ϵ) and C0.

    In figure 2, lk+, lk denote the paths of the line integral from Dζ to D~k(ϵ) used to avoid the jumps of the branch cuts of log(ζζk). This is crucial as the function Re[fΩ] differs by 2πqk across the branch cut and thus the energy remains undefined unless we select a path that does not intersect the branch cut. The explanation for the integral path for doubly connected domain is given in appendix A and for simply connected domain in [14].

    Miyoshi et al. [14] later derived an explicit formula for the Frank free energy for the unit disc and arbitrarily shaped domains up to O(ϵlogϵ). By evaluating the contour integral as shown on the left side of figure 2, they showed that the energy for the unit disc that excludes small regions D(ϵ)k=1NDk(ϵ) is given by

    F=K2(2πk=1Nqk2logϵ+2πFd)+O(ϵlogϵ),(3.3)

    where the first term in the brackets depends only on qk and ϵ (and not the defect locations) while the second term is given by

    Fd=k=1Nqk2Im[R(ζk,ζk)]+k=1NlkNqkqlIm[G(ζk,ζl)],(3.4)

    where G(.,.) is given in equation (2.8) and

    R(ζ,ζk)G(ζ,ζk)+ilog(ζζk).(3.5)

    The function R(ζ,ζk) is analytic at ζ=ζk and is single-valued. It is obtained by removing the contribution of the logarithmic singularity at ζ=ζk.

    Miyoshi et al. [14] also calculated the energy for an arbitrarily shaped simply connected domain Ω mapped from the unit disc Dζ by a conformal mapping z=g(ζ). They assumed that the boundary of the domain is sufficiently smooth and g(ζ)0 for ζDζ. The total energy F for the region Ω\D(ϵ) is written as follows [14]:

    F=K2(2πk=1Nqk2logϵ+2πFd+Fg+O(ϵlogϵ)),(3.6)

    where

    Fdk=1N(qk22qk)log|g(ζk)|+k=1Nqk2Im[R(ζk,ζk)]+k=1NlkNqkqlIm[G(ζk,ζl)],(3.7)
    FgDζ|g(ζ)g(ζ)|2dxdy+4πlog|g(0)|.(3.8)

    Now the energy Fd depends both on ζk and on the conformal map g(ζ), whereas Fg is independent of ζk. To obtain the extrema of F, only the derivatives of Fd with respect to ζk need to be considered; that is

    Fdζk=0,k=1,,N.(3.9)

    The equilibria for topological defects are obtained by finding {ζk|k=1,...,N} satisfying (3.9). A similar minimization technique was applied to find the locations of topological defects by Chandler & Spagnolie [9] and Duclos et al. [6].

    An important observation is that Im[G(ζk,ζl)] and Im[R(ζk,ζl)] satisfy reciprocity conditions; that is

    Im[G(ζk,ζl)]=(log|ζkζl|+log|1ζlζk|)=Im[G(ζl,ζk)],(3.10)
    Im[R(ζk,ζl)]=log|1ζlζk|=Im[R(ζl,ζk)].(3.11)

    Using (3.10) and (3.11), the following relations can be established:

    Rζ(ζ,ζk)|ζ=ζk=iζkIm[R(ζk,ζk)],Gζ(ζ,ζl)|ζ=ζk=2iζkIm[G(ζk,ζl)].(3.12)

    These properties will be important in the next section.

    4. Local force conditions for nematic liquid crystals

    This section contains the main new results of this paper and builds further on connections to point vortex dynamics and the Kirchhoff–Routh path function pointed out by Miyoshi et al. [14]. To give context to the following results it is worth surveying some results from the theory of point vortex motion.

    The reciprocity relations (3.10) and (3.11) written down at the end of the previous section are reminiscent of similar reciprocity relations satisfied by the so-called hydrodynamic Green’s function [13,1719,23] relevant when studying the dynamics of point vortices in two-dimensional fluid domains. For a simply connected domain, the hydrodynamic Green’s function coincides with the standard first-type Green’s function in potential theory: the function harmonic everywhere in the domain except for a single logarithmic singularity and vanishing everywhere on the domain boundary. In a multiply connected domain, the hydrodynamic Green’s function also has a single logarithmic singularity, vanishes on the outer boundary and has constant non-zero values on the internal boundaries of the holes; these values are chosen in order to ensure zero circulation conditions around the holes. The problem of point vortex dynamics in free space (no boundaries) was shown to be Hamiltonian by Kirchhoff [16] with the relevant Hamiltonian given by a regularized kinetic energy of the fluid. This regularized kinetic energy is akin to the regularized Frank free energy discussed in the previous section where small cut-off regions around the defects are incorporated in order to render the free energy finite. Lin [17,18] later showed that the motion of a finite set of point vortices in an arbitrary multiply connected domain is also Hamiltonian; he proved this using the existence of the hydrodynamic Green’s function in a multiply connected domain and wrote down the Hamiltonian in terms of this Green’s function and its finite part (or Robin function), as well as the conformal mapping from a canonical region. Lin’s work was of an abstract nature and features no example calculations of point vortex trajectories. Crowdy & Marshall [19] later made Lin’s abstract theory constructive by finding explicit expressions for the hydrodynamic Green’s function in the class of multiply connected circular domains Dζ mentioned earlier. The main tool in the theory of Crowdy & Marshall [19] is the prime function associated with Dζ; this is a very general mathematical object that can be associated with any multiply connected domain and has a wide range of applications across the physical sciences [13]. Using this mathematical framework it is easy to exhibit a wide range of point vortex trajectories [19,24,25].

