This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A One-Dimensional Variational Problem for Cholesteric Liquid Crystals with Disparate Elastic Constants

Dmitry Golovaty Michael Novack Peter Sternberg Department of Mathematics, The University of Akron, Akron, OH 44325 Department of Mathematics, The University of Texas at Austin 2515 Speedway, RLM 8.100, Austin, TX 78712 Department of Mathematics, Indiana University, Rawles Hall 831 East 3rd St., Bloomington, IN 47405
Abstract

We consider a one-dimensional variational problem arising in connection with a model for cholesteric liquid crystals. The principal feature of our study is the assumption that the twist deformation of the nematic director incurs much higher energy penalty than other modes of deformation. The appropriate ratio of the elastic constants then gives a small parameter ε\varepsilon entering an Allen-Cahn-type energy functional augmented by a twist term. We consider the behavior of the energy as ε\varepsilon tends to zero. We demonstrate existence of the local energy minimizers classified by their overall twist, find the Γ\Gamma-limit of the relaxed energies and show that it consists of the twist and jump terms. Further, we extend our results to include the situation when the cholesteric pitch vanishes along with ε\varepsilon.

keywords:
Cholesteric liquid crystal , Gamma-convergence , local minimizer

1 Introduction

We seek an understanding of the energy landscape for the one-dimensional variational problem

inf𝒜αEε(u),\inf_{\mathcal{A}_{\alpha}}E_{\varepsilon}(u), (1.1)

where u:[0,1]2u:[0,1]\to\mathbb{R}^{2} so that u=(u1,u2)u=(u_{1},u_{2}) with

Eε(u1,u2)=01ε2|u|2+14ε(|u|21)2+L2(u1u2u2u12πN)2dx,\displaystyle E_{\varepsilon}(u_{1},u_{2})=\int_{0}^{1}\frac{\varepsilon}{2}\left|{u^{\prime}}\right|^{2}+\frac{1}{4\varepsilon}(\left|{u}\right|^{2}-1)^{2}+\frac{L}{2}(u_{1}\,u_{2}^{\prime}-u_{2}\,u_{1}^{\prime}-2\pi N)^{2}\,dx, (1.2)
and
𝒜α:={uH1((0,1);2):u(0)=1,u(1)=eiα},\displaystyle\mathcal{A}_{\alpha}:=\{u\in H^{1}((0,1);\mathbb{R}^{2}):\;u(0)=1,\;u(1)=e^{i\alpha}\}, (1.3)

for some positive integer NN and some α[0,2π)\alpha\in[0,2\pi)

When convenient, as above, we will view u=(u1,u2)u=(u_{1},u_{2}) as a map into \mathbb{C}. On occasion we will also find it convenient to use the following notation for the twist term:

𝒯(u):=u1u2u2u1.\mathcal{T}(u):=u_{1}\,u_{2}^{\prime}-u_{2}\,u_{1}^{\prime}.

Our purpose in this article is to continue the analysis of a family of models with disparate elastic constants arising in the mathematics of liquid crystals [5, 6, 7, 8]. In particular, the problem (1.1) can be viewed as a highly simplified, relaxed version of the Oseen-Frank model for cholesteric liquid crystals, [2, 13, 20, 21, 22, 23] based on the elastic deformations of an 𝕊1\mathbb{S}^{1}- or 𝕊2\mathbb{S}^{2}-valued director nn, cf. [24]. Other models, of course, exist for nematic liquid crystals, including the QQ-tensor based Landau-de Gennes model, whose energy density consists of a bulk potential favoring either a uniaxial nematic state, an isotropic state, or both, depending on temperature, cf. [16]. We refer the reader to the recent literature [5, 12] that establishes a precise asymptotic relationship between the Oseen-Frank and the Landau-de Gennes models.

We recall now the form of the Oseen-Frank energy,

FOF(n):=\displaystyle F_{OF}(n):= Ω(K12(divn)2+K22((curln)n+q)2+K32|(curln)×n|2\displaystyle\int_{\Omega}\left(\frac{K_{1}}{2}(\mathrm{div}\,n)^{2}+\frac{K_{2}}{2}((\mathrm{curl}\,n)\cdot n+q)^{2}+\frac{K_{3}}{2}|(\mathrm{curl}\,n)\times n|^{2}\right.
+K2+K42(tr(n)2(divn)2))dx,\displaystyle\left.\quad\quad+\frac{K_{2}+K_{4}}{2}(\textup{tr}\,(\nabla n)^{2}-(\mathrm{div}\,n)^{2})\right)\,dx, (1.4)

where Ω3\Omega\subset\mathbb{R}^{3} represents the sample domain and the director nn maps Ω\Omega to 𝕊2\mathbb{S}^{2}. The material constants K1,K2,K3K_{1},K_{2},K_{3} and K4K_{4} are the elastic coefficients associated with the deformations of splay, twist, bend and saddle-splay, respectively [24]. Most important for this article is the second term, the twist, where q=2πpq=\frac{2\pi}{p} with pp being the pitch of the cholesteric helix. The distinction between nematic and cholesteric liquid crystals is manifested by the value of qq. The liquid crystal is in a nematic state when q=0q=0 and, absent boundary conditions, a global minimizer of FOFF_{OF} is a constant director field. On the other hand, a liquid crystal is in a cholesteric state whenever q0q\neq 0 and global minimizers of FOFF_{OF} in 3\mathbb{R}^{3} are rigid rotations of a uniformly twisted director field n=(nx,ny,0)=e2πizpn=(n_{x},n_{y},0)=e^{\frac{2\pi iz}{p}}.

In [8] we propose and analyze a model problem for nematic liquid crystals carrying a large energetic cost for splay. The model couples the Ginzburg-Landau potential to an elastic energy density with large elastic disparity, namely

infuH1(Ω;2)12Ω(ε|u|2+L(divu)2+1ε(1|u|2)2)𝑑x.\inf_{u\in H^{1}(\Omega;\mathbb{R}^{2})}\frac{1}{2}\int_{\Omega}\left(\varepsilon|\nabla u|^{2}+L(\mathrm{div}\,u)^{2}+\frac{1}{\varepsilon}(1-|u|^{2})^{2}\right)\,dx. (1.5)

Here one should view LL as playing a role analogous to K1K_{1} in (1.4). The minimization is taken over competitors satisfying an 𝕊1\mathbb{S}^{1}-valued Dirichlet condition on Ω\partial\Omega so as to avoid a trivial minimizer. This choice of potential clearly favors 𝕊1\mathbb{S}^{1}-valued states, which are a stand-in in our models for uniaxial nematic states. Analysis of (1.5) in the ε0\varepsilon\to 0 limit involves a ‘wall energy’ along a jump set JuJ_{u} penalizing jumps of any 𝕊1\mathbb{S}^{1}-valued competitor uu, and bulk elastic energy favoring low divergence. The conjectured Γ\Gamma-limit of (1.5) is

L2Ω(divu)2𝑑x+16JuΩ|u+u|3𝑑1,\frac{L}{2}\int_{\Omega}(\mathrm{div}\,u)^{2}\,dx+\frac{1}{6}\int_{J_{u}\cap\Omega}|u_{+}-u_{-}|^{3}\,d\mathcal{H}^{1}, (1.6)

where u+u_{+} and uu_{-} are the one-sided traces of uu along JuJ_{u} which exhibit a jump discontinuity in their tangential components.

The model considered in this paper is a cholesteric analog of the problem in [8]. Just as the functional considered in [8] can be viewed as a Ginzburg-Landau-type relaxation of the splay K1K_{1}-term in (1.4), the problem (1.1) can be understood as a similar relaxation of the twist K2K_{2}-term in the same energy. For example, in 2D this relaxation may take the form

inf𝒜Eε2D(u),\inf_{\mathcal{A}}E^{2D}_{\varepsilon}(u), (1.7)

where u:Ω3u:\Omega\to\mathbb{R}^{3} with

Eε2D(u)=Ωε2|u|2+14ε(|u|21)2+L2(ucurlu2πN)2dx,\displaystyle E^{2D}_{\varepsilon}(u)=\int_{\Omega}\frac{\varepsilon}{2}\left|{\nabla u}\right|^{2}+\frac{1}{4\varepsilon}(\left|{u}\right|^{2}-1)^{2}+\frac{L}{2}(u\cdot\mathrm{curl}\,{u}-2\pi N)^{2}\,dx, (1.8)
and
𝒜:={uH1(Ω;3):u|Ω=u0},\displaystyle\mathcal{A}:=\{u\in H^{1}(\Omega;\mathbb{R}^{3}):\;u|_{\partial\Omega}=u_{0}\}, (1.9)

for some domain Ω2\Omega\subset\mathbb{R}^{2}, some positive integer NN and boundary condition u0:Ω𝕊2u_{0}:\partial\Omega\to\mathbb{S}^{2}. Results of simulations for the gradient flow dynamics associated with the problem (1.7) lead to intricate textures, such as that shown in Fig. 1, resembling cholesteric fingerprint textures observed in experiments [17].

While attempting to tackle the problem (1.7), we found that the energy landscape in (1.1) is already rich enough to merit a separate investigation in one dimension that we undertake in this paper. We further assume that the component of uu along the axis of the twist vanishes so that the target space for the director is two-dimensional. Thus, though we will write u=(u1(x),u2(x))u=(u_{1}(x),u_{2}(x)) what we really have in mind is u=(0,u2(x),u3(x)).u=(0,u_{2}(x),u_{3}(x)). The thought experiment that allows us to impose this condition assumes that an electric field is applied along the axis of the twist and that the cholesteric has negative dielectric anisotropy that forces its molecules to orient perpendicular to the field, [11].

Existence and stability of minimizers for the three-component cholesteric director within the framework of the Oseen-Frank model in one dimension was considered in [1] and [4] under the assumption that all elastic constants have comparable values. In addition, in [4], the energy functional included the effects of an electric field. In the one-dimensional setting for highly disparate elastic constants, it turns out the inclusion of a third xx-dependent component leads to an energy where distinguishing textures are lost for ε1\varepsilon\ll 1 and the energy landscape becomes highly degenerate, see Remark 3.4. Thus, we find that the one-dimensional, two-component model (1.2) leads to stable states more reminiscent of those described above for the two-dimensional problem.

Refer to caption
Figure 1: Numerical solution for the gradient flow associated with (1.7) obtained in COMSOL [3]. The arrows represent the director uu, the blue and the red curves are level sets u3=0.92u_{3}=-0.92 and u3=0.92u_{3}=0.92, respectively. The simulation was started from a uniform twist state with the axis of the twist oriented in a vertical direction. The director is assumed to be oriented to the right and to the left on the top and the bottom boundaries, respectively. Periodic boundary conditions are imposed on vertical components of the boundary. Here N=10N=10, L=1L=1, and ε=0.005\varepsilon=0.005.

The richness of the energy landscape is first revealed in Section 2 where the key result is Theorem 2.3, showing that local minimizers of EεE_{\varepsilon} exist for every positive integer value of twist–essentially for every winding number.

Section 3 contains our principal result of this investigation, namely that similar to our work on (1.5) in [8], the Γ\Gamma-limit E0E_{0} given by (3.4) of the relaxed energy EεE_{\varepsilon} is the twist energy defined over 𝕊1\mathbb{S}^{1}-valued maps along with a jump energy, cf. Theorems 3.1 and 3.2. One distinction, however, between our Γ\Gamma-limit here and (1.6) is that in the present study the jump cost, now associated with jumps in the phase, is impervious to the size of the jump. We demonstrate in Theorem 3.5 and Corollary 3.6 that in certain parameter regimes depending on LL and α\alpha, global energy minimizers with jumps are energetically favorable. Indeed, this is the most dramatic effect of the assumption of disparate elastic constants present in our model. The relatively expensive cost of twist leads the global minimizer of (1.1), which of course is necessarily smooth, to rapidly change its phase, a process that can only be achieved with finite energetic cost by having the modulus simultaneously plunge towards zero.

In Section 4 we establish an energy barrier between the local minimizers of different winding numbers exposed in Theorem 2.3, cf. Theorem 4.1. This readily leads to the existence of saddle points in Theorem 4.2 via the Mountain Pass Theorem, thus filling out the energy landscape for EεE_{\varepsilon}.

Finally, in Section 5 we investigate the energy (5.1) motivated by studies of so-called twist bend nematics, where twisting of the director occurs at much shorter scales than in cholesterics [18]. Here we model this situation by tying the pitch (or the period of the twist) 1/N1/N to the Ginzburg-Landau parameter ε\varepsilon so that twisting “averages out” in the limit ε0\varepsilon\to 0. We show in Theorem 5.2 that, in fact, the weak limit of uniformly energy bounded director fields is equal to zero but we are nonetheless able to recover some information about fine scale behavior of these fields. Then in Theorems 5.3 and 5.4 we establish Γ\Gamma-convergence in this setting.

2 Global and local minimizers that stay bounded away from zero

We begin with the observation for problem (1.2)-(1.3) that a global minimizer exists for fixed ε>0\varepsilon>0.

Theorem 2.1.

For each fixed ε>0\varepsilon>0 there exists a minimizer of EεE_{\varepsilon} within the class 𝒜α.\mathcal{A}_{\alpha}.

Proof.

Existence follows readily from the direct method as follows. Suppressing the ε\varepsilon-dependence, let {uj}={(u1j,u2j)}\{u^{j}\}=\{(u_{1}^{j},u_{2}^{j})\} denote a minimizing sequence:

Eε(u1j,u2j)m:={infEε(u):u𝒜α}.E_{\varepsilon}(u_{1}^{j},u_{2}^{j})\to m:=\{\inf E_{\varepsilon}(u):\;u\in\mathcal{A}_{\alpha}\}.

Compactness of a minimizing sequence follows from the immediate energy bounds

01|uj|2dx<C,01|uj|4dx<C,01(u1ju2ju2ju1j)2dx<C.\int_{0}^{1}\left|{u^{j}\,{}^{\prime}}\right|^{2}\,dx<C,\quad\int_{0}^{1}\left|{u^{j}}\right|^{4}\,dx<C,\quad\int_{0}^{1}\big{(}u_{1}^{j}u_{2}^{j}\,{}^{\prime}-u_{2}^{j}u_{1}^{j}\,{}^{\prime}\big{)}^{2}\,dx<C.

So, in particular we have a uniform H1H^{1}-bound on {uj}\{u^{j}\}. Thus, up to subsequences, we get uniform (in fact Holder) convergence of uju¯=(u¯1,u¯2),u^{j}\to\bar{u}=(\bar{u}_{1},\bar{u}_{2}), and uju¯u^{j}\,{}^{\prime}\rightharpoonup\bar{u}^{\prime} weakly in L2((0,1))L^{2}((0,1)) for some u¯𝒜α.\bar{u}\in\mathcal{A}_{\alpha}.

Turning to the issue of lower-semicontinuity, we note that verification for the first two terms in EεE_{\varepsilon} is standard. For the third term we observe that

u1ju2ju2ju1ju¯1u¯2u¯2u¯1weakly inL2,u_{1}^{j}u_{2}^{j}\,{}^{\prime}-u_{2}^{j}u_{1}^{j}\,{}^{\prime}\rightharpoonup\bar{u}_{1}\bar{u}_{2}\,^{\prime}-\bar{u}_{2}\bar{u}_{1}\,^{\prime}\;\mbox{weakly in}\;L^{2},

through the pairing of weak L2L^{2} and uniform convergence.

Then we have

01(u1ju2ju2ju1j2πN)2dx=\displaystyle\int_{0}^{1}(u_{1}^{j}\,u_{2}^{j}\,{}^{\prime}-u_{2}^{j}\,u_{1}^{j}\,{}^{\prime}-2\pi N)^{2}\,dx=
01(u1ju2ju2ju1j)2dx4πN01(u1ju2ju2ju1j)dx+4π2N2.\displaystyle\int_{0}^{1}(u_{1}^{j}\,u_{2}^{j}\,{}^{\prime}-u_{2}^{j}\,u_{1}^{j}\,{}^{\prime})^{2}\,dx-4\pi N\int_{0}^{1}\big{(}u_{1}^{j}\,u_{2}^{j}\,{}^{\prime}-u_{2}^{j}\,u_{1}^{j}\,{}^{\prime}\big{)}\,dx+4\pi^{2}N^{2}.

The middle term is continuous given the strong convergence of uju^{j} to u¯.\bar{u}. For the first term, we appeal to the lower-semicontinuity of the L2L^{2} norm under weak L2L^{2} convergence. Thus, Eε(u¯)=m.E_{\varepsilon}(\bar{u})=m.

It turns out that characterization of the global minimizer in the case where α=0\alpha=0, so that the boundary conditions are simply u(0)=u(1)=1u(0)=u(1)=1, is much simpler than when α(0,2π)\alpha\in(0,2\pi). In particular, we have the following result.

Theorem 2.2.

Let uεu_{\varepsilon} denote a global minimizer of EεE_{\varepsilon} within the admissible class 𝒜0\mathcal{A}_{0}. Then ρε(x):=|uε(x)|\rho_{\varepsilon}(x):=\left|{u_{\varepsilon}(x)}\right| converges to 11 uniformly on [0,1][0,1] as ε0.\varepsilon\to 0.

Proof.

We proceed by contradiction and assume that for some δ>0\delta>0 there exists a sequence εj0\varepsilon_{j}\to 0 and values xj[0,1]x_{j}\in[0,1] such that

ρεj(xj)1δ.\rho_{\varepsilon_{j}}(x_{j})\leq 1-\delta.

The case where ρεj(xj)1+δ\rho_{\varepsilon_{j}}(x_{j})\geq 1+\delta is handled similarly.