    It is important to point out that while Kirchhoff–Routh theory can be used to write down the Hamiltonians (or Kirchhoff–Routh path functions) associated with point vortex motion in bounded multiply connected domains, it is not necessary to use it and, as mentioned earlier, there are local force conditions at each point vortex that will produce the same dynamics. Restricting interest now to equilibria, which means in the point vortex problem that the arrangement of point vortices sits in steady equilibrium without moving, one can choose either to minimize the Kirchhoff–Routh path function (i.e. find where its partial derivatives with respect to all the point vortex locations vanishes) to find the equilibrium configuration or one can choose to write down and solve the local force conditions. Both will lead to the same point vortex equilibria but are mathematically very different processes.

    The key point is that there is an analogy between point vortex dynamics and the theory of topological defects in nematic liquid crystals and, so far, all studies to find defect equilibria rely on minimizing the Frank free energy (corresponding to solving point vortex problems using Kirchhoff–Routh path functions). But there are also analogous local force conditions in the liquid crystal problem that can be used instead, and these appear to have gone unnoticed until now.

    To make all this more precise, and to actually construct the relevant local force conditions in the nematic liquid crystal problem, it is convenient to define a function f^k(ζ) that is a complex function fDζ(ζ) except for a logarithmic singularity at ζ=ζk; that is

    f^k(ζ)fDζ(ζ)+iqklog(ζζk).(4.1)

    In the presence of multiple vortices, the equilibria of point vortices can be obtained under the condition that the vortex speed at the point vortex is equal to zero. For general geometries, it is necessary to consider an additional term that comes from conformal mappings. The additional term under conformal mappings is obtained from a local analysis around ζ=ζk. We prove that the same technique can be used for finding the energy extrema in the presence of topological defects in the unit disc and arbitrarily shaped domains.

    (a) Local force conditions in the unit disc

    In this case, the complex function and the energy are given by (2.7) and (3.3), respectively. Taking the derivative of F in (3.3) with respect to ζk gives

    Fζk=K22πqk(qkζkIm[R(ζk,ζk)]+2lkNqlζkIm[G(ζk,ζl)]),(4.2)

    where the reciprocity condition Im[G(ζk,ζl)]=Im[G(ζl,ζk)] is used. The derivative of (4.1) with respect to ζ evaluated at ζ=ζk gives

    f^kζ|ζ=ζk=qkRζ(ζ,ζk)|ζ=ζk+lkNqlGζ(ζ,ζl)|ζ=ζk=i(qkζkIm[R(ζk,ζk)]+2lkNqlζkIm[G(ζk,ζl)]),(4.3)

    where the reciprocity conditions (3.10) and (3.11) are used. From (4.3) and (4.2),

    Fζk=iπqkKf^kζ(ζ)|ζ=ζk.(4.4)

    This relation is important: it means that the extrema of the free energy, i.e. solving for F/ζk=0, can be obtained by finding zeros of the derivatives of the set of regularized complex functions f^k(ζ) at each of the defect locations. This is entirely analogous to the situation in the field of vortex dynamics where this equivalence is better known.

    (b) Local force conditions for arbitrary simply connected domains

    Using a conformal map z=g(ζ) from ζ to z, the complex function for the alignment angle is given by (2.9). The derivative of the energy formula (3.6) with respect to ζk gives

    Fζk=K22π(12(qk22qk)g(ζk)g(ζk)+qk2ζkIm[R(ζk,ζk)]+2qklkNqlζkIm[G(ζk,ζl)])=πqkK((qk21)g(ζk)g(ζk)+qkζkIm[R(ζk,ζk)]+2lkNqlζkIm[G(ζk,ζl)]).(4.5)

    Through direct calculation, it can be demonstrated that

    Fzk=iπqkKf^Ω,kz(z)|z=zk=iπqkKg(ζk)(i(qk21)g(ζk)g(ζk)+f^kζ(ζ)|ζ=ζk),(4.6)

    where f^k(ζ) is defined in (4.1) and

    f^Ω,k(z)fΩ(z)+iqklog(zzk).(4.7)

    The derivation of (4.6) is explained in appendix A.

    Therefore, to find the equilibria of topological defects, it is sufficient to find the set of {ζk} that satisfies the set of local force conditions

    V^k(ζk)=1g(ζk)(i(qk21)g(ζk)g(ζk)+f^kζ|ζ=ζk)=0,k=1,,N.(4.8)

    It is not necessary to deal with the Frank free energy directly. We use this fact to find the energy extrema of some examples in the next subsections and to corroborate the results by cross-checking with previous work that has used the Frank free energy.

    (c) Example 1: equilibria in a unit disc

    First let Ω be the unit disc in the complex ζ-plane. Under the assumption that the configuration of static topological defects is symmetrical, there exist two topological defects with a charge of +1/2 located at ζ1=+r and ζ2=r, 0<r<1, r, respectively. The result (4.3) implies that a local condition for equilibrium is

    f^1ζ|ζ=ζ1=i[r2(1r2)14rr2(1+r2)]=i5r414(1r4)=0,(4.9)

    with a similar condition holding, by symmetry, at ζ2. This clearly gives r=51/4, a result verified experimentally by Duclos et al. [6]. It is important to emphasize that this result has been retrieved here without any consideration of the Frank free energy.

    (d) Example 2: equilibria in a quadrature domain

    A doublet or triplet [4,14] is the name given to a domain made up of multiple intersecting discs. In Chapter 11 of [13] Crowdy shows how a class of domains known as quadrature domains can be viewed geometrically as collections of circular discs that have merged so their contact points have been smoothed out into thin necks. A doublet domain Ω can therefore be approximated by the following conformal map from a unit ζ disc to a simply connected quadrature domain:

    g(ζ)=Rζζ2a2,a>1,g(ζ)=Rζ2a22Rζ2(ζ2a2)2=R(ζ2+a2)(ζ2a2)2,(4.10)

    where

    R=a411a+a2(4.11)

    is chosen so that Ω is approximately a domain made up of two merged discs with radius 1. Due to the symmetry of Ω, assume that there exist topological defects with a charge of +1/2 located at ζ1=r and ζ2=r, 0<r<1 in the ζ-plane. From (4.6) it follows that the equilibrium condition at ζ=ζ1 is

    i(q121)g(ζ1)g(ζ1)+f^1ζ|ζ=ζ1=i2[ζ11ζ1ζ1+1ζ1ζ2ζ21ζ2ζ1]3i4g(ζ1)g(ζ1)=i215r42r(1r4)3i4g(r)g(r)=0,(4.12)

    with a similar condition derived at ζ2. This gives the following condition to determine r:

    15r42r(1r4)3r(r2+3a2)r4a4=0,(4.13)

    which coincides with a polynomial formula obtained by considering the derivative of the free energy as considered recently in [14].