We begin with the observation that

Eε(uε)Eε(ei2πNx)=2(πN)2ε.E_{\varepsilon}(u_{\varepsilon})\leq E_{\varepsilon}(e^{i2\pi Nx})=2(\pi N)^{2}\varepsilon. (2.1)

It then follows that for some C0>0C_{0}>0 independent of ε\varepsilon one has

01(ρε)2+ρε4dx<C0,\int_{0}^{1}(\rho_{\varepsilon}^{\prime})^{2}\,+\rho_{\varepsilon}^{4}\,dx<C_{0},

which in turn implies a bound of the form

ρεH1(0,1)<C1=C1(C0).Hence,ρεC0,1/2(0,1)<C1.\left\|\rho_{\varepsilon}\right\|_{H^{1}(0,1)}<C_{1}=C_{1}(C_{0}).\quad\mbox{Hence,}\quad\left\|\rho_{\varepsilon}\right\|_{C^{0,1/2}(0,1)}<C_{1}.

Then invoking the Hölder bound above, we have

|ρε(x)ρε(xj)|C1|xxj|1/2\left|{\rho_{\varepsilon}(x)-\rho_{\varepsilon}(x_{j})}\right|\leq C_{1}\left|{x-x_{j}}\right|^{1/2}

and so for |xxj|(δ2C1)2\left|{x-x_{j}}\right|\leq\big{(}\frac{\delta}{2C_{1}}\big{)}^{2} one would have

ρε(x)ρε(xj)+C1|xxj|1/21δ2.\rho_{\varepsilon}(x)\leq\rho_{\varepsilon}(x_{j})+C_{1}\left|{x-x_{j}}\right|^{1/2}\leq 1-\frac{\delta}{2}.

This in turn would imply

Eε(uε)14ε01(ρε21)2𝑑x\displaystyle E_{\varepsilon}(u_{\varepsilon})\geq\frac{1}{4\varepsilon}\int_{0}^{1}(\rho_{\varepsilon}^{2}-1)^{2}\,dx 14ε{x:|xxj|(δ2C1)2}(ρε21)2𝑑x\displaystyle\geq\frac{1}{4\varepsilon}\int_{\left\{x:\,\left|{x-x_{j}}\right|\leq\big{(}\frac{\delta}{2C_{1}}\big{)}^{2}\right\}}(\rho_{\varepsilon}^{2}-1)^{2}\,dx
δ464C12ε.\displaystyle\geq\frac{\delta^{4}}{64C_{1}^{2}\varepsilon}.

This cannot hold in light of (2.1) for ε<ε0\varepsilon<\varepsilon_{0} where

ε0=δ282C1πN.\varepsilon_{0}=\frac{\delta^{2}}{8\sqrt{2}C_{1}\pi N}.


Next we turn to the construction of local minimizers of EεE_{\varepsilon} within the class 𝒜α\mathcal{A}_{\alpha} for α[0,2π)\alpha\in[0,2\pi). Like the global minimizers constructed for the case α=0\alpha=0 in Theorem 2.2, the modulus of these local minimizers will converge uniformly to 11 as ε0\varepsilon\to 0.

Theorem 2.3.

For every positive integer MM and every α[0,2π)\alpha\in[0,2\pi), there exists an ε0>0\varepsilon_{0}>0 such that for all ε<ε0\varepsilon<\varepsilon_{0} there is an H1H^{1}-local minimizer uε,M=ρε,Meiθε,Mu_{\varepsilon,M}=\rho_{\varepsilon,M}e^{i\theta_{\varepsilon,M}} of EεE_{\varepsilon} within the class 𝒜α\mathcal{A}_{\alpha} such that

lim supε0ρε,M1L(0,1)ε<,\displaystyle\limsup_{\varepsilon\to 0}\frac{\left\|{\rho_{\varepsilon,M}}-1\right\|_{L^{\infty}(0,1)}}{\varepsilon}<\infty, (2.2)
limε0θε,M=2πM+αuniformly inx[0,1],\displaystyle\lim_{\varepsilon\to 0}\theta_{\varepsilon,M}^{\prime}=2\pi M+\alpha\;\mbox{uniformly in}\;x\in[0,1], (2.3)
and
limε0Eε(uε,M)=L2(2π(MN)+α)2.\displaystyle\lim_{\varepsilon\to 0}E_{\varepsilon}(u_{\varepsilon,M})=\frac{L}{2}\left(2\pi(M-N)+\alpha\right)^{2}. (2.4)
Remark 2.4.

We will find later that in some parameter regimes, corresponding to α\alpha small and M=NM=N, these local minimizers turn out in fact to be global minimizers. However, when MNM\not=N or when M=NM=N but α\alpha exceeds a critical value, they will not.

Proof.

To capture these local minimizers we will rephrase our problem by switching to polar coordinates via the substitution

u1=ρcosθ,u2=ρsinθ.u_{1}=\rho\cos\theta,\quad u_{2}=\rho\sin\theta.

The boundary conditions corresponding to (1.3) are

ρ(0)=1=ρ(1),θ(0)=0,θ(1)=2πM+αfor some integerM>0.\rho(0)=1=\rho(1),\quad\theta(0)=0,\;\theta(1)=2\pi M+\alpha\quad\mbox{for some integer}\;M>0. (2.5)

We find that in these variables,

Eε=Eε(ρ,θ)=01ε2((ρ)2+ρ2(θ)2)+14ε(ρ21)2+L2(ρ2θ2πN)2dx.E_{\varepsilon}=E_{\varepsilon}(\rho,\theta)=\int_{0}^{1}\frac{\varepsilon}{2}\big{(}(\rho^{\prime})^{2}+\rho^{2}(\theta^{\prime})^{2}\big{)}+\frac{1}{4\varepsilon}(\rho^{2}-1)^{2}+\frac{L}{2}(\rho^{2}\theta^{\prime}-2\pi N)^{2}\,dx.

We will minimize Eε(ρ,θ)E_{\varepsilon}(\rho,\theta) subject to (2.5) via a constrained minimization procedure. To this end, for any number ρ0(0,1)\rho_{0}\in(0,1) we introduce the admissible class

ρ0:={ρH1(0,1):ρ(0)=1=ρ(1),ρ(x)ρ0on[0,1]}\mathcal{H}_{\rho_{0}}:=\{\rho\in H^{1}(0,1):\;\rho(0)=1=\rho(1),\;\rho(x)\geq\rho_{0}\;\mbox{on}\;[0,1]\} (2.6)

and for any positive integer MM and any α[0,2π)\alpha\in[0,2\pi) we denote

M,α:={θH1(0,1):θ(0)=0,θ(1)=2πM+α}.\mathcal{H}_{M,\alpha}:=\{\theta\in H^{1}(0,1):\;\theta(0)=0,\;\theta(1)=2\pi M+\alpha\}. (2.7)

We note that for each fixed ε>0\varepsilon>0 and ρ0(0,1)\rho_{0}\in(0,1), the direct method provides for a minimizing pair (ρε,M,θε,M)({\rho_{\varepsilon,M}},{\theta_{\varepsilon,M}}) to the constrained problem:

με,M:=infρρ0,θM,αEε(ρ,θ).{\mu_{\varepsilon,M}}:=\inf_{\rho\in\mathcal{H}_{\rho_{0}},\,\theta\in\mathcal{H}_{M,\alpha}}E_{\varepsilon}(\rho,\theta). (2.8)

The only point to be made here is that the lower bound ρjρ0\rho_{j}\geq\rho_{0} on a minimizing sequence {ρj,θj}\{\rho_{j},\theta_{j}\} allows for H1H^{1} control of {θj}\{\theta_{j}\}. Also the H1H^{1} control on {ρj}\{\rho_{j}\} yields uniform convergence of a subsequence so that the constraint is satisfied by the limiting ρε,M{\rho_{\varepsilon,M}}.

We remark for later use that με,M{\mu_{\varepsilon,M}} is bounded independent of ε\varepsilon since

με,MEε(1,(2πM+α)x)=L2(2π(MN)+α)2+O(ε){\mu_{\varepsilon,M}}\leq E_{\varepsilon}(1,(2\pi M+\alpha)x)=\frac{L}{2}\left(2\pi(M-N)+\alpha\right)^{2}+O(\varepsilon) (2.9)

We will now argue that for any integer M>0M>0 and any ρ0(0,1)\rho_{0}\in(0,1), these solutions to the constrained problem in fact satisfy ρε,M(x)>ρ0\rho_{\varepsilon,M}(x)>\rho_{0} for all x[0,1]x\in[0,1] when ε\varepsilon is sufficiently small. Hence, they correspond to H1H^{1}-local minimizers of Eε(u)E_{\varepsilon}(u) subject to the boundary conditions (1.3) since the representation uε,M=ρε,Meiθε,Mu_{\varepsilon,M}=\rho_{\varepsilon,M}e^{i\theta_{\varepsilon,M}} is global.

CLAIM: For any positive integer MM, any α[0,2π)\alpha\in[0,2\pi), and any ρ0(0,1)\rho_{0}\in(0,1) we have

ρε,M(x)>ρ0for allx[0,1]providedεis sufficiently small.\rho_{\varepsilon,M}(x)>\rho_{0}\;\mbox{for all}\;x\in[0,1]\quad\mbox{provided}\;\varepsilon\;\mbox{is sufficiently small.} (2.10)

To pursue this claim, we first observe that since the constraint falls only on ρε,M{\rho_{\varepsilon,M}}, this minimizing pair (ρε,M,θε,M)({\rho_{\varepsilon,M}},{\theta_{\varepsilon,M}}) must satisfy

limt0+Eε(ρε,M+tf,θε,M)Eε(ρε,M,θε,M)t0,\lim_{t\to 0^{+}}\frac{E_{\varepsilon}\big{(}\rho_{\varepsilon,M}+tf,\theta_{\varepsilon,M}\big{)}-E_{\varepsilon}\big{(}\rho_{\varepsilon,M},\theta_{\varepsilon,M}\big{)}}{t}\geq 0, (2.11)

for all fH01(0,1)f\in H^{1}_{0}(0,1) such that f(x)0f(x)\geq 0 on [0,1][0,1], and

ddtt=0Eε(ρε,M,θε,M+tψ)=0for allψH01(0,1).\frac{d}{dt}_{t=0}E_{\varepsilon}\left({\rho_{\varepsilon,M}},{\theta_{\varepsilon,M}}+t\psi\right)=0\quad\mbox{for all}\;\psi\in H^{1}_{0}(0,1). (2.12)

Computing these quantities we find that (2.11) takes the form

01ερε,Mf+(ερε,M(θε,M)2+1ε(ρε,M21)ρε,M2L(2πNρε,M2θε,M)ρε,Mθε,M)fdx0\int_{0}^{1}\varepsilon{\rho^{\prime}_{\varepsilon,M}}f^{\prime}+\left(\varepsilon{\rho_{\varepsilon,M}}{\left({\theta^{\prime}_{\varepsilon,M}}\right)}^{2}+\frac{1}{\varepsilon}\left(\rho^{2}_{\varepsilon,M}-1\right){\rho_{\varepsilon,M}}\right.\\ \left.-2L\left(2\pi N-{\rho^{2}_{\varepsilon,M}}{\theta^{\prime}_{\varepsilon,M}}\right){\rho_{\varepsilon,M}}{\theta^{\prime}_{\varepsilon,M}}\right)f\,dx\geq 0 (2.13)

for all nonnegative fH01(0,1)f\in H^{1}_{0}(0,1), and (2.12) takes the form

[(εθε,ML(2πNρε,M2θε,M))ρε,M2]=0.\left[\left(\varepsilon\theta^{\prime}_{\varepsilon,M}-L\left(2\pi N-{\rho^{2}_{\varepsilon,M}}{\theta^{\prime}_{\varepsilon,M}}\right)\right){\rho^{2}_{\varepsilon,M}}\right]^{\prime}=0. (2.14)

Thus,

(εθε,ML(2πNρε,M2θε,M))ρε,M2=Cεfor some constantCε,\big{(}\varepsilon{\theta^{\prime}_{\varepsilon,M}}-L(2\pi N-{\rho^{2}_{\varepsilon,M}}{\theta^{\prime}_{\varepsilon,M}})\big{)}{\rho^{2}_{\varepsilon,M}}=C_{\varepsilon}\quad\mbox{for some constant}\;C_{\varepsilon}, (2.15)

allowing us to solve for θε,M{\theta^{\prime}_{\varepsilon,M}} to find

θε,M=2πNLρε,M2+CεLρε,M4+ερε,M2.{\theta^{\prime}_{\varepsilon,M}}=\frac{2\pi NL{\rho^{2}_{\varepsilon,M}}+C_{\varepsilon}}{L{\rho^{4}_{\varepsilon,M}}+\varepsilon{\rho^{2}_{\varepsilon,M}}}. (2.16)

Integrating (2.16) over the interval [0,1][0,1] and using the boundary conditions on θε,M{\theta_{\varepsilon,M}} we obtain a formula for CεC_{\varepsilon}:

Cε=2πM+α2πLN01(Lρε,M2+ε)1𝑑x01(Lρε,M4+ερε,M2)1𝑑x.C_{\varepsilon}=\frac{2\pi M+\alpha-2\pi LN\int_{0}^{1}(L{\rho^{2}_{\varepsilon,M}}+\varepsilon)^{-1}\,dx}{\int_{0}^{1}(L{\rho^{4}_{\varepsilon,M}}+\varepsilon{\rho^{2}_{\varepsilon,M}})^{-1}\,dx}. (2.17)

Now by (2.9),

01(ρε,M21)|ρε,M|𝑑x201ε2(ρε,M)2+14ε(ρε,M21)2dx2με,M.\int_{0}^{1}({\rho^{2}_{\varepsilon,M}}-1)\left|{{\rho^{\prime}_{\varepsilon,M}}}\right|\,dx\leq\sqrt{2}\int_{0}^{1}\frac{\varepsilon}{2}({\rho^{\prime}_{\varepsilon,M}})^{2}+\frac{1}{4\varepsilon}({\rho^{2}_{\varepsilon,M}}-1)^{2}\,dx\leq\sqrt{2}\,{\mu_{\varepsilon,M}}.

Since ρε,M(0)=1{\rho_{\varepsilon,M}}(0)=1, it then follows from (2.9) and this total variation bound that ρε,M{\rho_{\varepsilon,M}} is bounded above uniformly in ε\varepsilon. Thus, by (2.17), the same is true of |Cε|\left|{C_{\varepsilon}}\right|.

Next we use (2.16) to find that

θε,M(2πNL+CεL+ε)=2πNLρε,M2+CεLρε,M4+ερε,M2(2πNL+CεL+ε)\displaystyle{\theta^{\prime}_{\varepsilon,M}}-\bigg{(}\frac{2\pi NL+C_{\varepsilon}}{L+\varepsilon}\bigg{)}=\frac{2\pi NL{\rho^{2}_{\varepsilon,M}}+C_{\varepsilon}}{L{\rho^{4}_{\varepsilon,M}}+\varepsilon{\rho^{2}_{\varepsilon,M}}}-\bigg{(}\frac{2\pi NL+C_{\varepsilon}}{L+\varepsilon}\bigg{)}
=(2πNL2ρε,M2+Cε[L(1+ρε,M2)+ε]ρε,M2(Lρε,M2+ε)(L+ε))(1ρε,M2)=:Λε(1ρε,M2)\displaystyle=\bigg{(}\frac{2\pi NL^{2}{\rho^{2}_{\varepsilon,M}}+C_{\varepsilon}\left[L(1+{\rho^{2}_{\varepsilon,M}})+\varepsilon\right]}{{\rho^{2}_{\varepsilon,M}}(L{\rho^{2}_{\varepsilon,M}}+\varepsilon)(L+\varepsilon)}\bigg{)}(1-{\rho^{2}_{\varepsilon,M}})=:\Lambda_{\varepsilon}(1-{\rho^{2}_{\varepsilon,M}})

where |Λε|C=C(N,M,L)\left|{\Lambda_{\varepsilon}}\right|\leq C=C(N,M,L) independent of ε\varepsilon by the uniform bounds on CεC_{\varepsilon} and ρε,M{\rho_{\varepsilon,M}}. Hence,

01|θε,M(2πNL+CεL+ε)|C01(1ρε,M2)𝑑x\displaystyle\int_{0}^{1}\left|{{\theta^{\prime}_{\varepsilon,M}}-\bigg{(}\frac{2\pi NL+C_{\varepsilon}}{L+\varepsilon}\bigg{)}}\right|\leq C\int_{0}^{1}(1-{\rho^{2}_{\varepsilon,M}})\,dx
2Cε(0114ε(1ρε,M2)2𝑑x)1/22Cμε,Mε.\displaystyle\leq 2C\sqrt{\varepsilon}\left(\int_{0}^{1}\frac{1}{4\varepsilon}(1-{\rho^{2}_{\varepsilon,M}})^{2}\,dx\right)^{1/2}\leq 2C\sqrt{{\mu_{\varepsilon,M}}}\sqrt{\varepsilon}. (2.18)

Since

2πM+α=01(θε,M(2πNL+CεL+ε))𝑑x+2πNL+CεL+ε2\pi M+\alpha=\int_{0}^{1}\left({\theta^{\prime}_{\varepsilon,M}}-\bigg{(}\frac{2\pi NL+C_{\varepsilon}}{L+\varepsilon}\bigg{)}\right)\,dx+\frac{2\pi NL+C_{\varepsilon}}{L+\varepsilon}

we can then invoke (2.18) to conclude that

Cε=2πL(MN)+Lα+O(ε).C_{\varepsilon}=2\pi L(M-N)+L\alpha+O(\sqrt{\varepsilon}). (2.19)

Substituting this back into (2.16) we find

θε,M=2πLM+Lα+2πLN(ρε,M21)Lρε,M4+ερε,M2+O(ε).{\theta^{\prime}_{\varepsilon,M}}=\frac{2\pi LM+L\alpha+2\pi LN({\rho^{2}_{\varepsilon,M}}-1)}{L\rho_{\varepsilon,M}^{4}+\varepsilon{\rho^{2}_{\varepsilon,M}}}+O(\sqrt{\varepsilon}). (2.20)

With these estimates we can now establish Claim (2.10).

In light of the boundary conditions, we need only consider x(0,1).x\in(0,1). First, suppose by contradiction, that {x:ρε,M=ρ0}\{x:\,{\rho_{\varepsilon,M}}=\rho_{0}\} contains an isolated point x0(0,1)x_{0}\in(0,1). Since the obstacle in (2.8) is smooth, it follows from standard regularity theory of obstacle problems (see e.g. [19]) that ρε,M{\rho_{\varepsilon,M}} makes C1,1C^{1,1} contact with the obstacle y(x)1y(x)\equiv 1. However, we also have that ρε,M{\rho_{\varepsilon,M}} satisfies the Euler-Lagrange equation on either side of x0x_{0}, that is,

ερε,M′′=ερε,M(θε,M)2+1ε(ρε,M21)ρε,M2L(2πNρε,M2θε,M)ρε,Mθε,M\varepsilon\rho^{\prime\prime}_{\varepsilon,M}=\varepsilon{\rho_{\varepsilon,M}}({\theta^{\prime}_{\varepsilon,M}})^{2}+\frac{1}{\varepsilon}({\rho^{2}_{\varepsilon,M}}-1){\rho_{\varepsilon,M}}-2L(2\pi N-{\rho^{2}_{\varepsilon,M}}{\theta^{\prime}_{\varepsilon,M}}){\rho_{\varepsilon,M}}{\theta^{\prime}_{\varepsilon,M}} (2.21)

cf. (2.13). Consequently the limits xx0+x\to x_{0}^{+} and xx0x\to x_{0}^{-} agree for ρε,M′′(x){\rho^{\prime\prime}_{\varepsilon,M}}(x) so we find that in fact ρε,MC2{\rho_{\varepsilon,M}}\in C^{2} in a neighborhood of x0x_{0} with

ρε,M′′(x0)=ε(θε,M(x0))2+1ε(ρ021)ρ02L(2πNθε,M(x0))θε,M(x0).{\rho^{\prime\prime}_{\varepsilon,M}}(x_{0})=\varepsilon({\theta^{\prime}_{\varepsilon,M}}(x_{0}))^{2}+\frac{1}{\varepsilon}(\rho^{2}_{0}-1)\rho_{0}-2L(2\pi N-{\theta^{\prime}_{\varepsilon,M}}(x_{0})){\theta^{\prime}_{\varepsilon,M}}(x_{0}).

Invoking (2.20) evaluated at x=x0x=x_{0}, we see

θε,M2πM+α+2πN(ρ021)ρ04+O(ε){\theta^{\prime}_{\varepsilon,M}}\sim\frac{2\pi M+\alpha+2\pi N(\rho_{0}^{2}-1)}{\rho_{0}^{4}}+O(\sqrt{\varepsilon}) (2.22)

so that

ρε,M′′(x0)1ε(ρ021)ρ0+O(1){\rho^{\prime\prime}_{\varepsilon,M}}(x_{0})\sim\frac{1}{\varepsilon}(\rho^{2}_{0}-1)\rho_{0}+O(1) (2.23)

But since ρε,M{\rho_{\varepsilon,M}} has a minimum at x0x_{0}, this contradicts the requirement that ρε,M′′(x0)0{\rho^{\prime\prime}_{\varepsilon,M}}(x_{0})\geq 0 when ε\varepsilon is sufficiently small.

Next we suppose by way of contradiction that {x:ρε,M=ρ0}\{x:\rho_{\varepsilon,M}=\rho_{0}\} contains an interval I[0,1]I\subset[0,1]. Fix a smooth non-negative function ff compactly supported in II. Then by (2.13) we must have

I(ε(θε,M)2+1ε(ρ021)ρ02L(2πNθε,M)θε,M)f𝑑x0,\int_{I}\bigg{(}\varepsilon({\theta^{\prime}_{\varepsilon,M}})^{2}+\frac{1}{\varepsilon}(\rho^{2}_{0}-1)\rho_{0}-2L(2\pi N-{\theta^{\prime}_{\varepsilon,M}}){\theta^{\prime}_{\varepsilon,M}}\bigg{)}f\,dx\geq 0,

again leading to a contradiction for ε\varepsilon small. Claim (2.10) is established and the local minimality of uε,Mu_{\varepsilon,M} follows.

We remark in passing that for the case M<NM<N, one can establish the stronger statement that in fact ρε,M(x)>1\rho_{\varepsilon,M}(x)>1 for all x(0,1)x\in(0,1) by choosing ρ0=1\rho_{0}=1 in the definition of the constrained set (2.6). Then the same contradiction argument works with (2.22) replaced by

θε,M2πM+α+O(ε){\theta^{\prime}_{\varepsilon,M}}\sim 2\pi M+\alpha+O(\sqrt{\varepsilon})

and (2.23) replaced by

ρε,M′′(x0)2L(2π(NM)α)(2πM+α)+O(ε).{\rho^{\prime\prime}_{\varepsilon,M}}(x_{0})\sim-2L\big{(}2\pi(N-M)-\alpha\big{)}\big{(}2\pi M+\alpha\big{)}+O(\sqrt{\varepsilon}).

Finally, in light of the uniform in ε\varepsilon bound on θε,M\theta_{\varepsilon,M}^{\prime} provided by (2.20), we observe that for any fixed values of MM and NN, the minimizing ρε,M{\rho_{\varepsilon,M}} must satisfy (2.2), since otherwise, a presumed maximum of ρε,M{\rho_{\varepsilon,M}} at x0x_{0} that is bigger than 11 or a presumed minimum that is less than 11 would violate (2.21). Then applying (2.2) to (2.20), we obtain (2.3) as well. We then may conclude that

lim infε0Eε(ρε,M,θε,M)\displaystyle\liminf_{\varepsilon\to 0}E_{\varepsilon}(\rho_{\varepsilon,M},\theta_{\varepsilon,M}) \displaystyle\geq lim infε0L201(ρε,M2θε,M2πN)2𝑑x\displaystyle\liminf_{\varepsilon\to 0}\frac{L}{2}\int_{0}^{1}(\rho_{\varepsilon,M}^{2}\theta_{\varepsilon,M}^{\prime}-2\pi N)^{2}\,dx
=\displaystyle= L2(2π(MN)+α)2,\displaystyle\frac{L}{2}\left(2\pi(M-N)+\alpha\right)^{2},

and so (2.4) follows, in view of (2.9). ∎

3 Γ\Gamma-convergence of EεE_{\varepsilon}

As we shall see, the local minimizers described in Theorem 2.3 are also global minimizers only in certain parameter regimes. In order to fill out the characterization of global minimizers in all parameter regimes, we will turn to the machinery of Γ\Gamma-convergence.

Our candidate for a limiting functional will be infinite unless uH1((0,1)J;S1)u\in H^{1}((0,1)\setminus J;S^{1}) where JJ is a finite collection of points, say 0<x1<x2<<xk<10<x_{1}<x_{2}<\ldots<x_{k}<1 for some non-negative integer kk, along with perhaps x=0x=0 and/or x=1x=1 depending on whether or not the traces of uu satisfy the desired boundary conditions inherited from EεE_{\varepsilon}; that is, we include x=0x=0 in JJ only if u(0+)1u(0^{+})\not=1 and we include x=1x=1 in JJ only if u(1)eiαu(1^{-})\not=e^{i\alpha}. For such a uu we will assume JJ is the minimal such set of points, meaning that if any point in J(0,1)J\cap(0,1) were eliminated, the function uu would no longer represent an H1H^{1} function in the compliment of the smaller set of points. In particular, if uH1((0,1))u\in H^{1}((0,1)) and has the proper traces, then J=.J=\emptyset.

Then we define E0:L2((0,1);2)E_{0}:L^{2}\big{(}(0,1);\mathbb{R}^{2}\big{)}\to\mathbb{R} via

E0(u):={L201(u1u2u2u12πN)2𝑑x+2230(J)ifuH1((0,1)J;S1)+ otherwise.\displaystyle E_{0}(u):=\left\{\begin{array}[]{cc}\displaystyle\frac{L}{2}\int_{0}^{1}\displaystyle(u_{1}\,u_{2}^{\prime}-u_{2}\,u_{1}^{\prime}-2\pi N)^{2}\,dx+\frac{2\sqrt{2}}{3}\mathcal{H}^{0}(J)&\mbox{if}\;u\in H^{1}((0,1)\setminus J;S^{1})\\ \\ +\infty&\mbox{ otherwise}.\end{array}\right. (3.4)

Here 0\mathcal{H}^{0} refers to zero-dimensional Hausdorff measure, i.e. counting measure.

Then we claim:

Theorem 3.1.

{Eε}\{E_{\varepsilon}\} Γ\Gamma-converges to E0E_{0} in L2((0,1);2)L^{2}\big{(}(0,1);\mathbb{R}^{2}\big{)}.

We also have the following compactness result.

Theorem 3.2.

If {uε}ε>0\{u_{\varepsilon}\}_{\varepsilon>0} satisfies

Eε(uε)C0<,E_{\varepsilon}(u_{\varepsilon})\leq C_{0}<\infty, (3.5)

then there exists a function uH1((0,1)J;S1)u\in H^{1}((0,1)\setminus J^{\prime};S^{1}) where JJ^{\prime} is a finite, perhaps empty, set of points in (0,1)(0,1) such that along a subsequence ε0\varepsilon_{\ell}\to 0 one has

uεuinL2((0,1);2).u_{\varepsilon_{\ell}}\to u\;\mbox{in}\;L^{2}\big{(}(0,1);\mathbb{R}^{2}\big{)}. (3.6)

Furthermore, writing u(x)=eiθ(x)u(x)=e^{i\theta(x)} for θH1((0,1)J)\theta\in H^{1}((0,1)\setminus J^{\prime}), we have that for every compact set K(0,1)JK\subset\subset(0,1)\setminus J^{\prime}, there exists an ε0(K)>0\varepsilon_{0}(K)>0 such that for every ε<ε0\varepsilon_{\ell}<\varepsilon_{0} one has |uε|>0\left|{u_{\varepsilon_{\ell}}}\right|>0 on KK and there is a lifting whereby uε(x)=ρε(x)eiθε(x)u_{\varepsilon_{\ell}}(x)=\rho_{\varepsilon_{\ell}}(x)e^{i\theta_{\varepsilon_{\ell}}(x)} on KK, with

θεθweakly inHloc1((0,1)J).\theta_{\varepsilon_{\ell}}\rightharpoonup\theta\;\mbox{weakly in}\;H^{1}_{loc}\big{(}(0,1)\setminus J^{\prime}\big{)}. (3.7)
Remark 3.3.

It is not necessarily the case that JJ^{\prime} is minimal for uu; that is, it can happen that uH1((0,1)J;S1)u\in H^{1}((0,1)\setminus J;S^{1}) for some proper subset JJJ\subset J^{\prime} and in that case it is the minimal such set JJ which one uses to evaluate the Γ\Gamma-limit E0E_{0} at uu. However, one cannot guarantee the validity of (3.7) with JJ^{\prime} replaced by such a minimal JJ. For example, in a neighborhood of, say, x=1/2x=1/2 whose size shrinks with ε\varepsilon, an energy-bounded sequence {uε}\{u_{\varepsilon}\} could undergo a rapid jump in phase by 2π2\pi while the modulus of uεu_{\varepsilon} plunges to zero–or even stays positive but very small– in this neighborhood. Then the limiting uu could have well-behaved lifting across x=1/2x=1/2 while for all ε>0\varepsilon>0, the function uεu_{\varepsilon} would not.

Remark 3.4.

The appearance of a jump set contribution to the Γ\Gamma-limit E0E_{0} is associated with the cost of a Modica-Mortola type transition layer for the modulus from value 11 down to 0 and back, accompanied by a rapid shift in the phase. If one instead considers a three-component model for u=(u1(x),u2(x),u3(x))u=(u_{1}(x),u_{2}(x),u_{3}(x)) then such a phase shift can be achieved with asymptotically vanishing cost by plunging u2(x)2+u3(x)2u_{2}(x)^{2}+u_{3}(x)^{2} to zero while compensating with u1(x)u_{1}(x) to keep |u|1\left|{u}\right|\approx 1. This apparently leads to an absence of local minimizers with such meta-stable states eventually ‘melting’ under a gradient flow to global minimizers given asymptotically by (3.34) of Theorem 3.5 below. In fact, the degeneracy in such a three-component model is worse than just this: If one introduces cylindrical coordinates so that (u1,u2,u3)=(ρcosθ,ρsinθ,u3)(u_{1},u_{2},u_{3})=(\rho\cos\theta,\rho\sin\theta,u_{3}) and then one writes ρ=cosϕ\rho=\cos\phi and u3=sinϕu_{3}=\sin\phi for some angle ϕ(x)\phi(x), a three-component version of EεE_{\varepsilon} would take the form

1201εϕ(x)2+εθ(x)2+L(cos2ϕ(x)θ(x)2πN)2dx.\frac{1}{2}\int_{0}^{1}\varepsilon\phi^{\prime}(x)^{2}+\varepsilon\theta^{\prime}(x)^{2}+L\big{(}\cos^{2}\phi(x)\theta^{\prime}(x)-2\pi N\big{)}^{2}\,dx.

Note then that for ε\varepsilon small there is no control on ϕ\phi^{\prime}, nor is there control on θ\theta^{\prime} when ϕπ/2\phi\approx\pi/2.

We now present the proofs of Theorem 3.1 and Theorem 3.2. We will begin with the proof of Theorem 3.2 since elements of it will be called upon in the proof of Theorem 3.1.

Proof of Theorem 3.2.

We fix an integer q2q\geq 2 and consider a sequence satisfying (3.5). Denoting ρε:=|uε|\rho_{\varepsilon}:=\left|{u_{\varepsilon}}\right|, since uεu_{\varepsilon} is H1H^{1}, we have that ρε\rho_{\varepsilon} is continuous and we may define the open sets

ε:={y[0,1]:ρε(y)>12q}.\mathcal{I}_{\varepsilon}:=\{y\in[0,1]:\rho_{\varepsilon}(y)>1-2^{-q}\}.

As open sets on the real line, each is a countable disjoint union of open intervals

ε=m=1mε=m=1(amε,bmε),\mathcal{I}_{\varepsilon}=\cup_{m=1}^{\infty}\mathcal{I}_{m}^{\varepsilon}=\cup_{m=1}^{\infty}(a_{m}^{\varepsilon},b_{m}^{\varepsilon}),

with

ρε(amε)=ρε(bmε)=12q.\rho_{\varepsilon}(a_{m}^{\varepsilon})=\rho_{\varepsilon}(b_{m}^{\varepsilon})=1-2^{-q}.

Note that by the energy bound (3.5),

𝟙ε𝟙(0,1) in L1((0,1)).\mathbbm{1}_{\mathcal{I}_{\varepsilon}}\to\mathbbm{1}_{(0,1)}\textup{ in }L^{1}((0,1)). (3.8)

Now we consider the open sets

(0,1)¯ε=̊εc.(0,1)\setminus\overline{\mathcal{I}}_{\varepsilon}=\mathring{\mathcal{I}}_{\varepsilon}^{c}.

and similarly decompose ̊εc\mathring{\mathcal{I}}_{\varepsilon}^{c} into a countable union of intervals

m=1(bmε,am+1ε).\cup_{m=1}^{\infty}(b_{m}^{\varepsilon},a_{m+1}^{\varepsilon}).