    5. Local force conditions for annular regions

    The local force conditions are independent of domain connectivity so it is possible to extend the approach to domains with higher connectivity. To demonstrate this requires showing that imaginary parts of the functions G(.,.) and R(.,.) satisfy the same reciprocity conditions as for the simply connected case; see (5.7) and (5.8) to follow.

    Following Miyazako & Sakajo [7], the following analytical formula for the complex function associated with the alignment angle field with topological defects N located at ζ=ζk, k=1,,N is given, as a function of a variable ζ in a concentric annular region Dζ={ζ|ρ<|ζ|<1}, by

    fDζ(ζ)=k=1NqkG(ζ,ζk)+(i+Qnlogρ)logζπ2,(5.1)

    where

    G(ζ,ζk)i(logP(ζ/ζk,ρ)+logP(ζζk,ρ)+log(ζk)),(5.2)

    where P(.,.) is closely related to the prime function for Dζ [13]. One explicit infinite product representation of it (there are others—see [13]) is

    P(ζ,ρ)=(1ζ)P^(ζ,ρ),P^(ζ,ρ)n=1(1ρ2nζ)(1ρ2nζ1).(5.3)

    Due to the existence of the inner region, the sum of the topological charges qk should satisfy

    k=1Nqk=0.(5.4)

    This is (2.5) with M=1. The value Qn corresponds to the circulation (torque) around the inner boundary, which is calculated using an arbitrary integer n and the arguments of ζk as follows:

    Qn=nπk=1NqkIm[logζk].(5.5)

    It is easy to prove that the function (5.1) satisfies the tangential boundary condition on both inner and outer boundaries.

    Note that the real part of ilogζ in (5.1) has a branch cut between the inner and outer boundaries. However, both the director field and alignment angle in annular regions is continuous as the angles ϕ1 and ϕ2 are equivalent if there exists an integer m such that ϕ1=ϕ2+mπ, as mentioned in §2.

    For simplicity, define a function R(ζ,ζk) that is analytic at ζ=ζk as follows:

    R(ζ,ζk)G(ζ,ζk)+ilog(ζζk)=i(logP^(ζ/ζk,ρ)+logP(ζζk,ρ)).(5.6)

    Similar to the case of a simply connected domain, the imaginary parts of the functions G(.,.) and R(.,.) satisfy the following reciprocity relations:

    Im[G(ζl,ζk)]=(log|P(ζl/ζk,ρ)|+log|P(ζlζk,ρ)|+log|ζk|)=(log|ζl/ζkP(ζk/ζl,ρ)|+log|P(ζkζl,ρ)|+log|ζk|)=Im[G(ζk,ζl)],(5.7)
    Im[R(ζl,ζk)]=(log|P^(ζl/ζk,ρ)|+log|P(ζlζk,ρ)|)=Im[R(ζk,ζl)],(5.8)

    where the following property of the P(.,.) functions [13] are used:

    P(ζ1,ρ)=ζ1P(ζ,ρ),P^(ζ1,ρ)=P^(ζ,ρ).(5.9)

    Due to the reciprocity conditions (5.7) and (5.8), it can be seen that

    Rζ(ζ,ζk)|ζ=ζk=iζkIm[R(ζk,ζk)],Gζ(ζ,ζl)|ζ=ζk=2iζkIm[G(ζk,ζl)].(5.10)

    The Frank free energy in annular regions is calculated in a similar manner to the simply connected case except on a branch cut between the inner and outer boundaries. The path of the contour integral is shown on the right in figure 2. The branch cut between C0 and C1 must be selected such that it does not intersect the cuts between D~k and C0. The integral path is also chosen not to intersect the branch cut, as mentioned in appendix A. Using the same method as that used for simply connected domains in [14] and considering the effect on the branch cut, the Frank free energy for Dζ excluding small discs Dk(ϵ) with radius ϵ around ζ=ζk is given by

    F=K2(2πk=1Nqk2logϵ+2πFd+O(ϵlogϵ)),(5.11)

    where

    Fd=(1+Qn2(logρ)2)logρ2k=1NqkRe[logζk]+k=1Nqk2Im[R(ζk,ζk)]+k=1NlkNqkqlIm[G(ζk,ζl)].(5.12)

    We also consider an arbitrarily shaped doubly connected domain Ω in the z-plane mapped from the annular region Dζ by conformal mapping z=g(ζ), where the boundary of the domain is sufficiently smooth. Using the same technique as that used for simply connected domains (2.9), the complex function for the doubly connected domains fΩ(z) is given by

    fΩ(z)=fDζ(ζ)ilogg(ζ)=fDζ(g1(z))ilogg(g1(z)).(5.13)

    The total energy F of the region Ω\D(ϵ), D(ϵ)k=1NDk(ϵ), can be written as follows:

    F=K2(2πk=1Nqk2logϵ+2πFd+Fg+O(ϵlogϵ)),(5.14)

    where

    Fd(1+Qn2(logρ)2)logρ2k=1NqkRe[logζk]+k=1Nqk2Im[R(ζk,ζk)]+k=1NlkNqkqlIm[G(ζk,ζl)]+k=1N(qk22qk)log|g(ζk)|,(5.15)
    FgDζ|g(ζ)g(ζ)|2dS.(5.16)

    The detailed derivations of (5.12) and (5.14) are given in appendix B.