Now some of the intervals (bmε,am+1ε)(b_{m}^{\varepsilon},a_{m+1}^{\varepsilon}) could contain a point cmεc_{m}^{\varepsilon} such that

ρ(cmε)=2q,\rho(c_{m}^{\varepsilon})=2^{-q},

and we collect those intervals and label them (bmjε,amj+1ε)(b_{m_{j}}^{\varepsilon},a_{m_{j}+1}^{\varepsilon}), where jj belongs to an index set SεS_{\varepsilon}. A priori SεS_{\varepsilon} could be finite or infinite. Let BεB_{\varepsilon} be the union of these “bad intervals.” These are the intervals over which it is possible that a limit of uεu_{\varepsilon} exhibits a jump discontinuity. We first prove that the number of these intervals is finite and bounded uniformly in ε\varepsilon. We observe that

C0\displaystyle C_{0} Bεε2|uε|2+14ε(|uε|21)2dx\displaystyle\geq\int_{B_{\varepsilon}}\frac{\varepsilon}{2}\left|{u_{\varepsilon}^{\prime}}\right|^{2}+\frac{1}{4\varepsilon}(\left|{u_{\varepsilon}}\right|^{2}-1)^{2}\,dx
jSεbmjεcmjεε2(ρε)2+14ε(ρε21)2dy+cmjεamj+1εε2(ρε)2+14ε(ρε21)2dy\displaystyle\geq\sum_{j\in S_{\varepsilon}}\int_{b_{m_{j}}^{\varepsilon}}^{c_{m_{j}}^{\varepsilon}}\frac{\varepsilon}{2}(\rho_{\varepsilon}^{\prime})^{2}+\frac{1}{4\varepsilon}(\rho_{\varepsilon}^{2}-1)^{2}\,dy+\int_{c_{m_{j}}^{\varepsilon}}^{a_{{m_{j}}+1}^{\varepsilon}}\frac{\varepsilon}{2}(\rho_{\varepsilon}^{\prime})^{2}+\frac{1}{4\varepsilon}(\rho_{\varepsilon}^{2}-1)^{2}\,dy
jSεbmjεcmjε|ρε||ρε21|2𝑑y+cmjεamj+1ε|ρε||ρε21|2𝑑y\displaystyle\geq\sum_{j\in S_{\varepsilon}}\int_{b_{m_{j}}^{\varepsilon}}^{c_{m_{j}}^{\varepsilon}}\frac{|\rho_{\varepsilon}^{\prime}||\rho_{\varepsilon}^{2}-1|}{\sqrt{2}}\,dy+\int_{c_{m_{j}}^{\varepsilon}}^{a_{{m_{j}}+1}^{\varepsilon}}\frac{|\rho_{\varepsilon}^{\prime}||\rho_{\varepsilon}^{2}-1|}{\sqrt{2}}\,dy
jSε22q12q|z21|𝑑z.\displaystyle\geq\sum_{j\in S_{\varepsilon}}\sqrt{2}\int_{2^{-q}}^{1-2^{-q}}|z^{2}-1|\,dz. (3.9)

Rearranging (3.9) yields an estimate on the size of SεS_{\varepsilon}:

0(Sε)(22q12q|z21|𝑑z)1C0.\mathcal{H}^{0}(S_{\varepsilon})\leq\left(\sqrt{2}\int_{2^{-q}}^{1-2^{-q}}|z^{2}-1|\,dz\right)^{-1}C_{0}. (3.10)

Next, on (0,1)Bε(0,1)\setminus B_{\varepsilon}, we observe that ρε2q,\rho_{\varepsilon}\geq 2^{-q}, which allows us define a lifting of uεu_{\varepsilon} as ρεeiθε\rho_{\varepsilon}e^{i\theta_{\varepsilon}} and to find a positive constant C1C_{1} such that

(0,1)Bε(θε)2𝑑y\displaystyle\int_{(0,1)\setminus B_{\varepsilon}}(\theta_{\varepsilon}^{\prime})^{2}\,dy C1+C1(0,1)BεL2(ρε2θε2πN)2𝑑y\displaystyle\leq C_{1}+C_{1}\int_{(0,1)\setminus B_{\varepsilon}}\frac{L}{2}(\rho_{\varepsilon}^{2}\theta_{\varepsilon}^{\prime}-2\pi N)^{2}\,dy
C1+C1Eε(uε)C1+C1C0<.\displaystyle\leq C_{1}+C_{1}E_{\varepsilon}(u_{\varepsilon})\leq C_{1}+C_{1}C_{0}<\infty. (3.11)

On each of the (finitely many) intervals comprising (0,1)Bε(0,1)\setminus B_{\varepsilon} we may choose our lifting such that the value of θε\theta_{\varepsilon} at, say, the left endpoint of the interval lies in [0,2π)[0,2\pi) and from the fundamental theorem of calculus and Cauchy-Schwarz it then follows from (3.11) that θεL((0,1)Bε)\left\|\theta_{\varepsilon}\right\|_{L^{\infty}((0,1)\setminus B_{\varepsilon})} is bounded uniformly in ε\varepsilon by a constant depending on C0C_{0} and C1C_{1}. Consequently, we have a bound of the form

θεH1((0,1)Bε)<C2,\left\|\theta_{\varepsilon}\right\|_{H^{1}((0,1)\setminus B_{\varepsilon})}<C_{2}, (3.12)

for some constant C2C_{2} independent of ε\varepsilon.

Now we are going to obtain a subsequence of ε\varepsilon approaching zero along which the bad intervals converge to a finite set of points. To this end, we start with the sequence of all the endpoints of the left-most subinterval in BεB_{\varepsilon} and extract a subsequential limit, calling it x1x_{1}. Then, along this subsequence of εs\varepsilon^{\prime}s, we move on to the left endpoints of the second subinterval of BεB_{\varepsilon}, and passing to a further subsequence, arrive at a limit point x2x_{2}, etc. In light of (3.10), this procedure generates a finite number of points x1<x2<xkx_{1}<x_{2}\ldots<x_{k} in [0,1][0,1]. (If this procedure ever yields xj=xj+1x_{j}=x_{j+1} then we drop xj+1x_{j+1} from this list.) In this manner, we arrive at a subsequence, ε0\varepsilon_{\ell}\to 0 such that:

0(Sε) is independent of  and equal to some fixed k,\mathcal{H}^{0}(S_{\varepsilon_{\ell}})\textup{ is independent of }\ell\textup{ and equal to some fixed }k\in\mathbb{N},

and, in light of (3.8), the subintervals of BεB_{\varepsilon_{\ell}} collapse to these kk points as ε0\varepsilon_{\ell}\to 0; that is

BεJ:={x1,x2,,xk}asε0.B_{\varepsilon_{\ell}}\to J^{\prime}:=\{x_{1},x_{2},\ldots,x_{k}\}\;\mbox{as}\;\varepsilon_{\ell}\to 0. (3.13)

.

If we then fix any finite union of closed intervals K1[0,1]JK_{1}\subset\subset[0,1]\setminus J^{\prime}, it follows from (3.13) that

K1Bε=K_{1}\cap B_{\varepsilon_{\ell}}=\emptyset (3.14)

for ε<ε0\varepsilon<\varepsilon_{0} with ε0=ε0(K1)\varepsilon_{0}=\varepsilon_{0}(K_{1}) small enough. Therefore, uεu_{\varepsilon_{\ell}} has a lifting on the various intervals comprising K1BεK_{1}\cap B_{\varepsilon_{\ell}} and invoking (3.12), we have, after passing to a further subsequence, (with notation suppressed) that

θεθin H1(K1),θεθinL2(K1)\theta_{\varepsilon_{\ell}}\rightharpoonup\theta\quad\textup{in }H^{1}(K_{1}),\;\theta_{\varepsilon_{\ell}}\to\theta\quad\mbox{in}\;L^{2}(K_{1}) (3.15)

for some θH1(K1)\theta\in H^{1}(K_{1}) such that

θH1(K1)C2.\|\theta\|_{H^{1}(K_{1})}\leq C_{2}. (3.16)

Repeating this procedure on a nested sequence of sets

K1K2Kp[0,1]JK_{1}\subset\subset K_{2}\subset\subset\cdots\subset\subset K_{p}\subset\subset\cdots[0,1]\setminus J^{\prime} (3.17)

which exhaust [0,1]J[0,1]\setminus J^{\prime}, and passing to further subsequences via a diagonalization procedure we arrive at a subsequence (still denoted here by ε0\varepsilon_{\ell}\to 0) such that (3.7) holds for some θH1((0,1)J)\theta\in H^{1}\big{(}(0,1)\setminus J^{\prime}\big{)}.

Finally, we define uH1((0,1)J;S1)u\in H^{1}\big{(}(0,1)\setminus J^{\prime};S^{1}\big{)} via u(x):=eiθ(x)u(x):=e^{i\theta(x)} and verify (3.6). The uniform bound (3.5) implies that ρε1\rho_{\varepsilon}\to 1 in L2((0,1))L^{2}((0,1)) and also that

C001|1|ρε|2||ρε|𝑑x|xεy(1ρε2)ρε𝑑x|C_{0}\geq\int_{0}^{1}\left|1-\left|{\rho_{\varepsilon}}\right|^{2}\right|\left|{\rho_{\varepsilon}^{\prime}}\right|\,dx\geq\left|\int_{x_{\varepsilon}}^{y}(1-\rho_{\varepsilon}^{2})\rho_{\varepsilon}^{\prime}\,dx\right| (3.18)

for any y(0,1)y\in(0,1) where xε(0,1)x_{\varepsilon}\in(0,1) is any point selected such that, say, ρε(xε)2.\rho_{\varepsilon}(x_{\varepsilon})\leq 2. It follows that ρεL(0,1)<M\left\|\rho_{\varepsilon}\right\|_{L^{\infty}(0,1)}<M for some M=M(C0)M=M(C_{0}) independent of ε\varepsilon. Hence, for any η>0\eta>0 if we select a compact set K[0,1]JK\subset[0,1]\setminus J^{\prime} such that |[0,1]K|<η\left|{[0,1]\setminus K}\right|<\eta, we can appeal to (3.15) to conclude (3.6) since

lim supl01|uεlu|2𝑑xlim suplK|uεlu|2𝑑x+lim suplKc|uεlu|2𝑑x\displaystyle\limsup_{l\to\infty}\int_{0}^{1}\left|{u_{\varepsilon_{l}}-u}\right|^{2}\,dx\leq\limsup_{l\to\infty}\int_{K}\left|{u_{\varepsilon_{l}}-u}\right|^{2}\,dx+\limsup_{l\to\infty}\int_{K^{c}}\left|{u_{\varepsilon_{l}}-u}\right|^{2}\,dx
lim suplKc|uεlu|2𝑑x2Kc(M2+1)𝑑x<2(M2+1)η.\displaystyle\leq\limsup_{l\to\infty}\int_{K^{c}}\left|{u_{\varepsilon_{l}}-u}\right|^{2}\,dx\leq 2\int_{K^{c}}(M^{2}+1)\,dx<2(M^{2}+1)\eta.

Proof of Theorem 3.1.

We will first assume that uεuu_{\varepsilon}\to u in L2((0,1);2)L^{2}\big{(}(0,1);\mathbb{R}^{2}\big{)} and establish the inequality

lim infEε(uε)E0(u).\liminf E_{\varepsilon}(u_{\varepsilon})\geq E_{0}(u). (3.19)

To this end, we may certainly assume that

lim infEε(uε)C0<for someC0>0,\liminf E_{\varepsilon}(u_{\varepsilon})\leq C_{0}<\infty\quad\mbox{for some}\;C_{0}>0,

since otherwise (3.19) is immediate. Let {uε}\{u_{\varepsilon_{\ell}}\} be a subsequence which achieves the limit inferior. As in (3.9) in the proof of Theorem 3.2, we can then assert that for any integer q2q\geq 2 and up to a further subsequence for which we suppress the notation, one has the lower bound

lim infBεqε2|uε|2+14ε(|uε|21)2dx(22q12q|z21|𝑑z)0(Jq)\liminf_{\ell\to\infty}\int_{B^{q}_{\varepsilon_{\ell}}}\frac{\varepsilon_{\ell}}{2}\left|{u_{\varepsilon_{\ell}}\,^{\prime}}\right|^{2}+\frac{1}{4\varepsilon_{\ell}}(\left|{u_{\varepsilon_{\ell}}}\right|^{2}-1)^{2}\,dx\geq\left(\sqrt{2}\int_{2^{-q}}^{1-2^{-q}}|z^{2}-1|\,dz\,\right)\mathcal{H}^{0}(J^{q}) (3.20)

along with

θεθinHloc1((0,1)Jq).\theta_{\varepsilon_{\ell}}\rightharpoonup\theta\quad\mbox{in}\;H^{1}_{\rm{loc}}\big{(}(0,1)\setminus J^{q}\big{)}. (3.21)

Here we have emphasized the qq dependence to write JqJ^{q} for the finite set of points in [0,1][0,1] and BεqB^{q}_{\varepsilon_{\ell}} for the set of ‘bad intervals’ collapsing to JqJ^{q} over which |uε|\left|{u_{\varepsilon_{\ell}}}\right| dips from values of 12q1-2^{-q} to 2q2^{-q}. Next, we note that for any two positive integers q1<q2q_{1}<q_{2} one has the containment Bεq2Bεq1B^{q_{2}}_{\varepsilon}\subset B^{q_{1}}_{\varepsilon} and so, for any sequence ε0\varepsilon_{\ell}\to 0, the finite set of points arising as the limit of Bεq2B^{q_{2}}_{\varepsilon_{\ell}} must be a subset of the corresponding limit of the finite collection of collapsing intervals comprising Bεq1B^{q_{1}}_{\varepsilon_{\ell}}. Also, since the limiting phase θ\theta of uu will be in Hloc1H^{1}_{loc} of the complement of any such limit of bad intervals, and since JJ is assumed to be the minimal one, we have

0(J)0(Jq)<C1for anyq<,\mathcal{H}^{0}(J)\leq\mathcal{H}^{0}(J^{q})<C_{1}\quad\mbox{for any}\;q<\infty,

for some C1=C1(C0)C_{1}=C_{1}(C_{0}) in light of (3.10). Thus, passing to the limit qq\to\infty in (3.20) gives

lim01ε2|uε|2+14ε(|uε|21)2dx2230(J).\lim_{\ell\to\infty}\int_{0}^{1}\frac{\varepsilon_{\ell}}{2}\left|{u_{\varepsilon_{\ell}}\,^{\prime}}\right|^{2}+\frac{1}{4\varepsilon_{\ell}}(\left|{u_{\varepsilon_{\ell}}}\right|^{2}-1)^{2}\,dx\geq\frac{2\sqrt{2}}{3}\mathcal{H}^{0}(J). (3.22)

Turning to the lower-semi-continuity of the twist term, we can repeat the argument of Theorem 3.2 to obtain that, again up to a further subsequence which we do not notate,

θεθinHloc1((0,1)J~q),\theta_{\varepsilon_{\ell}}\rightharpoonup\theta\quad\mbox{in}\;H^{1}_{\rm{loc}}\big{(}(0,1)\setminus\tilde{J}^{q}\big{)}, (3.23)

where J~q\tilde{J}_{q} is the finite set of points in [0,1][0,1] which is the limit of bad intervals B~εq\tilde{B}_{\varepsilon_{\ell}}^{q} where |uε|12q1|u_{\varepsilon_{\ell}}|\leq 1-2^{-q-1} and dips from 12q11-2^{-q-1} to 12q1-2^{-q}. Of course, it could turn out that J~q=\tilde{J}^{q}=\emptyset, in which case the convergence of θε\theta_{\varepsilon_{\ell}} to θ\theta occurs weakly in Hloc1((0,1)).H^{1}_{\rm{loc}}\big{(}(0,1)\big{)}. We also note that

ρε21 in L2((0,1)),\rho_{\varepsilon}^{2}\to 1\textup{ in }L^{2}((0,1)), (3.24)

which combined with (3.23) implies that for any K(0,1)JqK\subset\subset(0,1)\setminus J^{q}

limKρε2θε𝑑x=Kθ𝑑x.\lim_{\ell\to\infty}\int_{K}\rho_{\varepsilon_{\ell}}^{2}\theta_{\varepsilon_{\ell}}^{\prime}\,dx=\int_{K}\theta^{\prime}\,dx. (3.25)

Then, using (3.25), the weak convergence of θε\theta_{\varepsilon_{\ell}}^{\prime} to θ\theta^{\prime}, and the fact that ρε12q\rho_{\varepsilon_{\ell}}\geq 1-2^{-q} on KK for large \ell, we can estimate

lim inf01(𝒯(uε)2πN)2𝑑x\displaystyle\liminf_{\ell\to\infty}\int_{0}^{1}\displaystyle(\mathcal{T}(u_{\varepsilon_{\ell}})-2\pi N)^{2}\,dx lim infKρε4(θε)24πNρε2θε+4π2N2dx\displaystyle\geq\liminf_{\ell\to\infty}\int_{K}\displaystyle\rho_{\varepsilon_{\ell}}^{4}(\theta_{\varepsilon_{\ell}}^{\prime})^{2}-4\pi N\rho_{\varepsilon_{\ell}}^{2}\theta_{\varepsilon_{\ell}}^{\prime}+4\pi^{2}N^{2}\,dx
lim infK(12q)4(θε)24πNρε2θε+4π2N2dx\displaystyle\geq\liminf_{\ell\to\infty}\int_{K}(1-2^{-q})^{4}(\theta_{\varepsilon_{\ell}}^{\prime})^{2}-4\pi N\rho_{\varepsilon_{\ell}}^{2}\theta_{\varepsilon_{\ell}}^{\prime}+4\pi^{2}N^{2}\,dx
K(12q)4(θ)24πNθ+4π2N2dx.\displaystyle\geq\int_{K}(1-2^{-q})^{4}(\theta^{\prime})^{2}-4\pi N\theta^{\prime}+4\pi^{2}N^{2}\,dx. (3.26)

Choosing larger and larger KK and using that 0(Jq)<\mathcal{H}^{0}(J^{q})<\infty, we find

lim inf01(Tw(uε)2πN)2𝑑x01(12q)4(θ)24πNθ+4π2N2dx.\liminf_{\ell\to\infty}\int_{0}^{1}\displaystyle(Tw(u_{\varepsilon_{\ell}})-2\pi N)^{2}\,dx\geq\int_{0}^{1}(1-2^{-q})^{4}(\theta^{\prime})^{2}-4\pi N\theta^{\prime}+4\pi^{2}N^{2}\,dx.

Finally, sending qq\to\infty yields

lim inf01(𝒯(uε)2πN)2𝑑x01(𝒯(u)2πN)2𝑑x.\liminf_{\ell\to\infty}\int_{0}^{1}\displaystyle(\mathcal{T}(u_{\varepsilon_{\ell}})-2\pi N)^{2}\,dx\geq\int_{0}^{1}(\mathcal{T}(u)-2\pi N)^{2}\,dx. (3.27)

Combining (3.22) with (3.27) completes the proof of lower semi-continuity.

Moving on now to the construction of the recovery sequence for any uL2((0,1);2),u\in L^{2}\big{(}(0,1);\mathbb{R}^{2}\big{)}, if uH1((0,1)J;S1)u\not\in H^{1}((0,1)\setminus J;S^{1}) for any finite set JJ, then E0(u)=E_{0}(u)=\infty and taking the trivial recovery sequence vεuv_{\varepsilon}\equiv u will suffice.