    Using the expression (5.12), it can be shown that for a concentric annulus,

    Fζk=iπqkKf^kζ(ζ)|ζ=ζk,(5.17)

    where the regularized function f^k(ζ)fDζ(ζ)+iqklog(ζζk) with fDζ(ζ) defined in (5.1). For arbitrarily shaped domains, by direct calculations of the derivative of f^Ω,k(z)fΩ(z)+iqklog(zzk) and of the energy (5.14), it can be shown that

    Fzk=iπqkKf^Ω,kz(z)|z=zk.(5.18)

    The derivations are explained in appendices C and D. The equations (5.17) and (5.18) are the equivalent local force conditions that minimize the energy for doubly connected domains. Some examples are given in the next subsection.

    (a) Example 3: equilibria of pairs of topological defects in annular regions

    An important check is to verify that the local force conditions in a concentric annulus just derived retrieve the same equilibria as computed recently by Miyazako & Sakajo [12] by minimizing the Frank free energy. The first case is where the 1/2 defect is located at ζ=r1 and the +1/2 defect is located at ζ=r2, where ρ<r1<r2<1 in concentric annular regions with inner radius ρ [12]. It is convenient to define the derivative of the logarithm of P(.,.) with respect to ζ:

    K(ζ,ρ)ζζlogP(ζ,ρ)=ζ1ζ+n=1[ρ2nζ1ρ2nζ+ρ2nζ11ρ2nζ1],(5.19)
    K^(ζ,ρ)ζζlogP^(ζ,ρ)=n=1[ρ2nζ1ρ2nζ+ρ2nζ11ρ2nζ1].(5.20)

    To find the equilibria of these defects, the locations ζ=r1 and ζ=r2 are found based on the local force conditions, that is,

    f^1ζ(r1)=i(K(r12,ρ)2r112r1[K(r1/r2,ρ)+K(r1r2,ρ)]1r1)=0,(5.21)
    f^2ζ(r2)=i(K(r22,ρ)2r2+12r2[K(r2/r1,ρ)+K(r2r1,ρ)]1r2)=0,(5.22)

    where the fact that K^(1,ρ)=0 has been used. The locations r1 and r2 can be found using a standard nonlinear solver.

    Figure 3 shows the equilibria of topological defects in annular regions with ρ=0.1, ρ=0.15 and the colour maps of alignment angles Re[fDζ(ζ)]. The locations of the topological defects were calculated from the conditions (5.21) and (5.22) for ρ=0.1 and ρ=0.15. Figure 4 shows the values of r1 and r2. It is numerically confirmed that when ρ is greater than approximately 0.1530, there is no equilibrium for defect pairs at locations r1 and r2. This transition was calculated by Miyazako & Sakajo [12]. The numerical results obtained using the local force conditions for annular regions are consistent with their results (when a parameter n in their notation is chosen to be zero).

    annular region with pair of topological defects.

    Figure 3. Colour maps of Re[fDζ(ζ)] in an annular region with a pair of topological defects. A defect with a charge of 1/2 is located at ζ=r1 and a defect with a charge of +1/2 is located at ζ=r2, where ρ<r1<r2<1. The locations of topological defects can be using the conditions (5.21) and (5.22). When ρ=0.1 and ρ=0.15, the two defects are in equilibrium as the conditions are satisfied at these locations. When ρ0.1530, our numerical calculation shows that no equilibria can be found.

    annular region with a pair of topological defects.

    Figure 4. The values of r1 and r2 that satisfy local force conditions in an annular region with a pair of topological defects. The values r1 and r2 are found from conditions (5.21) and (5.22). It is found that no equilibrium can be obtained for a pair of defects located at ζ=r1 and ζ=r2 when ρ0.1530. The results obtained using the local force conditions are the same as those reported by Miyazako & Sakajo [12].

    It is important to check that when ρ0 and r10, the location of the other defect r2 is determined by the equilibrium of only one defect with a charge of +1/2 in the presence of a defect with a charge of +1/2 fixed at r1=0. The local force condition at ζ=r2 is

    f^2ζ|ζ=r2=i(r22(1r22)12r2)=i(2r2212r2(1r22))=0,(5.23)

    which gives r2=1/2. This is consistent with the numerical results shown in figure 4 and the value predicted by Miyazako & Sakajo [12].

    The second case is where two 1/2 defects are located at ζ=r1, r1, respectively, and two +1/2 defects are located at ζ=r2, r2, respectively, where ρ<r1<r2<1. The third case is where three 1/2 defects are located at ζ=r1, r1ω, r1ω2, respectively, and three +1/2 defects are located at ζ=r2, r2ω,r2ω2, respectively, where ρ<r1<r2<1 and ωexp(2πi/3). Both cases were calculated by Miyazako & Sakajo using the contour integral for the free energy [12]. To find the equilibria of these defects, we consider the equivalent local force conditions. Figure 5 shows the results for two pairs of topological defects. When ρ=0.1 and ρ=0.14, four defects are in equilibrium, as equilibrium conditions are satisfied at these locations. However, when ρ0.144, no solutions for r1 and r2 can be found so no equilibrium exists in this case. This result is numerically verified as shown in the colour map on the right side of figure 5. We have also confirmed that it is impossible to obtain any formations of equilibria for the three pairs of +1/2 and 1/2 topological defects with threefold symmetry as shown in figure 6.

    annular region with two pairs.

    Figure 5. Colour maps of Re[fDζ(ζ)] in an annular region with two pairs of +1/2 and 1/2 topological defects. Two defects with a charge of 1/2 are located at ζ=r1, r1, and two defects with a charge of +1/2 are located at ζ=r2, r2, where ρ<r1<r2<1. The optimal values of r1 and r2 can be found by the local force conditions. When ρ=0.1 and ρ=0.14, the two defects are in equilibrium as equilibrium conditions are satisfied at these locations.

    annular region with inner radius.