Thus we may assume uH1((0,1)J;S1)u\in H^{1}((0,1)\setminus J;S^{1}) for a finite set JJ and our task is to construct a sequence {vε}H1((0,1);2)\{v_{\varepsilon}\}\subset H^{1}\big{(}(0,1);\mathbb{R}^{2}\big{)} such that

vεuinL2((0,1);2)andlimε0Eε(vε)=E0(u).v_{\varepsilon}\to u\;\mbox{in}\;L^{2}\big{(}(0,1);\mathbb{R}^{2}\big{)}\quad\mbox{and}\quad\lim_{\varepsilon\to 0}E_{\varepsilon}(v_{\varepsilon})=E_{0}(u). (3.28)

In case the traces of uu satisfy the desired boundary conditions for admissibility in EεE_{\varepsilon}, that is, in case u(0+)=1u(0^{+})=1 and u(1)=eiαu(1^{-})=e^{i\alpha} so that x=0x=0 and x=1x=1 do not lie in JJ, our construction will take the form vε=ρεuv_{\varepsilon}=\rho_{\varepsilon}u for a sequence {ρε}H1((0,1);[0,1])\{\rho_{\varepsilon}\}\subset H^{1}\big{(}(0,1);[0,1]\big{)} to be described below. We first describe the construction for this case and then discuss how it is slightly altered in case 0 or 11 lie in JJ. Denoting JJ by {x1,x2,,xk}\{x_{1},x_{2},\ldots,x_{k}\} with then J(0,1)J\subset(0,1) by assumption, we then take ρε\rho_{\varepsilon} to satisfy the following conditions:

  1. (i)

    ρε\rho_{\varepsilon} is smooth on [0,1][0,1].

  2. (ii)

    ρε0\rho_{\varepsilon}\equiv 0 on (xjε2,xj+ε2)(x_{j}-\varepsilon^{2},x_{j}+\varepsilon^{2}).

  3. (iii)

    ρε\rho_{\varepsilon} makes a standard Modica-Mortola style transition from 11 to 0 on Ij1I_{j}^{1}, an interval of size say O(ε)O(\sqrt{\varepsilon}) with right endpoint xjε2x_{j}-\varepsilon^{2}, and makes a transition from 0 back to 11 on an interval of size O(ε)O(\sqrt{\varepsilon}) with left endpoint xj+ε2x_{j}+\varepsilon^{2} that we denote by Ij2I_{j}^{2}, cf. [15].

  4. (iv)

    ρε1\rho_{\varepsilon}\equiv 1 on (0,1)j(Ij1(xjε2,xj+ε2)Ij2)(0,1)\setminus\cup_{j}(I_{j}^{1}\cup(x_{j}-\varepsilon^{2},x_{j}+\varepsilon^{2})\cup I_{j}^{2}).

In case either u(0+)1u(0^{+})\not=1 or u(1)eiαu(1^{-})\not=e^{i\alpha} so that 0 and/or 11 lies in JJ, this procedure must be slightly altered near the endpoints. For example, if u(0+)1u(0^{+})\not=1 then one requires ρε\rho_{\varepsilon} to make a Modica-Mortola style transition from 11 down to 0 on the interval [0,ε][0,\sqrt{\varepsilon}], ρε0\rho_{\varepsilon}\equiv 0 on [ε,ε+ε2][\sqrt{\varepsilon},\sqrt{\varepsilon}+\varepsilon^{2}] and a Modica-Mortola transition from 0 back up to 11 on [ε+ε2,2ε+ε2].[\sqrt{\varepsilon}+\varepsilon^{2},2\sqrt{\varepsilon}+\varepsilon^{2}]. Then we define

θε(x)={0ifx[0,ε+ε2/2)θifx>ε+ε2/2,\theta_{\varepsilon}(x)=\left\{\begin{matrix}0&\;\mbox{if}\;x\in[0,\sqrt{\varepsilon}+\varepsilon^{2}/2)\\ \theta&\;\mbox{if}\;x>\sqrt{\varepsilon}+\varepsilon^{2}/2,\end{matrix}\right.

where u=eiθu=e^{i\theta}, and take vε=ρεeiθε.v_{\varepsilon}=\rho_{\varepsilon}e^{i\theta_{\varepsilon}}. A similar recipe is taken in a neighborhood of x=1x=1 in case u(1)eiα.u(1^{-})\not=e^{i\alpha}.

Computing the transition energy of such a construction is classical and can be found in e.g. [14, 15]. One finds from conditions (ii)-(iv), that

01ε2(ρε)2+14ε(ρε21)2dx2230(J).\int_{0}^{1}\frac{\varepsilon}{2}(\rho_{\varepsilon}^{\prime})^{2}+\frac{1}{4\varepsilon}(\rho_{\varepsilon}^{2}-1)^{2}\,dx\to\frac{2\sqrt{2}}{3}\mathcal{H}^{0}(J).

Furthermore, since θH1((0,1)J)\theta\in H^{1}\big{(}(0,1)\setminus J), and ρε1\rho_{\varepsilon}\to 1 in L2((0,1))L^{2}\big{(}(0,1)\big{)} it is easily seen that

limε001ε2ρε2(θε)2𝑑x=0andlimε0L201𝒯(vε)𝑑x=L2𝒯(u)dx.\lim_{\varepsilon\to 0}\int_{0}^{1}\frac{\varepsilon}{2}\rho_{\varepsilon}^{2}(\theta_{\varepsilon}^{\prime})^{2}\,dx=0\quad\mbox{and}\quad\lim_{\varepsilon\to 0}\frac{L}{2}\int_{0}^{1}\mathcal{T}(v_{\varepsilon})\,dx=\frac{L}{2}\mathcal{T}(u)\,dx.

The proof of (3.28) is complete. ∎

We observe that for uH1((0,1)J;S1)u\in H^{1}((0,1)\setminus J;S^{1}), one has

E0(u)=L201(θ2πN)2𝑑x+2230(J),E_{0}(u)=\frac{L}{2}\int_{0}^{1}(\theta^{\prime}-2\pi N)^{2}\,dx+\frac{2\sqrt{2}}{3}\mathcal{H}^{0}(J), (3.29)

where u=eiθu=e^{i\theta} for θH1((0,1)J)\theta\in H^{1}((0,1)\setminus J). Using this formulation, it is then straight-forward to identify the global minimizers of the Γ\Gamma-limit, and consequently the limits of global minimizers of EεE_{\varepsilon} as well:

Theorem 3.5.

The global minimizer(s) of E0E_{0} are given by:

  1. (i)

    the function

    u(x)=ei(2πN+α)xu(x)=e^{i(2\pi N+\alpha)x} (3.30)

    having constant twist and no jumps when

    Lα2<423 and α[0,π].L\alpha^{2}<\frac{4\sqrt{2}}{3}\textit{ and }\alpha\in[0,\pi]. (3.31)
  2. (ii)

    the function

    u(x)=ei(2π(N1)+α)xu(x)=e^{i(2\pi(N-1)+\alpha)x} (3.32)

    having constant twist and no jumps when

    L(2πα)2<423 and α[π,2π).L(2\pi-\alpha)^{2}<\frac{4\sqrt{2}}{3}\textit{ and }\alpha\in[\pi,2\pi). (3.33)
  3. (iii)

    the one-parameter set of functions given by

    u(x)={ei2πNxifx<x0,ei(2πNx+α)ifx>x0,u(x)=\left\{\begin{matrix}e^{i2\pi Nx}&\;\mbox{if}\;x<x_{0},\\ e^{i(2\pi Nx+\alpha)}&\;\mbox{if}\;x>x_{0},\end{matrix}\right. (3.34)

    for any x0(0,1)x_{0}\in(0,1), that have one jump and twist 2πN2\pi N away from the jump when

    Lα2>423 and L(2πα)2>423.L\alpha^{2}>\frac{4\sqrt{2}}{3}\textit{ and }L(2\pi-\alpha)^{2}>\frac{4\sqrt{2}}{3}. (3.35)

Since any limit of global minimizers of a Γ\Gamma-converging sequence must itself be a global minimizer of the Γ\Gamma-limit, one immediately concludes the following result based on Theorem 3.5 and the compactness result Theorem 3.2:

Corollary 3.6.

Let {uε}\{u_{\varepsilon}\} denote a family of minimizers of EεE_{\varepsilon} subject to the boundary conditions (1.3). Then if (3.35) holds, we have uεuu_{\varepsilon}\to u in L2L^{2} for some uu in the one-parameter family given by (3.34), while if (3.31) or (3.33) holds, there will be a subsequence uεjuu_{\varepsilon_{j}}\to u in L2L^{2} with u=ei(2πN+α)xu=e^{i(2\pi N+\alpha)x} or u=ei(2π(N1)+α)xu=e^{i(2\pi(N-1)+\alpha)x}, respectively.

Remark 3.7.

It is in the case where Lα2>423L\alpha^{2}>\frac{4\sqrt{2}}{3} and L(2πα)2>423L(2\pi-\alpha)^{2}>\frac{4\sqrt{2}}{3} that one really sees the most dramatic effect of the assumption of disparate elastic constants present in our model. The relatively expensive cost of twist leads the global minimizer of EεE_{\varepsilon}, which of course is necessarily smooth, to rapidly change its phase, a process that can only be achieved with small energetic cost by having the modulus simultaneously plunge towards zero.

Remark 3.8.

We have not attempted to determine the optimal location of the jump location x0x_{0} for minimizers of EεE_{\varepsilon} in scenario (3.34). We suspect this might entail much higher order energetic considerations–perhaps even at an exponentially small order–but we are not sure.

Proof of Theorem 3.5.

When α=0\alpha=0 then clearly the global minimizer is uniquely given by u=ei2πNxu=e^{i2\pi Nx} since it has zero energy. Consider then the case α(0,2π).\alpha\in(0,2\pi). By selecting any point x0(0,1)x_{0}\in(0,1), and taking uu to be given by (3.34), we see that there is always a competitor with one jump having energy given simply by 223\frac{2\sqrt{2}}{3}. Any competitor jumping more than once has energy no lower than twice that value. On the other hand, minimization of E0E_{0} among competitors with J=J=\emptyset is standard, since criticality implies θ\theta^{\prime} is constant. Given the boundary conditions, this requires u=ei(2πM+α)xu=e^{i(2\pi M+\alpha)x} for some MM\in\mathbb{Z} to be determined. The energy of such a uu is L2(2π(MN)+α)2\frac{L}{2}(2\pi(M-N)+\alpha)^{2}. Since α(0,2π)\alpha\in(0,2\pi), the minimum over MM is L2(2π(NN)+α)2=L2α2\frac{L}{2}(2\pi(N-N)+\alpha)^{2}=\frac{L}{2}\alpha^{2} if α<2πα\alpha<2\pi-\alpha and L2(2π(N1N)+α)2=L2(2πα)2\frac{L}{2}(2\pi(N-1-N)+\alpha)^{2}=\frac{L}{2}(2\pi-\alpha)^{2} if 2πα<α2\pi-\alpha<\alpha. Comparing these two energies to that of the one-jump competitors in (3.34), the theorem follows. We note that if α=π\alpha=\pi in this regime, there are two global minimizers. ∎

Next we state a result on local minimizers of the Γ\Gamma-limit. These functions are the ε0\varepsilon\to 0 limit of the non-vanishing local minimizers captured in Theorem 2.3.

Theorem 3.9.

For any positive integer MM the function u=ei(2πM+α)xu=e^{i(2\pi M+\alpha)x} is an isolated L2L^{2}-local minimizer of E0E_{0}.

By invoking Theorem 4.1 of [10], one can conclude from Theorem 3.9 and Theorem 3.1 that there exist local minimizers of EεE_{\varepsilon} for ε\varepsilon small that converge to this isolated local minimizer of E0E_{0}. This provides for an alternative proof of existence for these local minimizers to the one given in Proposition 2.3. However, the approach in Theorem 2.3 yields much more detailed information on the structure of these functions via (2.2), (2.3) and (2.4).

Proof of Theorem 3.9.

We fix a positive integer MM and a number α[0,2π)\alpha\in[0,2\pi). We will consider the case M<NM<N. The case MNM\geq N is similar. Of course, in case M=NM=N and (3.31) holds, then in fact u=ei(2πN+α)xu=e^{i(2\pi N+\alpha)x} is the global minimizer, as was already addressed in Theorem 3.5. Let us denote θM:=2πMx+αx\theta_{M}:=2\pi Mx+\alpha x. In light of (3.29), our goal is to show that for some δ>0\delta>0, one has E0(θ)>E0(θM)E_{0}(\theta)>E_{0}(\theta_{M}) whenever θH1((0,1)J;S1)\theta\in H^{1}\big{(}(0,1)\setminus J;S^{1}\big{)} for some finite set JJ provided 0<θθML2(0,1)<δ0<\left\|\theta-\theta_{M}\right\|_{L^{2}(0,1)}<\delta.

We begin with the easiest case where J=J=\emptyset and where θ(0)=θM(0),θ(1)=θM(1)\theta(0)=\theta_{M}(0),\;\theta(1)=\theta_{M}(1). Writing v:=θθMv:=\theta-\theta_{M}, we calculate

E0(θ)E0(θM)=L201(θM+v2πN)2𝑑xL201(θM2πN)2𝑑x\displaystyle E_{0}(\theta)-E_{0}(\theta_{M})=\frac{L}{2}\int_{0}^{1}\big{(}\theta_{M}^{\prime}+v^{\prime}-2\pi N\big{)}^{2}\,dx-\frac{L}{2}\int_{0}^{1}\big{(}\theta_{M}^{\prime}-2\pi N\big{)}^{2}\,dx
=2πL(MN+α/2π)01v𝑑x+L201(v)2𝑑x=L201(v)2𝑑x>0,\displaystyle=2\pi L(M-N+\alpha/2\pi)\int_{0}^{1}v^{\prime}\,dx+\frac{L}{2}\int_{0}^{1}(v^{\prime})^{2}\,dx=\frac{L}{2}\int_{0}^{1}(v^{\prime})^{2}\,dx>0,

since in the case under consideration, v(0)=0=v(1)v(0)=0=v(1).

Now we turn to the general case where JJ\not=\emptyset. To this end, consider a competitor θH1(j=1(aj,bj))\theta\in H^{1}\big{(}\cup_{j=1}^{\ell}(a_{j},b_{j})\big{)} where then J=(0,1)j=1(aj,bj)J=(0,1)\setminus\cup_{j=1}^{\ell}(a_{j},b_{j}), along with perhaps x=0x=0 and/or x=1x=1, depending upon whether a competitor satisfies the boundary conditions. Thus, depending upon the boundary conditions of a competitor, we note that

0(J){1,,+1}.\mathcal{H}^{0}(J)\in\{\ell-1,\ell,\ell+1\}. (3.36)

Again we introduce v:=θθMv:=\theta-\theta_{M} and after a rearrangement of the indices, we suppose that for j=1,2,j=1,2\ldots,\ell^{\prime}, one has the condition

kj:=v(bj)v(aj)<26πN:=k0,k_{j}:=v(b_{j})-v(a_{j})<\frac{\sqrt{2}}{6\pi N}:=k_{0}, (3.37)

while for j=+1,,j=\ell^{\prime}+1,\ldots,\ell, the opposite inequality holds. We allow for the possibility that either =0\ell^{\prime}=0 or =\ell^{\prime}=\ell.

Then we again calculate the energy difference E0(θ)E0(θM)E_{0}(\theta)-E_{0}(\theta_{M}) by splitting up the sum as follows:

E0(θ)E0(θM)=2230(J)+2πL(MN+α/2π)j=1ajbjv𝑑x\displaystyle E_{0}(\theta)-E_{0}(\theta_{M})=\frac{2\sqrt{2}}{3}\mathcal{H}^{0}(J)+2\pi L(M-N+\alpha/2\pi)\sum_{j=1}^{\ell^{\prime}}\int_{a_{j}}^{b_{j}}v^{\prime}\,dx
+2πL(MN+α/2π)j=+1ajbjv𝑑x+L2j=1ajbj(v)2𝑑x\displaystyle+2\pi L(M-N+\alpha/2\pi)\sum_{j=\ell^{\prime}+1}^{\ell}\int_{a_{j}}^{b_{j}}v^{\prime}\,dx+\frac{L}{2}\sum_{j=1}^{\ell}\int_{a_{j}}^{b_{j}}(v^{\prime})^{2}\,dx
>2230(J)2πLNk0\displaystyle>\frac{2\sqrt{2}}{3}\mathcal{H}^{0}(J)-2\pi LNk_{0}\ell^{\prime}
2πLNj=+1kj+L2j=+1ajbj(v)2𝑑x\displaystyle-2\pi LN\sum_{j=\ell^{\prime}+1}^{\ell}k_{j}+\frac{L}{2}\sum_{j=\ell^{\prime}+1}^{\ell}\int_{a_{j}}^{b_{j}}(v^{\prime})^{2}\,dx
>j=+1(2πLNkj+L2ajbj(v)2𝑑x),\displaystyle>\sum_{j=\ell^{\prime}+1}^{\ell}\big{(}-2\pi LNk_{j}+\frac{L}{2}\int_{a_{j}}^{b_{j}}(v^{\prime})^{2}\,dx\big{)}, (3.38)

in light of (3.37) and (3.36).

If =\ell^{\prime}=\ell then the last sum is vacuous and the proof is complete. If not, then we now fix any j{+1,,}j\in\{\ell^{\prime}+1,\ldots,\ell\} for which the reverse inequality to (3.37) holds, and observe that

δj2:=ajbjv2𝑑x(aj,bj){|v|>kj/4}v2𝑑xkj216meas((aj,bj){|v|>kj/4}).\delta_{j}^{2}:=\int_{a_{j}}^{b_{j}}v^{2}\,dx\geq\int_{(a_{j},b_{j})\cap\{\left|{v}\right|>k_{j}/4\}}v^{2}\,dx\geq\frac{k_{j}^{2}}{16}{\rm meas}\,\big{(}(a_{j},b_{j})\cap\{\left|{v}\right|>k_{j}/4\}\big{)}. (3.39)

Also,

kj4<(aj,bj){|v|>kj/4}|v|dx\displaystyle\frac{k_{j}}{4}<\int_{(a_{j},b_{j})\cap\{\left|{v}\right|>k_{j}/4\}}\left|{v^{\prime}}\right|\,dx
meas((aj,bj){|v|>kj/4})1/2((aj,bj){|v|>kj/4}(v)2𝑑x)1/2.\displaystyle\leq{\rm meas}\,\big{(}(a_{j},b_{j})\cap\{\left|{v}\right|>k_{j}/4\}\big{)}^{1/2}\big{(}\int_{(a_{j},b_{j})\cap\{\left|{v}\right|>k_{j}/4\}}(v^{\prime})^{2}\,dx\big{)}^{1/2}.