    Figure 6. Colour map of Re[fDζ(ζ)] in an annular region with inner radius ρ in the presence of three pairs of topological defects. Accounting for the symmetry, three defects with a charge of 1/2 are located at ζ=r1, r1ω, r1ω2, and three defects with a charge of +1/2 are located at ζ=r2, r2ω, r2ω2, where ρ<r1<r2<1 and ωexp(2πi/3). This case is shown not to admit an equilibrium for any ρ.

    6. Final remarks

    The analytical formulas derived in this paper give local conditions for the equilibria of topological defects that, we have shown, are equivalent to minimizing the Frank free energy. Throughout the paper these have been referred to as local force conditions and it is important to justify use of this terminology mainly because they have been derived here purely mathematically without reference to any physical interpretation.

    One reason for associating the local conditions with forces is due to the analogy with point vortex dynamics where this local condition at each point vortex is indeed known to be equivalent to that point vortex having no net force on it; see the arguments given in Saffman [16] based on Newton’s second law of mechanics.

    However, it is also possible to use the principle of virtual work to argue that the analogous local conditions at defects in nematic liquid crystals are also associated with local forces. We assume that the energy increase δF caused by small displacements of zk is given by

    δF=Fzkdzk+Fzk¯dzk¯.(6.1)

    By the principle of virtual work,

    δF=(fxdxk+fydyk)=(fxify)dzk2(fx+ify)dzk¯2,(6.2)

    where dzk=dxk+idyk and (fx,fy) is the force on the defect located at zk. By equating (6.1) and (6.2), the force on the defect located at z=zk can be associated with the local derivative of the complex function as follows:

    fxify=2Fzk=2iπqkKf^Ω,kz|z=zk=2iπqkKg(ζk)(i(qk21)g(ζk)g(ζk)+f^kζ(ζ)|ζ=ζk),(6.3)

    where, importantly, we used (4.6) for the second equality. Because this is the force on a topological defect in a confined geometry it is natural to refer to our equivalent set of local conditions for minimizing the Frank free energy as local force conditions.

    Chandler & Spagnolie [10] calculated the forces on immersed bodies with tangential or weak anchoring boundary conditions also using the principle of virtual work. Defect dynamics governed by the field have also been studied [26]. The results were compared with those obtained using Toner-Tu models based on statistical mechanics. In view of this, it is worth pointing out another observation: the same local force conditions derived here can be derived in a quite different way by modelling the defect core with radius ϵ as an immersed impenetrable body. Chandler & Spagnolie [9] derived a formula for the forces on an immersed body, a formula akin to the Blasius force formula in fluid mechanics (those authors also discuss several other analogies with fluid dynamics, but they do not touch on the analogy with point vortex dynamics explored in the present paper or in [14]). Following their theory, the force on a defect with defect core modelled as a small immersed circular body Dk(ϵ) of radius ϵ1 is calculated as

    fxify=iK2Dk(ϵ)(fΩz)2dz=iK2Dk(ϵ)[iqkzzk+f^Ω,kz(z)|z=zk+O(zzk)]2dz=iK(2πqk)f^Ω,kz(z)|z=zk,(6.4)

    where the residue theorem has been used in the final step. This expression is the same as (6.3).

    It should be noted that another derivation for (6.3) was given by Denniston [27]. There the alignment angle field ϕ is decomposed into ϕq and ϕex, where ϕex denotes the external field of the function harmonic at the defect core and ϕq satisfies the following boundary condition on the boundary of the defect core with a small radius:

    ϕqqkθ+const(6.5)

    as zzk+ϵeiθ. The main result of Denniston [27] is the force given as the charge multiplied by the gradient of ϕex evaluated at z=zk times a rotational matrix. The final formula (6.3) can be seen as a complex form of the result and a simple extension of it to multiply connected domains with tangential anchoring boundary conditions. The equation for forces on topological defects (6.3) has also been derived by Giomi et al. [28]. They formulated the force on a defect influenced by the disclination of the other defects in the presence of an external fluid. Nevertheless, the formula derived here is new in the sense that we show the force on the defect is also expressed by the rotational number 2πqk multiplied by the derivative of the complex function based on the analogy with point vortex dynamics that has given rise to the new results.

    While the focus here has been on equilibria, the dynamics of defects are expressed in terms of the gradient of the Frank free energy for topological charges and torques for defects. Tang et al. considered both the charge and torque for defects and calculated the free energy-driven dynamics [8]. This approach was extended to calculate the dynamics of defects located above a wall and under some active stress [20]. The orientation of defects has also been considered [29] to capture the force field of two interacting defects. Čoper & Kos also derived a system of equations for torques associated with many defects and calculated the forces in unbounded domains [15]. Future research will focus on using the torque to describe the complex behaviour of defects in confined geometries.

    Finally, although only simply and doubly connected geometries have been treated in detail in this paper, there appears to be no obstruction to generalization of the main result to domains of arbitrary finite connectivity, that is, to nematic liquid crystal domains of connectivity M+1 for any M0. In that case, conformal mapping from a circular domain Dζ with M circular holes is relevant. All the function theoretic machinery based on the prime function appropriate to such circular domains exists [13] and it is likely that a rich constructive framework extending the work herein can be built for nematic liquid crystals much the same as has already been done in point vortex dynamics [19,24,25]. This is the subject of ongoing work by the authors.

    Data accessibility

    Data available from the Zenodo Repository [30].

    Declaration of AI use

    We have not used AI-assisted technologies in creating this article.

    Authors’ contributions

    H.M.: conceptualization, formal analysis, funding acquisition, methodology, writing— original draft; D.C.: conceptualization, formal analysis, project administration, supervision, writing— review and editing; H.M.: conceptualization, funding acquisition, project administration, resources, writing—review and editing; T.N.: conceptualization, funding acquisition, methodology, project administration, writing—review and editing.

    All authors gave final approval for publication and agreed to be held accountable for the work performed therein.

    Conflict of interest declaration

    We declare we have no competing interests.

    Funding

    The first author was supported by a Postdoctoral Research Fellowship from the Japan Society for the Promotion of Science (JSPS-24KJ0041). The second author was partially funded by EPSRC Grant EP/V062298/1. The third author was partially supported by the Japan Society for the Promotion of Science KAKENHI (grant numbers JP23H00086 and JP23H04406).

    Appendix A. Derivation of local force condition for arbitrary simply connected domains

    We associate (4.5) with a conformal mapping z=g(ζ) for arbitrarily shaped domains. The Taylor expansion of g(ζ) around ζ=ζk gives

    1ζζk=g(ζk)zzk+g(ζk)2g(ζk)+O(zzk).(A 1)

    The derivative of the complex potential fΩ(z) defined in (2.9) with respect to ζ is given by

    fΩζ=iqkζζk+qkR(ζ,ζk)ζ+lkNqlG(ζ,ζl)ζig(ζ)g(ζ)=iqk(g(ζk)zzk+g(ζk)2g(ζk))+qkR(ζ,ζk)ζ+lkNqlG(ζ,ζl)ζig(ζ)g(ζ)+O(zzk).(A 2)

    By the chain rule and the expansion of the conformal map g(ζ), we have

    fΩζ=fΩzdzdζ=fΩz(g(ζk)+(zzk)g(ζk)g(ζk)+).(A 3)

    Using (A 2) and (A 3), the derivative of fΩ with respect to z is given by

    fΩz=fΩζ(g(ζk)+(zzk)g(ζk)g(ζk))1+O(zzk)=iqkzzk+V^k+O(zzk),(A 4)

    where V^k is calculated using the following Taylor expansion:

    V^k=1g(ζk)(i(qk21)g(ζk)g(ζk)+qkR(ζ,ζk)ζ|ζ=ζk+lkNqlG(ζ,ζl)ζ|ζ=ζk)=1g(ζk)(i(qk21)g(ζk)g(ζk)+f^kζ|ζ=ζk),(A 5)

    where f^k(ζ) is defined in (4.1). To derive the first line in (A 5), the reciprocity conditions (3.10) and (3.11) are used. Hence, the final expression (4.6) is obtained.

    Appendix B. Proof of energy formula for doubly connected domains

    Here, we prove the energy formula for annular regions (5.12). By direct calculation, it can be shown that the above formula can be written using f^k(ζ) defined in (4.1):

    Fd=(1+Qn2(logρ)2)logρ+k=1NqkIm[f^k(ζk)]k=1NIm[(i+Qnlogρ)qklogζk].(B 1)

    Following [14], the energy can be written as a contour integral along the branch cuts as follows:

    F=K2DζD(ϵ)|ϕ|2dS=K212i(DζD(ϵ))fDζζfDζ¯dζ,(B 2)

    where fDζ(ζ) is defined in (5.1). The path of the contour integral is shown in figure 7. The boundary of the unit disc is denoted as C0 and that of the inner disc is denoted as C1. The integral paths lk±, k=1,,N and lρ± are chosen so that the paths do not cross the branch cuts of the logarithmic singularities at ζ=ζk, k=1,,N. We also choose branch cuts from the inner disc denoted as lρ± so that they do not cross the other branch cuts and Dk(ϵ). We denote the intersection between lk± and C0 as ak± and that between lk± and Dk(ϵ) as bk±. We also denote the intersection between lρ± and C0 as 1± and that between lρ± and C1 as ρ±. The imaginary parts of log(ak±ζk) and log(bk±ζk) are denoted as αk± and βk±, respectively. The study by Miyoshi et al. [14] explains the procedure for simply connected domains in detail.

    Paths of contour integral.

    Figure 7. Paths of contour integral in Dζ\D(ϵ) (left) and g1(Ω\D(ϵ))=Dζ\D~(ϵ) (right). (Left) The disc with inner radius ϵ1 should be eliminated to calculate the free energy defined by (3.1). The path of integrals are chosen such that they do not cross any other branch cuts. The inner disc D2(ϵ) should be modified when a conformal map g(ζ) is considered as shown in the right figure.

    1. Line integral along lρ±:

    • When ζ± is on lρ±, the function fDζ satisfies

    fDζ(ζ+)fDζ(ζ)=2iπ(i+Qnlogρ).(B 3)

    Using relation (B 3), the line integral along lρ± can be evaluated as follows:

    12ilρ+fDζζfDζ¯dζ12ilρfDζζfDζ¯dζ=π(i+Qnlogρ)[fDζ(ρ+)fDζ(1+)].(B 4)
    1. Line integral alonglk±

    • We use the same technique as that described in [14]. On the path ζ±lk±, the function fDζ(ζ) satisfies

    fDζ(ζ+)fDζ(ζ)=2πqk.(B 5)

    Using relation (B.5) and log(bkζk)=logϵ+iβk, the line integral along lk± is given by

    12ilk+fDζζfDζ¯dζ12ilkfDζζfDζ¯dζ=iπqklkfDζζdζ=πqk2logϵiπqkf^k(ζk)iπqk2βk+iπqkfDζ(ak)+O(ϵ),(B 6)

    where f^k(ζ) is defined by (4.1).