Combining this with (3.39) yields the inequality

ajbj(v)2𝑑xkj4256δj2\int_{a_{j}}^{b_{j}}(v^{\prime})^{2}\,dx\geq\frac{k_{j}^{4}}{256\delta_{j}^{2}}

which we now substitute into (3.38) to conclude that

E0(θ)E0(θM)>j=+1(Lkj4512δj22πLNkj).E_{0}(\theta)-E_{0}(\theta_{M})>\sum_{j=\ell^{\prime}+1}^{\ell}\big{(}\frac{Lk_{j}^{4}}{512\delta_{j}^{2}}-2\pi LNk_{j}\big{)}. (3.40)

Choosing δ\delta (which we recall denotes (vL2(0,1)\left\|v\right\|_{L^{2}(0,1)}) such that

δ2<k031024πN,\delta^{2}<\frac{k_{0}^{3}}{1024\pi N},

and using that δjδ\delta_{j}\leq\delta while kjk0k_{j}\geq k_{0} for all jj, we obtain positivity of the right-hand side of (3.40).

4 An energy barrier leading to saddle points

The local minimizers provided by Theorem 2.3 can be viewed as the least energy critical points of EεE_{\varepsilon} within a given degree or winding number class given by the amount of twist. One might anticipate then that to pass continuously from one of these classes to another requires both the emergence of a zero in the order parameter and the expenditure of a certain amount of energy. What is more, one might expect the presence of saddle points in some sense interspersed between the distinct degree classes. That is the content of the two results in this section.

In the first theorem we demonstrate that the energy barrier between any two local minimizers uε,M1u_{\varepsilon,M_{1}} and uε,M2u_{\varepsilon,M_{2}} with M1M2M_{1}\neq M_{2} is at least 223\frac{2\sqrt{2}}{3} when ε\varepsilon is sufficiently small. To this end, given a Λ>0,\Lambda>0, we define the energy sublevel set

EεΛ:={u𝒜α:Eε(u)<Λ}.E_{\varepsilon}^{\Lambda}:=\left\{u\in\mathcal{A}_{\alpha}:E_{\varepsilon}(u)<\Lambda\right\}.

We have the following:

Theorem 4.1.

Let M1,M2M_{1},M_{2}\in\mathbb{N} be such that M1M2M_{1}\neq M_{2} and assume that uε,M1u_{\varepsilon,M_{1}} and uε,M2u_{\varepsilon,M_{2}} are local minimizers of EεE_{\varepsilon} as obtained in Theorem 2.3. Suppose that

γε:[0,1]𝒜αwithγε(0)=uε,M1andγε(1)=uε,M2\gamma^{\varepsilon}:[0,1]\to\mathcal{A}_{\alpha}\quad\mbox{with}\;\gamma^{\varepsilon}(0)=u_{\varepsilon,M_{1}}\;\mbox{and}\;\gamma^{\varepsilon}(1)=u_{\varepsilon,M_{2}} (4.1)

is a continuous path in 𝒜α\mathcal{A}_{\alpha} that connects uε,M1u_{\varepsilon,M_{1}} and uε,M2u_{\varepsilon,M_{2}}. Fix an h>0h>0 and set Λh:=2π2L(NM1α/2π)2+223h\Lambda_{h}:=2\pi^{2}L{(N-M_{1}-\alpha/2\pi)}^{2}+\frac{2\sqrt{2}}{3}-h. There exists an εh>0\varepsilon_{h}>0 such that the curve γε\gamma^{\varepsilon} leaves the set EεΛhE^{\Lambda_{h}}_{\varepsilon} whenever ε<εh.\varepsilon<\varepsilon_{h}.

Proof.

Fix any h(0,1)h\in(0,1) and any curve γε\gamma^{\varepsilon} satisfying (4.1). Denote

γε(t):=utε=ρtεeiθtε\gamma^{\varepsilon}(t):=u^{\varepsilon}_{t}=\rho^{\varepsilon}_{t}e^{i\theta^{\varepsilon}_{t}}

for every t[0,1]t\in[0,1]. The non-vanishing functions eiαxuε,M1e^{-i\alpha x}u_{\varepsilon,M_{1}} and eiαxuε,M2e^{-i\alpha x}u_{\varepsilon,M_{2}} have winding numbers M1M_{1} and M2M_{2} respectively on [0,1][0,1] and so utε(x)u^{\varepsilon}_{t}(x) has to vanish for some x(0,1)x\in(0,1) and t(0,1)t\in(0,1). Since γε\gamma^{\varepsilon} is continuous and utε()u^{\varepsilon}_{t}(\cdot) is a continuous function for every t[0,1]t\in[0,1], it follows that, given any δ(0,1/2),\delta\in(0,1/2), we can find tδε(0,1)t^{\varepsilon}_{\delta}\in(0,1) such that minx(0,1)ρtδεε(x)=δ\min_{x\in(0,1)}\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}}(x)=\delta and the winding number for eiαxutδεεe^{-i\alpha x}u^{\varepsilon}_{t^{\varepsilon}_{\delta}} is still equal to M1.M_{1}.

Now suppose by way of contradiction that γε([0,1])EεΛh\gamma^{\varepsilon}([0,1])\subset E_{\varepsilon}^{\Lambda_{h}}. We would like to estimate Eε(utδεε).E_{\varepsilon}(u^{\varepsilon}_{t^{\varepsilon}_{\delta}}). First, by minimizing Eε(ρtδεεeiθ)E_{\varepsilon}(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}}e^{i\theta}) over θ𝒯M1,α\theta\in\mathcal{T}_{M_{1},\alpha}, note that the same approach that led to (2.20) can be followed to show that there exists a θ¯ε𝒯M1,α{\bar{\theta}}_{\varepsilon}\in{\mathcal{T}}_{M_{1},\alpha} such that

θ¯ε=2πLM1+Lα+2πLN((ρtδεε)21)L(ρtδεε)4+ε(ρtδεε)2+O(ε){\bar{\theta}}_{\varepsilon}^{\prime}=\frac{2\pi LM_{1}+L\alpha+2\pi LN((\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{2}-1)}{L(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{4}+\varepsilon{(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{2}}}+O(\sqrt{\varepsilon}) (4.2)

on (0,1)(0,1), and necessarily

Eε(ρtδεeiθ¯ε)Eε(utδε).E_{\varepsilon}\left(\rho^{\varepsilon}_{t_{\delta}}e^{i\bar{\theta}_{\varepsilon}}\right)\leq E_{\varepsilon}\left(u^{\varepsilon}_{t_{\delta}}\right). (4.3)

Using the standard Modica-Mortola arguments, we now have

01ε2((ρtδεε))2+14ε((ρtδεε)21)2dxc(δ),\int_{0}^{1}\frac{\varepsilon}{2}((\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{\prime})^{2}+\frac{1}{4\varepsilon}\big{(}(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{2}-1\big{)}^{2}\,dx\geq c(\delta),

where limδ0c(δ)=223.\lim_{\delta\to 0}c(\delta)=\frac{2\sqrt{2}}{3}. Further, we can appeal to (4.2)-(4.3) and the assumption that γε(tδε)EεΛh\gamma^{\varepsilon}(t^{\varepsilon}_{\delta})\in E_{\varepsilon}^{\Lambda_{h}} to show that

L201(2πN(ρtδεε)2θ¯ε)2𝑑x=L201(2πN2πLM1+Lα+2πLN((ρtδεε)21)L(ρtδεε)2+ε)2𝑑x+O(ε)=2π2L01(NM1α/2π+(ε/L)N(ρtδεε)2+ε/L)2𝑑x+O(ε)=2π2L(NM1α/2π)2+O(ε).\frac{L}{2}\int_{0}^{1}\left(2\pi N-(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{2}{\bar{\theta}}^{\prime}_{\varepsilon}\right)^{2}\,dx=\frac{L}{2}\int_{0}^{1}\left(2\pi N-\frac{2\pi LM_{1}+L\alpha+2\pi LN\big{(}{(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}}})^{2}-1\big{)}}{L(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{2}+\varepsilon}\right)^{2}\,dx+O(\sqrt{\varepsilon})\\ =2\pi^{2}L\int_{0}^{1}\left(\frac{N-M_{1}-\alpha/2\pi+(\varepsilon/L)N}{(\rho^{\varepsilon}_{t^{\varepsilon}_{\delta}})^{2}+\varepsilon/L}\right)^{2}\,dx+O(\sqrt{\varepsilon})\\ =2\pi^{2}L{(N-M_{1}-\alpha/2\pi)}^{2}+O(\sqrt{\varepsilon}). (4.4)

It then follows from (4.3) that

Eε(utδεε)2π2L(NM1α/2π)2+c(δ)+O(ε).E_{\varepsilon}(u^{\varepsilon}_{t^{\varepsilon}_{\delta}})\geq 2\pi^{2}L{(N-M_{1}-\alpha/2\pi)}^{2}+c(\delta)+O(\sqrt{\varepsilon}).

It is clear, however, that one can select a positive δ\delta sufficiently small, and then an εh>0\varepsilon_{h}>0 such that the last expression exceeds Λh\Lambda_{h} whenever ε<εh\varepsilon<\varepsilon_{h}. ∎

The energy threshold provided by Theorem 4.1 leads to a straight-forward application of the Mountain Pass Theorem to establish saddle points for EεE_{\varepsilon}.

Theorem 4.2.

For every positive integer MM and α[0,2π)\alpha\in[0,2\pi) there exists a critical point vεv_{\varepsilon} of EεE_{\varepsilon} within the class 𝒜α\mathcal{A}_{\alpha}. Furthermore, the corresponding critical value Eε(vε)E_{\varepsilon}(v_{\varepsilon}) satisfies the asymptotic condition

Eε(vε)2π2L(NMα/2π)2+223asε0.E_{\varepsilon}(v_{\varepsilon})\to 2\pi^{2}L{(N-M-\alpha/2\pi)}^{2}+\frac{2\sqrt{2}}{3}\quad\mbox{as}\;\varepsilon\to 0. (4.5)
Proof.

First, we note that the arguments in Theorem 4.1 can easily be adapted with the same energy threshold to a curve that connects the states UM:=ei(2πM+α)xU_{M}:=e^{i(2\pi M+\alpha)x} and UM1:=ei(2πM1+α)xU_{M_{1}}:=e^{i(2\pi M_{1}+\alpha)x} for any two positive integers MM and M1M_{1}. Fixing ε>0\varepsilon>0 one defines the potential critical value cεc_{\varepsilon} via

cε:=infγΓεmaxt[0,1]Eε(γ(t)),c_{\varepsilon}:=\inf_{\gamma\in\Gamma_{\varepsilon}}\;\max_{t\in[0,1]}\;E_{\varepsilon}(\gamma(t)),

where Γε\Gamma_{\varepsilon} is the set of continuous curves γ\gamma such that

γ:[0,1]𝒜αwithγ(0)=UMandγ(1)=UM+1.\gamma:[0,1]\to\mathcal{A}_{\alpha}\quad\mbox{with}\;\gamma(0)=U_{M}\;\mbox{and}\;\gamma(1)=U_{M+1}. (4.6)

Beginning with the case M<NM<N we have that

Eε(UM)2π2L(NMα/2π)2E_{\varepsilon}(U_{M})\sim 2\pi^{2}L{(N-M-\alpha/2\pi)}^{2}

while

Eε(UM+1)2π2L(NM1α/2π)2,E_{\varepsilon}(U_{M+1})\sim 2\pi^{2}L{(N-M-1-\alpha/2\pi)}^{2},

so that, in particular, Eε(UM+1)<Eε(UM)E_{\varepsilon}(U_{M+1})<E_{\varepsilon}(U_{M}). Then the implication of Theorem 4.1 is that EεE_{\varepsilon} exhibits the requisite mountain pass structure since for any h>0h>0 one has

maxt[0,1]Eε(γ(t))2π2L(NMα/2π)2+223h>Eε(UM)>Eε(UM+1)\max_{t\in[0,1]}\;E_{\varepsilon}(\gamma(t))\geq 2\pi^{2}L{(N-M-\alpha/2\pi)}^{2}+\frac{2\sqrt{2}}{3}-h>E_{\varepsilon}(U_{M})>E_{\varepsilon}(U_{M+1}) (4.7)

for any γΓε\gamma\in\Gamma_{\varepsilon}, provided ε\varepsilon is sufficiently small.

Subtracting off the boundary conditions by writing any competitor u𝒜αu\in\mathcal{A}_{\alpha} as u=u~+(x)u=\tilde{u}+\ell(x) where (x):=1+x(eiα1),\ell(x):=1+x\big{(}e^{i\alpha}-1\big{)}, we can work in the space H01((0,1))H^{1}_{0}\big{(}(0,1)\big{)}. It remains to verify the Palais-Smale condition. Under assumptions

Eε(u~k+)<C0andδEε(u~k+)0ask0,E_{\varepsilon}(\tilde{u}_{k}+\ell)<C_{0}\quad\mbox{and}\quad\left\|\delta E_{\varepsilon}(\tilde{u}_{k}+\ell)\right\|\to 0\quad\mbox{as}\;k\to 0,

for {u~k}H01((0,1))\{\tilde{u}_{k}\}\subset H^{1}_{0}\big{(}(0,1)\big{)}, it immediately follows from the uniform energy bound that after passing to a subsequence (with notation suppressed), one has

u~ku~ε,Mweakly inH1andu~ku~ε,Muniformlyask,\tilde{u}_{k}\rightharpoonup\tilde{u}_{\varepsilon,M}\;\mbox{weakly in}\;H^{1}\quad\mbox{and}\quad\tilde{u}_{k}\to\tilde{u}_{\varepsilon,M}\;\mbox{uniformly}\quad\mbox{as}\;k\to\infty, (4.8)

for some u~ε,MH01\tilde{u}_{\varepsilon,M}\in H^{1}_{0}. Then one writes EεE_{\varepsilon} as the sum of the Allen-Cahn energy and the twist energy I(u~):=L201𝒯(u~+)𝑑xI(\tilde{u}):=\frac{L}{2}\int_{0}^{1}\mathcal{T}(\tilde{u}+\ell)\,dx and one follows the standard proof used to verify that the Allen-Cahn functional satisfies Palais-Smale (see e.g. [9], Prop. 3.3). The key step in upgrading the weak H1H^{1} convergence to strong convergence is writing out the difference δEε(u~k+;u~k)δEε(u~k+;u~ε,M)\delta E_{\varepsilon}(\tilde{u}_{k}+\ell;\tilde{u}_{k})-\delta E_{\varepsilon}(\tilde{u}_{k}+\ell;\tilde{u}_{\varepsilon,M}), and in light of the convergences (4.8), the extra twist terms in this difference pose no additional trouble. We conclude from the Mountain Pass Theorem that a critical point vε,M:=u~ε,M+v_{\varepsilon,M}:=\tilde{u}_{\varepsilon,M}+\ell exists with Eε(vε,M)=cεE_{\varepsilon}(v_{\varepsilon,M})=c_{\varepsilon}.

Now we turn to the proof of condition (4.5). Again, we know from Theorem 4.1 that for any h>0h>0, one has the inequality (4.7) for ε\varepsilon small enough, so that

lim infcε2π2L(NMα/2π)2+223\liminf c_{\varepsilon}\geq 2\pi^{2}L{(N-M-\alpha/2\pi)}^{2}+\frac{2\sqrt{2}}{3} (4.9)

On the other hand, we can build a continuous path γε:[0,1]𝒜α\gamma^{\varepsilon}:[0,1]\to\mathcal{A}_{\alpha} as follows:
1) Writing UM=eiθMU_{M}=e^{i\theta_{M}} as tt varies between 0 and say 1/31/3, the modulus gradually depresses towards 0 in a small interval of xx-values about x=1/2x=1/2 via the standard Modica-Mortola construction, so that γε(1/3)0\gamma^{\varepsilon}(1/3)\equiv 0 for say 1/2ε2x1/2+ε21/2-\varepsilon^{2}\leq x\leq 1/2+\varepsilon^{2}. For this interval of tt-values one leaves the phase θM\theta_{M} unchanged. As we have previously noted, such a procedure can be executed with

Eε(γε(t))2π2L(NMα/2π)2+223+O(ε)fort[0,1/3).E_{\varepsilon}(\gamma^{\varepsilon}(t))\leq 2\pi^{2}L{(N-M-\alpha/2\pi)}^{2}+\frac{2\sqrt{2}}{3}+O(\varepsilon)\;\mbox{for}\;t\in[0,1/3).

Of course along the subinterval where the modulus vanishes, the value of the phase is irrelevant but we find it convenient in this exposition to define the phase throughout the whole interval for each function γε(t)\gamma^{\varepsilon}(t).
2) At t=1/3t=1/3 we introduce a removable discontinuity in the phase θM\theta_{M} at x=1/2x=1/2 where the modulus vanishes. Then, as tt increases from t=1/3t=1/3 to t=2/3t=2/3, one takes the phase to gradually converge to θM+1\theta_{M+1} and θM+12π\theta_{M+1}-2\pi on [0,1/2)[0,1/2) and (1/2,1](1/2,1], respectively, while leaving the modulus unchanged. Since M<NM<N, the O(1)O(1) energy contribution of the twist will decrease under this process. As tt approaches 2/32/3, we converge to UM+1U_{M+1} except for the small interval about x=1/2x=1/2 where the modulus is depressed.
3) In the time interval t[2/3,1]t\in[2/3,1] one smoothly raises the modulus back up to 11 on 1/2ε2x1/2+ε21/2-\varepsilon^{2}\leq x\leq 1/2+\varepsilon^{2} so that at t=1t=1 one has γε(1)=UM+1\gamma^{\varepsilon}(1)=U_{M+1}. Again, this process decreases energy so that throughout the interval 0t10\leq t\leq 1 one maintains the estimate

Eε(γε(t))2π2L(NMα/2π)2+223+O(ε).E_{\varepsilon}(\gamma^{\varepsilon}(t))\leq 2\pi^{2}L{(N-M-\alpha/2\pi)}^{2}+\frac{2\sqrt{2}}{3}+O(\varepsilon).