    1. Line integral alongDk(ϵ)

    • We use f^k(ζ) and the derivative of the function. Using the parameterization around Dk as ζ=ζk+ϵeiθ, θ[βk,βk+], we have

    12iDk(ϵ)fDζζfDζ¯dζ=12βkβk+[iqkϵeiθ+f^kζ(ζk+ϵeiθ)][iqklogϵeiθ+f^k(ζk+ϵeiθ)]¯ϵeiθdθ=πqk2logϵ+iπqkf^k(ζk)¯+iπqk2βk+iπ2qk2+O(ϵlogϵ).(B 7)
    1. Line integral alongC1

    • For ζ=ρeiθC1, θ[π,+π), the use of (5.9) gives

    Re[fDζ(ζ)]=Re[k=1NqkG(ζ,ζk)]+θ+Qnπ2=θ+nππ2.(B 8)

    The line integral along C1 becomes

    12iC1fDζζfDζ¯dζ=12iC1fDζζ(2θ+π(2n1)fDζ)dζ,(B 9)

    where a minus sign is placed in front of the integral because the integral along C1 is taken in the clockwise direction. Using integration by parts, the first term of (B 9) is

    22iC1fDζζθdζ=2iπfDζ(ρ+)+2π2(i+Qnlogρ)iππfDζ(ρeiθ)dθ,(B 10)

    where we used (B 3). To calculate the final term of (B 10), we note that

    iππfDζ(ρeiθ)dθ=iππ[k=1NqkG(ζ,ζk)+(i+Qnlogρ)log(ρeiθ)]dθ+iπ2=2πk=1Nqklogζk2iπlogρ(i+Qnlogρ)+iπ2,(B 11)

    where we used (5.4) and

    ππlogP(ρeiθα,ρ)dθ=ππlogP(ρeiθ/β,ρ)dθ=0,(B 12)

    for any complex parameters α, β strictly inside Dζ. This can be proven by the standard residue theorem of logarithmic functions; that is

    C0log(1tζ)ζdζ=0,|t|<1.(B 13)

    The second term of (B 9) becomes

    π(2n1)12iC1fDζζdζ=π2(2n1)(i+Qnlogρ),(B 14)

    where we used (B 3). We use (B 3) again to show that for the third term of (B 9),

    +12iC1fDζζfDζdζ=14i[fDζ2]ζ=ρζ=ρ+=π(i+Qnlogρ)[fDζ(ρ+)πi(i+Qnlogρ)].(B 15)

    By combining (B 10), (B 14) and (B 15), the line integral around C1 is given by

    12iC1fDζζfDζdζ=π(i+Qnlogρ)fDζ(ρ+)2πk=1Nqklogζk+π(i+Qnlogρ)[2π(1n)iπQnlogρ2ilogρ]+iπ2.(B 16)
    1. Line integral alongC0

    We note that

    12iC0fDζζfDζdζ=12iC0fDζζ(2ϕ(θ)πfDζ)dζ,(B 17)

    where ϕ(θ)θ+χ(θ), where χ(θ) is a step function that has a jump at θ=αk± by 2πqk. The visualization of χ(θ) is shown in fig. 1 in [14]. The first term of (B 17) becomes

    1iC0fDζζϕdζ=ik=1N[fDζ(eiθ)(θ+χ(θ))]θ=αkθ=αk+i[fDζ(eiθ)(θ+χ(θ))]θ=πθ=π++iππfDζ(eiθ)dθ=2iπk=1Nqk(fDζ(ak)+αk+χ(αk)+2πqk)2iπfDζ(1+)2π2(i+Qnlogρ)+iππfDζ(eiθ)dθ=2iπk=1Nqk(fDζ(ak)+πqk)2iπfDζ(1+)2π2Qnlogρ+iπ2,(B 18)

    where we used

    iππfDζ(eiθ)dθ=iππRe[fDζ(eiθ)]dθ=2iπqkk=1N(αk+χ(αk)+πqk)iπ2,(B 19)

    where from the first equality, we used the functional property of P^(.,.) and P(.,.). The relation comes from

    ππ(log|P(eiθ/ζk,ρ)|+log|ζk|)dθ=ππ(log|eiθζk|+log|P^(eiθ/ζk,ρ)|)dθ=0,(B 20)
    ππlog|P(eiθζk,ρ)|dθ=ππ(log|1eiθζk|+log|P^(eiθζk,ρ)|)dθ=0,(B 21)

    where we used the residue theorem (B 13). Using the residue theorem at ζ=ζk and ζ=0, the second term in (B 17) becomes

    π2iC0fDζζdζ=π2(i+Qnlogρ),(B 22)

    where we used k=1Nqk=0. The third term of (B 17) becomes

    12iC0fDζζfDζdζ=14ik=1N[fDζ2]ζ=akζ=ak+14i[fDζ2]ζ=1ζ=1+=iπk=1Nqk(fDζ(ak)+πqk)π(i+Qnlogρ)(fDζ(1+)iπ(i+Qnlogρ)).(B 23)

    By combining (B 18), (B 22) and (B 23), the line integral along C0 is given by

    12iC0fDζζfDζdζ=iπk=1Nqk(fDζ(ak)+πqk)π(i+Qnlogρ)fDζ(1+)+iπ2Qnlogρ(i+Qnlogρ)+iπ2.(B 24)
    1. Summation

    The summation of (i)–(v) gives

    12i(Dζ\D(ϵ))fDζζfDζdζ=2πk=1Nqk2logϵ+2πk=1NqkIm[f^k(ζk)]2πk=1Nqklogζk2π(1+Qn2(logρ)2)logρ2π(i+Qnlogρ)k=1NIm[qklogζk].(B 25)

    By direct calculation, the formula (B 25) except the first term is the same as (B 1) multiplied by 2π.

    For general domains, we consider the conformal map z=g(ζ) from annular region Dζ to the domain Ω. We define the mapped region D~k(ϵ)g1(Dk(ϵ)). Since the Frank free energy is conformally invariant, the free energy is calculated in the annular region. The integral path is shown on the right side of figure 7. However, two simple modifications are needed; that is, the radius of inner disc D~k(ϵ) becomes ϵ/|g(ζk)|+O(ϵ2) and the complex function becomes (5.13). By using the expression in (5.13), the contour integral for the energy is given by

    F=K212i(Dζ\D~(ϵ))fΩζfΩdζ=K212i(Dζ\D~(ϵ))(fDζζ+hζ)(fDζ+h)dζ=K212i(Dζ\D~(ϵ))(fDζζfDζ+fDζζh+hζfDζ+hζh)dζ,(B 26)

    where h(ζ)ilogg(ζ). The first term of (B 26) divided by K/2 is the same as (B 25), except the radius ϵ now becomes ϵ/|g(ζk)| as the inner disc D~k(ϵ) is excluded. The final term of (B 26) is

    K212i(Dζ\D~(ϵ))hζhdζ=K2[Dζ|gg|2dS+O(ϵ2)].(B 27)

    To prove (5.14), some simple modifications to (5.12) are needed. Due to the single-valueness of the integrand along lk± and lρ±, the second term of (B 26) is

    lk+fDζζhdζlkfDζζhdζ=lρ+fDζζhdζlρfDζζhdζ=0.(B 28)

    Also,

    12iD~k(ϵ)fDζζhdζ=12βkβk+[iqkϵ~eiθ+f^kζ(ζk+ϵeiθ+ζk)]h(ϵ~eiθ+ζk)ϵ~eiθdθ+O(ϵlogϵ)=πiqkh(ζk)+O(ϵlogϵ)=πqklogg(ζk)+O(ϵlogϵ),(B 29)

    and

    12iC0fDζζhdζ12iC1fDζζhdζ=πk=1Nqklogg(ζk),(B 30)

    where we used the residue theorem, that is

    C0fDζζh¯dζC1fDζζh¯dζ=k=1Nakak+1+fDζ¯θhdθππfDζ¯θhdθ¯=2ππh(eiθ)dθ2ππh(ρeiθ)dθC0fDζζhdζ+C1fDζζhdζ¯,

    where we used fDζ/θ+fDζ/θ=2 on C0 and C1 except ζ=ak±, k=1,,N and ζ=1±.

    For the third term of (B 26), we note that

    12ilρ+hζfDζdζ12ilρhζfDζdζ=π(i+Qnlogρ)(h(ρ)h(1)),(B 31)

    where we used (B 3) and

    12ilk+hζfDζdζ12ilkhζfDζdζ=iπqk[h(bk)h(ak)]=iπqk[h(ζk)h(ak)]+O(ϵ),

    and

    12iD~k(ϵ)hζfDζdζ=O(ϵlogϵ).(B 32)

    The line integral along C0 gives

    12iC0hζfDζdζ=12iC0hζ(2ϕπfDζ)dζ=iπk=1Nqkh(ak)π(i+Qnlogρ)h(1)+iππh(eiθ)dθ12iC0(πhζhfDζζ)dζ,

    and the line integral along C1 is

    12iC1hζfDζdζ=π(i+Qnlogρ)h(ρ)iππh(ρeiθ)dθ+12iC1(π(2n1)hζhfDζζ)dζ.

    It is noted that

    iππh(eiθ)dθiππh(ρeiθ)dθ=C0h(ζ)ζdζC1h(ζ)ζdζ=0,(B 33)

    and due to the fact that the inner and outer boundaries are not self-intersecting,

    C0hζdζ=iC0ggdζ=C1hζdζ=iC1ggdζ=0.(B 34)

    Hence, the third term of (B 26) becomes

    12i(Dζ\D~(ϵ))hζfDζdζ=2πk=1Nqklogg(ζk),(B 35)

    where we used the residue theorem to calculate

    12iC0hfDζζdζ12iC1hfDζζdζ=πk=1Nqklogg(ζk).(B 36)

    Combining these, the formula (5.14) is derived.

    Appendix C. Local force conditions for a concentric annulus

    Using the energy formula, this section proves that the formula derived using the local force condition is equivalent to the condition for the local extrema of topological defects in annular regions. The derivative of F given in (5.12) with respect to ζk gives

    Fζk=πqkK[qkζkIm[R(ζk,ζk)]+2lkNqlζkIm[G(ζk,ζl)](1+iQnlogρ)1ζk],(C 1)

    where we used the reciprocal properties of G(ζ,ζk) and R(ζ,ζk) given in (5.7) and (5.8), respectively, and the following properties:

    Qnζk=iqk2ζk,ζk(Re[logζk])=12ζk.(C 2)

    Hence, the derivative of the energy in (C 1) becomes

    Fζk=iπqkK[qkRζk(ζ,ζk)|ζ=ζk+lkNqlGζ(ζ,ζl)|ζ=ζk+(i+Qnlogρ)1ζk].(C 3)

    Similar to §5, we now associate the derivative of the energy (C 1) with the local force conditions. The derivative of f^k(ζ) evaluated at ζ=ζk is

    f^kζ(ζ)|ζ=ζk=qkRζ(ζ,ζk)|ζ=ζk+lkNqlGζ(ζ,ζl)|ζ=ζk+(i+Qnlogρ)1ζk.(C 4)

    Hence, the derivative of the free energy is written as

    Fζk=iπqkKf^kζ(ζ)|ζ=ζk.(C 5)

    Appendix D. Local force conditions for arbitrary doubly connected domains

    Using the same technique as that used for simply connected domains in §3, the derivative of fΩ(z) defined in (5.13) is given using the chain rule for z=g(ζ) as follows:

    fΩz=iqkzzk+V^k+O(ζζk),(D 1)

    where V^k is given by the same expression (A 5) with different fDζ(ζ) defined in (5.1). By direct calculation of the derivative of f^Ω,k(z)fΩ(z)+iqklog(zzk), it can be shown that

    Fzk=iπqkKf^Ω,kz(z)|z=zk.(D 2)

    Therefore, to find the equilibria for topological defects in doubly connected domains whose mapping from a concentric annulus is known, it is sufficient to find {ζk} satisfying the local conditions

    g(ζk)V^k=iπqkK(i(qk21)g(ζk)g(ζk)+f^kζ(ζ)|ζ=ζk)=0,k=1,,N.(D 3)

    Footnotes

    Published by the Royal Society. All rights reserved.