Hence, we conclude that

lim supcεlim supEε(γε)2π2L(NMα/2π)2+223\limsup{c_{\varepsilon}}\leq\limsup{E_{\varepsilon}(\gamma^{\varepsilon})}\leq 2\pi^{2}L{(N-M-\alpha/2\pi)}^{2}+\frac{2\sqrt{2}}{3}

and together with (4.9) we arrive at (4.5). ∎

5 The case of unbounded twist

Finally, we consider the situation of an energy that encourages more and more twist in the ε0\varepsilon\to 0 limit. To this end, we replace NN in (1.2) by Nε:=1/εβN_{\varepsilon}:=1/\varepsilon^{\beta} where β\beta is a positive number chosen less than 1/21/2 in order to retain an energy bound that is uniform in ε\varepsilon. Thus, we study global and local minimizers of an energy E~ε\tilde{E}_{\varepsilon} given by

E~ε(u)=01ε2|u|2+14ε(|u|21)2+L2(u1u2u2u12πεβ)2dx,\tilde{E}_{\varepsilon}(u)=\int_{0}^{1}\frac{\varepsilon}{2}\left|{u^{\prime}}\right|^{2}+\frac{1}{4\varepsilon}(\left|{u}\right|^{2}-1)^{2}+\frac{L}{2}(u_{1}\,u_{2}^{\prime}-u_{2}\,u_{1}^{\prime}-2\pi\varepsilon^{-\beta})^{2}\,dx, (5.1)

again subject to the boundary conditions u(0)=1,u(1)=eiαu(0)=1,\;u(1)=e^{i\alpha} for some α[0,2π)\alpha\in[0,2\pi).

Of course existence of global minimizers for each ε>0\varepsilon>0 follows as in Theorem 2.1. One also can establish a version of the local minimizer result Theorem 2.3:

Theorem 5.1.

Fix any positive integer mm and any α[0,2π)\alpha\in[0,2\pi). Then there exists an ε0>0\varepsilon_{0}>0 such that for all ε<ε0\varepsilon<\varepsilon_{0} there exist non-vanishing local minimizers uε,±=ρε,±eiθε,±u_{\varepsilon,\pm}=\rho_{\varepsilon,\pm}e^{i\theta_{\varepsilon,\pm}} of E~ε\tilde{E}_{\varepsilon} within the class 𝒜α\mathcal{A}_{\alpha} such that

lim supρε,±1L(0,1)ε<asε0\displaystyle\limsup\frac{\left\|{\rho_{\varepsilon,\pm}}-1\right\|_{L^{\infty}(0,1)}}{\varepsilon}<\infty\;\mbox{as}\;\varepsilon\to 0 (5.2)
and (5.3)
θε,±2π(εβ±m)+αasε0uniformly inx[0,1].\displaystyle\theta_{\varepsilon,\pm}^{\prime}\to 2\pi\left(\left\lfloor\varepsilon^{-\beta}\right\rfloor\pm m\right)+\alpha\;\mbox{as}\;\varepsilon\to 0\;\mbox{uniformly in}\;x\in[0,1]. (5.4)
Proof.

The proof follows along similar lines as the proof of Theorem 2.3. First define Mε±=εβ±mM^{\pm}_{\varepsilon}=\left\lfloor\varepsilon^{-\beta}\right\rfloor\pm m. Then one writes competitors for constrained minimization of E~ε=E~ε(ρ,θ)\tilde{E}_{\varepsilon}=\tilde{E}_{\varepsilon}(\rho,\theta) in polar form (ρ,θ)(\rho,\theta) where ρ\rho satisfies (2.6) and θ(0)=0,θ(1)=2πMε±+α\theta(0)=0,\;\theta(1)=2\pi M^{\pm}_{\varepsilon}+\alpha. The requirement β<1/2\beta<1/2 assures that a version of the uniform energy bound (2.9) still holds. Similarly, a uniform bound on the constant of integration CεC_{\varepsilon} is achievable as in (2.19), with the bound now depending on mm. The rest of the argument is unchanged.∎

Next we consider the asymptotic behavior as ε0\varepsilon\to 0 of E~ε\tilde{E}_{\varepsilon}. Due to the fact that εβ\varepsilon^{-\beta}\to\infty as ε0\varepsilon\to 0, we expect that the elements of an energy bounded sequence will oscillate more and more rapidly as ε0\varepsilon\to 0.

Theorem 5.2.

Suppose that for some 0<β<1/20<\beta<1/2, {uε}𝒜α\{u_{\varepsilon}\}\subset\mathcal{A}_{\alpha} satisfies the uniform energy bound

E~ε(uε)C0<.\tilde{E}_{\varepsilon}(u_{\varepsilon})\leq C_{0}<\infty. (5.5)

Then |uε|21\left|{u_{\varepsilon}}\right|^{2}\to 1 in L2(0,1)L^{2}(0,1) and there exists a finite set J(0,1)J^{\prime}\subset(0,1) and a subsequence {uε}\{u_{\varepsilon_{\ell}}\} such that for every compact set K(0,1)JK\subset\subset(0,1)\setminus J^{\prime}, there exists an ε0(K)>0\varepsilon_{0}(K)>0 such that for every ε<ε0\varepsilon_{\ell}<\varepsilon_{0}, one has |uε|>0|u_{\varepsilon_{\ell}}|>0 on KK and there is a lifting whereby uε=ρεe2πivε/εβu_{\varepsilon_{\ell}}=\rho_{\varepsilon_{\ell}}e^{2\pi iv_{\varepsilon_{\ell}}/\varepsilon_{\ell}^{\beta}}, with

vεx strongly in Hloc1((0,1)J).v_{\varepsilon_{\ell}}\to x\textit{ strongly in }H^{1}_{loc}((0,1)\setminus J^{\prime}). (5.6)

In addition, we have

uε0 weakly in L2((0,1);),u_{\varepsilon}\rightharpoonup 0\textit{ weakly in }L^{2}((0,1);\mathbb{C}), (5.7)

so that the entire sequence converges weakly to 0.

Proof of Theorem 5.2.

By the same argument as the one leading up to (3.10), we can identify finite unions of open intervals BεB_{\varepsilon} such that on (0,1)Bε(0,1)\setminus B_{\varepsilon}, ρε2q\rho_{\varepsilon}\geq 2^{-q}. Also, by restricting to a subsequence {ε}\{\varepsilon_{\ell}\}, we can assume that the sets BεB_{\varepsilon_{\ell}} collapse to a finite set of points JJ^{\prime}. We may therefore define liftings θε\theta_{\varepsilon_{\ell}} such that on each of the finitely many intervals comprising (0,1)Bε(0,1)\setminus B_{\varepsilon_{\ell}}, the value of θε\theta_{\varepsilon_{\ell}} at the left endpoint of an interval is greater than the value of θε\theta_{\varepsilon_{\ell}} at the right endpoint of the previous interval, with a difference of no more than 2π2\pi. Also, we can without loss of generality suppose that 0 is in the domain of θε\theta_{\varepsilon_{\ell}} and set θε(0)=0\theta_{\varepsilon_{\ell}}(0)=0. If we define

vε:=εβθε2π,v_{\varepsilon_{\ell}}:=\frac{\varepsilon_{\ell}^{\beta}\theta_{\varepsilon_{\ell}}}{2\pi}, (5.8)

then we may rewrite the twist term in terms of vεv_{\varepsilon_{\ell}} and use the uniform energy bound to conclude that

12(0,1)BεLε2β(ρε2vε1)2C0.\frac{1}{2}\int_{(0,1)\setminus B_{\varepsilon_{\ell}}}\frac{L}{\varepsilon_{\ell}^{2\beta}}\left(\rho_{\varepsilon_{\ell}}^{2}v_{\varepsilon_{\ell}}^{\prime}-1\right)^{2}\leq C_{0}. (5.9)

Furthermore, due to the choice of θε\theta_{\varepsilon_{\ell}} on each subinterval of BεB_{\varepsilon_{\ell}}, we see that

the value of vεv_{\varepsilon_{\ell}} jumps by no more than εβ\varepsilon_{\ell}^{\beta} (5.10)

from the right endpoint of one subinterval to the left endpoint of the subsequent one. After passing to a further subsequence (with notation suppressed), we conclude from (5.9) that for any K1[0,1]JK_{1}\subset\subset[0,1]\setminus J^{\prime},

vε1 in L2(K1).v_{\varepsilon_{\ell}}^{\prime}\to 1\textup{ in }L^{2}(K_{1}). (5.11)

From (5.10), (5.11), and the condition vε(0)=0v_{\varepsilon_{\ell}}(0)=0, we deduce that

vεx in L(K1).v_{\varepsilon_{\ell}}\to x\textup{ in $L^{\infty}(K_{1})$}. (5.12)

As in (3.17), we may repeat this procedure on a nested sequence of compact sets to arrive at a subsequence (still denoted by vεv_{\varepsilon_{\ell}}) such that

vεx in Hloc1((0,1)J).v_{\varepsilon_{\ell}}\to x\textup{ in }H^{1}_{loc}((0,1)\setminus J^{\prime}).

To prove (5.7), we must demonstrate that for any wL2((0,1);)w\in L^{2}((0,1);\mathbb{C}),

01uεw¯𝑑x0,\int_{0}^{1}u_{\varepsilon}\overline{w}\,dx\to 0, (5.13)

where the bar denotes complex conjugation. Let us first obtain a subsequence uεu_{\varepsilon_{\ell}} satisfying (5.11) and (5.12) such that lim|uεw¯|\lim_{\ell\to\infty}\left|\int u_{\varepsilon_{\ell}}\overline{w}\right| achieves the limit superior. By Egorov’s theorem, after restricting to a subsequence such that vε1v_{\varepsilon_{\ell}}^{\prime}\to 1 almost everywhere, we can assume without loss of generality that vε1v_{\varepsilon_{\ell}}^{\prime}\to 1 almost uniformly on (0,1)J(0,1)\setminus J^{\prime}. Also, if we approximate ww by smooth functions, it is enough to show that for any K(0,1)JK\subset\subset(0,1)\setminus J^{\prime} on which vε1v_{\varepsilon_{\ell}}^{\prime}\to 1 uniformly,

01uεφ¯𝑑x=01ρεe2πivε/εβφ¯𝑑x0\int_{0}^{1}u_{\varepsilon_{\ell}}\overline{\varphi}\,dx=\int_{0}^{1}\rho_{\varepsilon_{\ell}}e^{2\pi iv_{\varepsilon_{\ell}}/\varepsilon^{\beta}_{\ell}}\overline{\varphi}\,dx\to 0 (5.14)

if φCc(K;)\varphi\in C_{c}^{\infty}(K;\mathbb{C}). Since ρε1\rho_{\varepsilon_{\ell}}\to 1 in L2L^{2}, (5.14) would follow from the condition

01e2πivε/εβφ¯𝑑x0,\int_{0}^{1}e^{2\pi iv_{\varepsilon_{\ell}}/\varepsilon^{\beta}_{\ell}}\overline{\varphi}\,dx\to 0,

which we now prove. Because vε1v_{\varepsilon_{\ell}}^{\prime}\to 1 uniformly on sptφ\mathrm{spt}\,\varphi, vεv_{\varepsilon_{\ell}} is a diffeomorphism of sptφ\mathrm{spt}\,\varphi for large \ell. Thus for \ell large enough, we can add and subtract e2πix/εβφ¯\int e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\overline{\varphi} and then change variables to get

|sptφe2πivε/εβφ¯|\displaystyle\left|\int_{\mathrm{spt}\,\varphi}e^{2\pi iv_{\varepsilon_{\ell}}/\varepsilon^{\beta}_{\ell}}\overline{\varphi}\right| |sptφe2πix/εβφ¯|+|sptφe2πix/εβφ¯sptφe2πivε/εβφ¯|\displaystyle\leq\left|\int_{\mathrm{spt}\,\varphi}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\overline{\varphi}\right|+\left|\int_{\mathrm{spt}\,\varphi}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\overline{\varphi}-\int_{\mathrm{spt}\,\varphi}e^{2\pi iv_{\varepsilon_{\ell}}/\varepsilon^{\beta}_{\ell}}\overline{\varphi}\right|
=|sptφe2πix/εβφ¯|+|sptφe2πix/εβφ¯vε(sptφ)e2πix/εβφ¯(vε1)|vε(vε1)||.\displaystyle=\left|\int_{\mathrm{spt}\,\varphi}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\overline{\varphi}\right|+\left|\int_{\mathrm{spt}\,\varphi}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\overline{\varphi}-\int_{v_{\varepsilon_{\ell}}(\mathrm{spt}\,\varphi)}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\frac{\overline{\varphi}(v_{\varepsilon_{\ell}}^{-1})}{|v_{\varepsilon_{\ell}}^{\prime}(v_{\varepsilon_{\ell}}^{-1})|}\right|.

The first term goes to zero as \ell\to\infty since e2πix/εβ0e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\rightharpoonup 0. In the second term we can add and subtract vε(sptφ)e2πix/εβφ¯(vε1)\int_{v_{\varepsilon_{\ell}}(\mathrm{spt}\,\varphi)}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\overline{\varphi}(v_{\varepsilon_{\ell}}^{-1}) and then change variables back, yielding

|sptφe2πix/εβφ¯\displaystyle\left|\int_{\mathrm{spt}\,\varphi}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\overline{\varphi}-\right. vε(sptφ)e2πix/εβφ¯(vε1)|vε(vε1)||\displaystyle\left.\int_{v_{\varepsilon_{\ell}}(\mathrm{spt}\,\varphi)}e^{2\pi ix/\varepsilon^{\beta}_{\ell}}\frac{\overline{\varphi}(v_{\varepsilon_{\ell}}^{-1})}{|v_{\varepsilon_{\ell}}^{\prime}(v_{\varepsilon_{\ell}}^{-1})|}\right|
vε(sptφ)sptφ|φφ(vε1)|+vε(sptφ)|φ(vε1)||vε(vε1)|||vε(vε1)|1|\displaystyle\leq\int_{v_{\varepsilon_{\ell}}(\mathrm{spt}\,\varphi)\cup\mathrm{spt}\,\varphi}\left|\varphi-\varphi(v_{\varepsilon_{\ell}}^{-1})\right|+\int_{v_{\varepsilon_{\ell}}(\mathrm{spt}\,\varphi)}\frac{|\varphi(v_{\varepsilon_{\ell}}^{-1})|}{|v_{\varepsilon_{\ell}}^{\prime}(v_{\varepsilon_{\ell}}^{-1})|}\left||v_{\varepsilon_{\ell}}^{\prime}(v_{\varepsilon_{\ell}}^{-1})|-1\right|
φLvε(sptφ)sptφ|xvε1|+φLsptφ|vε1|,\displaystyle\leq\|\varphi^{\prime}\|_{L^{\infty}}\int_{v_{\varepsilon_{\ell}}(\mathrm{spt}\,\varphi)\cup\mathrm{spt}\,\varphi}|x-v_{\varepsilon_{\ell}}^{-1}|+\|\varphi\|_{L^{\infty}}\int_{\mathrm{spt}\,\varphi}|v_{\varepsilon_{\ell}}^{\prime}-1|,

which approaches zero by (5.11) and (5.12). ∎

We would also like to describe the asymptotic behavior of minimizers in this regime by identifying a limiting problem. As demonstrated in the previous theorem, no meaningful limit can be extracted from simply looking at the sequence {uε}\{u_{\varepsilon}\}. Instead, we examine the “microscale” behavior of uεu_{\varepsilon} by eliminating the excess twist in the limit ε0\varepsilon\to 0, in the sense that we obtain a limiting asymptotic problem for the rescaled functions

w(x):=u(x)e2πiεβx.w(x):=u(x)e^{-2\pi i\lfloor\varepsilon^{-\beta}\rfloor x}.

Here εβ\lfloor\varepsilon^{-\beta}\rfloor denotes the integer part of εβ\varepsilon^{-\beta}.

In terms of ww, the energy E~ε(u)\tilde{E}_{\varepsilon}(u) is given by

E~ε(u)=Fε(w)\displaystyle\tilde{E}_{\varepsilon}(u)=F_{\varepsilon}(w) :=01ε2|(we2πiεβx)|2+14ε(|w|21)2\displaystyle:=\int_{0}^{1}\frac{\varepsilon}{2}\left|{(we^{2\pi i\lfloor\varepsilon^{-\beta}\rfloor x})^{\prime}}\right|^{2}+\frac{1}{4\varepsilon}(\left|{w}\right|^{2}-1)^{2}
+L2(w1w2w2w1+|w|22πεβ2πεβ)2dx.\displaystyle\qquad+\frac{L}{2}(w_{1}\,w_{2}^{\prime}-w_{2}\,w_{1}^{\prime}+|w|^{2}2\pi\lfloor\varepsilon^{-\beta}\rfloor-2\pi\varepsilon^{-\beta})^{2}\,dx.

The boundary conditions imposed on competitors for FεF_{\varepsilon} are the same as those for E~ε\tilde{E}_{\varepsilon}. The asymptotic behavior of minimizers of E~ε\tilde{E}_{\varepsilon} can therefore be completely understood in terms of FεF_{\varepsilon}, so we pursue an asymptotic limit for FεF_{\varepsilon}. Let us define the limiting functional as in Section 3, with slightly altered notation to emphasize the dependence on preferred twist:

E0,A(w):={L201(w1w2w2w12πA)2𝑑x+2230(J)ifwH1((0,1)J;S1)+ otherwise.\displaystyle E_{0,A}(w):=\left\{\begin{array}[]{cc}\displaystyle\frac{L}{2}\int_{0}^{1}\displaystyle(w_{1}\,w_{2}^{\prime}-w_{2}\,w_{1}^{\prime}-2\pi A)^{2}\,dx+\frac{2\sqrt{2}}{3}\mathcal{H}^{0}(J)&\mbox{if}\;w\in H^{1}((0,1)\setminus J;S^{1})\\ \\ +\infty&\mbox{ otherwise}.\end{array}\right. (5.18)

We recall that 0 and/or 11 belongs to JJ depending on whether or not the traces of uu satisfy the desired boundary conditions inherited from EεE_{\varepsilon}; that is, we include x=0x=0 in JJ only if u(0+)1u(0^{+})\not=1 and we include x=1x=1 in JJ only if u(1)eiαu(1^{-})\not=e^{i\alpha}.

Theorem 5.3.

Let 0<β<1/20<\beta<1/2 and suppose that for a subsequence {ε}0\{\varepsilon_{\ell}\}\to 0 and some A[0,1]A\in[0,1] we have

εβεβA.\varepsilon_{\ell}^{-\beta}-\lfloor\varepsilon_{\ell}^{-\beta}\rfloor\to A.

Then {Fε}\{F_{\varepsilon_{\ell}}\} Γ\Gamma-converges to E0,AE_{0,A} in L2((0,1);2)L^{2}\left((0,1);\mathbb{R}^{2}\right).

We also have the compactness result

Theorem 5.4.

If {uε}ε>0\{u_{\varepsilon}\}_{\varepsilon>0} satisfies

E~ε(uε)=Fε(wε)C0<,\tilde{E}_{\varepsilon}(u_{\varepsilon})=F_{\varepsilon}(w_{\varepsilon})\leq C_{0}<\infty, (5.19)

and

εβεβA\varepsilon_{\ell}^{-\beta}-\lfloor\varepsilon_{\ell}^{-\beta}\rfloor\to A (5.20)

for some 0<β<1/20<\beta<1/2, then there exists a function wH1((0,1)J;S1)w\in H^{1}((0,1)\setminus J^{\prime};S^{1}) where JJ^{\prime} is a finite, perhaps empty, set of points in (0,1)(0,1) such that along a subsequence ε0\varepsilon_{\ell}\to 0 one has

uεe2πiεβx=wεwinL2((0,1);).u_{\varepsilon_{\ell}}e^{-2\pi i\lfloor\varepsilon_{\ell}^{-\beta}\rfloor x}=w_{\varepsilon_{\ell}}\to w\;\mbox{in}\;L^{2}\big{(}(0,1);\mathbb{C}\big{)}. (5.21)

Furthermore, writing w(x)=eiθ(x)w(x)=e^{i\theta(x)} for θH1((0,1)J)\theta\in H^{1}((0,1)\setminus J^{\prime}), we have that for every compact set K(0,1)JK\subset\subset(0,1)\setminus J^{\prime}, there exists an ε0(K)>0\varepsilon_{0}(K)>0 such that for every ε<ε0\varepsilon_{\ell}<\varepsilon_{0} one has |uε|=|wε|>0\left|{u_{\varepsilon_{\ell}}}\right|=\left|{w_{\varepsilon_{\ell}}}\right|>0 on KK and there is a lifting whereby uε(x)e2πiεβx=wε(x)=ρε(x)eiθε(x)u_{\varepsilon_{\ell}}(x)e^{-2\pi i\lfloor\varepsilon_{\ell}^{-\beta}\rfloor x}=w_{\varepsilon_{\ell}}(x)=\rho_{\varepsilon_{\ell}}(x)e^{i\theta_{\varepsilon_{\ell}}(x)} on KK, with

θεθweakly inHloc1((0,1)J).\theta_{\varepsilon_{\ell}}\rightharpoonup\theta\;\mbox{weakly in}\;H^{1}_{loc}\big{(}(0,1)\setminus J^{\prime}\big{)}. (5.22)
Proof of Theorem 5.4.

The proof is based on the proof of Theorem 3.2. First, we estimate that

01ε2|(wεe2πiεβx)|2𝑑x\displaystyle\int_{0}^{1}\frac{\varepsilon}{2}\left|{(w_{\varepsilon}e^{2\pi i\lfloor\varepsilon^{-\beta}\rfloor x})^{\prime}}\right|^{2}\,dx =01ε2|wε+2πiwεβ|2𝑑x\displaystyle=\int_{0}^{1}\frac{\varepsilon}{2}\left|w_{\varepsilon}^{\prime}+2\pi iw\lfloor\varepsilon^{-\beta}\rfloor\right|^{2}\,dx
=01ε2|wε|2𝑑x+O(ε1/2β)dx\displaystyle=\int_{0}^{1}\frac{\varepsilon}{2}\left|w_{\varepsilon}^{\prime}\right|^{2}\,dx+O(\varepsilon^{1/2-\beta})\,dx (5.23)

for an energy bounded sequence {wε}\{w_{\varepsilon}\}. Therefore,

Fε(wε)=01ε2|wε|2\displaystyle F_{\varepsilon}(w_{\varepsilon})=\int_{0}^{1}\frac{\varepsilon}{2}\left|{w_{\varepsilon}^{\prime}}\right|^{2} +14ε(|wε|21)2\displaystyle+\frac{1}{4\varepsilon}(\left|{w_{\varepsilon}}\right|^{2}-1)^{2}
+L2(𝒯(wε)+|wε|22πεβ2πεβ)2dx+O(ε1/2β).\displaystyle+\frac{L}{2}(\mathcal{T}(w_{\varepsilon})+|w_{\varepsilon}|^{2}2\pi\lfloor\varepsilon^{-\beta}\rfloor-2\pi\varepsilon^{-\beta})^{2}\,dx+O(\varepsilon^{1/2-\beta}). (5.24)

The rest of the proof follows almost exactly as in Theorem 3.2. Indeed, the only difference between EεE_{\varepsilon} in that theorem and the right hand side of (5.24) here is the preferred twist 2πN2\pi N versus |wε|22πεβ2πεβ|w_{\varepsilon}|^{2}2\pi\lfloor\varepsilon^{-\beta}\rfloor-2\pi\varepsilon^{-\beta}, respectively. For the purpose of showing compactness, this distinction is immaterial, since it is only the uniform boundedness of the preferred twist 2πN2\pi N in L2L^{2} that was used in (3.11) to obtain compactness. Using β<1/2\beta<1/2, we can estimate

|wε|22πεβ2πεβL2\displaystyle\left\||w_{\varepsilon}|^{2}2\pi\lfloor\varepsilon^{-\beta}\rfloor-2\pi\varepsilon^{-\beta}\right\|_{L^{2}} (|wε|21)2πεβL2+2πεβ2πεβL2,\displaystyle\leq\left\|(|w_{\varepsilon}|^{2}-1)2\pi\lfloor\varepsilon^{-\beta}\rfloor\right\|_{L^{2}}+\left\|2\pi\lfloor\varepsilon^{-\beta}\rfloor-2\pi\varepsilon^{-\beta}\right\|_{L^{2}},
2π((|wε|21)ε1/2L2+1)\displaystyle\leq 2\pi\left(\left\|(|w_{\varepsilon}|^{2}-1)\varepsilon^{-1/2}\right\|_{L^{2}}+1\right)
2π(2C0+1),\displaystyle\leq 2\pi\left(2\sqrt{C_{0}}+1\right),

so we are done. ∎

Proof of Theorem 5.3.

We begin with the lower-semicontinuity condition. Let wεww_{\varepsilon}\to w in L2L^{2}. We can assume that

lim infε0Fε(wε)C0<,\liminf_{\varepsilon\to 0}F_{\varepsilon}(w_{\varepsilon})\leq C_{0}<\infty, (5.25)

otherwise the lower-semicontinuity is trivial. The proof is similar to the proof of (3.19) in Theorem 3.1. Also, due to (5.24), it is enough to show that

lim infε0\displaystyle\liminf_{\varepsilon\to 0} 01ε2|wε|2+14ε(|wε|21)2+L2(𝒯(wε)+|wε|22πεβ2πεβ)2dx\displaystyle\int_{0}^{1}\frac{\varepsilon}{2}\left|{w_{\varepsilon}^{\prime}}\right|^{2}+\frac{1}{4\varepsilon}(\left|{w_{\varepsilon}}\right|^{2}-1)^{2}+\frac{L}{2}(\mathcal{T}(w_{\varepsilon})+|w_{\varepsilon}|^{2}2\pi\lfloor\varepsilon^{-\beta}\rfloor-2\pi\varepsilon^{-\beta})^{2}\,dx
E0,A(w).\displaystyle\geq E_{0,A}(w). (5.26)

First, for the twist term, it must be verified that under the assumption (5.25),

lim infε001L2(𝒯(wε)+|wε|22π\displaystyle\liminf_{\varepsilon\to 0}\int_{0}^{1}\frac{L}{2}(\mathcal{T}(w_{\varepsilon})+|w_{\varepsilon}|^{2}2\pi εβ2πεβ)2dx\displaystyle\lfloor\varepsilon^{-\beta}\rfloor-2\pi\varepsilon^{-\beta})^{2}\,dx
L201(𝒯(w)2πA)2𝑑x.\displaystyle\geq\frac{L}{2}\int_{0}^{1}\displaystyle(\mathcal{T}(w)-2\pi A)^{2}\,dx. (5.27)

In Theorem 3.1, after (3.23), we proved the inequality

K(12q)4(θε)24πN(ρε)2θε+4π2N2dxK(12q)4(θ)24πNθ+4π2N2dx,\int_{K}(1-2^{-q})^{4}(\theta_{\varepsilon_{\ell}}^{\prime})^{2}-4\pi N(\rho_{\varepsilon_{\ell}})^{2}\theta_{\varepsilon_{\ell}}^{\prime}+4\pi^{2}N^{2}\,dx\geq\int_{K}(1-2^{-q})^{4}(\theta^{\prime})^{2}-4\pi N\theta^{\prime}+4\pi^{2}N^{2}\,dx,

where KK is a compact set on which θεθ\theta_{\varepsilon_{\ell}}^{\prime}\rightharpoonup\theta^{\prime} and ρε12q\rho_{\varepsilon_{\ell}}\geq 1-2^{-q}, followed by an exhaustion argument in KK and qq to prove lower-semicontinuity of the twist in (3.27). The corresponding inequality to be verified in this case is

K(12q)4(θε)2\displaystyle\int_{K}(1-2^{-q})^{4}(\theta_{\varepsilon_{\ell}}^{\prime})^{2} +4π(|we|2εβεβ)|wε|2θε+4π2(|we|2εβεβ)2dx\displaystyle+4\pi\left(|w_{e_{\ell}}|^{2}\lfloor\varepsilon_{\ell}^{-\beta}\rfloor-\varepsilon_{\ell}^{-\beta}\right)|w_{\varepsilon_{\ell}}|^{2}\theta_{\varepsilon_{\ell}}^{\prime}+4\pi^{2}\left(|w_{e_{\ell}}|^{2}\lfloor\varepsilon_{\ell}^{-\beta}\rfloor-\varepsilon_{\ell}^{-\beta}\right)^{2}\,dx
K(12q)4(θ)24πAθ+4π2A2dx,\displaystyle\geq\int_{K}(1-2^{-q})^{4}(\theta^{\prime})^{2}-4\pi A\theta^{\prime}+4\pi^{2}A^{2}\,dx, (5.28)

which is the left-hand side of (5.27) expanded out and estimated using |wε|12q|w_{\varepsilon_{\ell}}|\geq 1-2^{-q} on KK, on which θεθ\theta_{\varepsilon_{\ell}}^{\prime}\rightharpoonup\theta^{\prime}. The desired inequality (5.28) would follow immediately from the weak convergence of θε\theta_{\varepsilon_{\ell}}^{\prime} and the two conditions

εβ|wε|2εβA in L2\varepsilon_{\ell}^{-\beta}-|w_{\varepsilon_{\ell}}|^{2}\lfloor\varepsilon_{\ell}^{-\beta}\rfloor\to A\textup{ in }L^{2} (5.29)

and

|wε|2(εβ|wε|2εβ)A in L2,|w_{\varepsilon_{\ell}}|^{2}(\varepsilon_{\ell}^{-\beta}-|w_{\varepsilon_{\ell}}|^{2}\lfloor\varepsilon_{\ell}^{-\beta}\rfloor)\to A\textup{ in }L^{2}, (5.30)

which we check in turn. First for (5.29), we estimate

εβ|wε|2εβAL2εβεβAL2+(1|wε|2)εβL2.\displaystyle\left\|\varepsilon_{\ell}^{-\beta}-|w_{\varepsilon_{\ell}}|^{2}\lfloor\varepsilon_{\ell}^{-\beta}\rfloor-A\right\|_{L^{2}}\leq\left\|\varepsilon_{\ell}^{-\beta}-\lfloor\varepsilon_{\ell}^{-\beta}\rfloor-A\right\|_{L^{2}}+\left\|(1-|w_{\varepsilon_{\ell}}|^{2})\lfloor\varepsilon_{\ell}^{-\beta}\rfloor\right\|_{L^{2}}.

The first term goes to zero as ε0\varepsilon\to 0 due to (5.20), and the second vanishes due to the uniform energy bound (5.25), since β<1/2\beta<1/2. Moving on to (5.30), we can repeat the argument (3.18) to find that

wεLM(C0).\|w_{\varepsilon_{\ell}}\|_{L^{\infty}}\leq M(C_{0}).

The second condition (5.30) can be shown as consequence of this LL^{\infty} bound, (5.29), and (5.25) after writing

|wε|2(εβ|wε|2εβ)A=|wε|2(εβ|wε|2εβA)+(|wε|21)A.|w_{\varepsilon_{\ell}}|^{2}(\varepsilon_{\ell}^{-\beta}-|w_{\varepsilon_{\ell}}|^{2}\lfloor\varepsilon_{\ell}^{-\beta}\rfloor)-A=|w_{\varepsilon_{\ell}}|^{2}(\varepsilon_{\ell}^{-\beta}-|w_{\varepsilon_{\ell}}|^{2}\lfloor\varepsilon_{\ell}^{-\beta}\rfloor-A)+(|w_{\varepsilon_{\ell}}|^{2}-1)A.

Choosing larger and larger KK which exhaust (0,1)(0,1) and letting qq\to\infty as in Theorem 3.1, the proof of (5.27) is finished. The remainder of the lower-semicontinuity proof follows from the proof of Theorem 3.1 and (5.24). The recovery sequence is very similar to the proof of Theorem 3.1, which is evident due to the similarity of (5.24) with EεE_{\varepsilon}, so we omit the details. We only mention that on the set of size O(ε)O(\varepsilon) where |wε|1|w_{\varepsilon}|\neq 1, the assumption β<1/2\beta<1/2 is needed to make sure the twist term vanishes in the limit ε0\varepsilon\to 0. ∎

Finally, we identify the minimizers of E0,AE_{0,A}. As in Corollary 3.6, this provides a description of all subsequential limits of a family of minimizers {uε}\{u_{\varepsilon}\} for FεF_{\varepsilon} and thus E~ε\tilde{E}_{\varepsilon}. We omit the proof since it follows the same strategy as the proof of Corollary 3.6.

Theorem 5.5.

Let N=N(A,α)N=N(A,\alpha) be the closest integer to Aα2πA-\frac{\alpha}{2\pi}, so that N{1,0,1}N\in\{-1,0,1\}. Then the global minimizer(s) of E0,AE_{0,A} are given by

  1. (i)

    the function

    u(x)=ei(2πN+α)xu(x)=e^{i(2\pi N+\alpha)x} (5.31)

    having constant twist and no jumps when

    L(2π(NA)+α)2<423.L(2\pi(N-A)+\alpha)^{2}<\frac{4\sqrt{2}}{3}. (5.32)
  2. (ii)

    the one-parameter set of functions given by

    u(x)={ei2πAxifx<x0,ei(2πAx+α2πA)ifx>x0,u(x)=\left\{\begin{matrix}e^{i2\pi Ax}&\;\mbox{if}\;x<x_{0},\\ e^{i(2\pi Ax+\alpha-2\pi A)}&\;\mbox{if}\;x>x_{0},\end{matrix}\right. (5.33)

    for any x0(0,1)x_{0}\in(0,1), that have one jump and twist 2πA2\pi A away from the jump, when

    L(2π(NA)+α)2>423.L(2\pi(N-A)+\alpha)^{2}>\frac{4\sqrt{2}}{3}. (5.34)

6 Acknowledgments.

DG acknowledges the support from NSF DMS-1729538. PS acknowledge the support from a Simons Collaboration grant 585520.

References

  • [1] Bedford, S. Global minimisers of cholesteric liquid crystal systems. arXiv preprint arXiv:1411.3599 (2014).
  • [2] Bernardino, N. R., Pereira, M. C. F., Silvestre, N. M., and da Gama, M. M. T. Structure of the cholesteric–isotropic interface. Soft matter 10, 47 (2014), 9399–9402.
  • [3] COMSOL Multiphysics® v. 5.3. http://www.comsol.com/. COMSOL AB, Stockholm, Sweden.
  • [4] Gartland, E. C., Huang, H., Lavrentovich, O. D., Palffy-Muhoray, P., Smalyukh, I. I., Kosa, T., and Taheri, B. Electric-field induced transitions in a cholesteric liquid-crystal film with negative dielectric anisotropy. Journal of Computational and Theoretical Nanoscience 7, 4 (2010), 709–725.
  • [5] Golovaty, D., Kim, Y.-K., Lavrentovich, O. D., Novack, M., and Sternberg, P. Phase transitions in nematics: textures with tactoids and disclinations. Mathematical Modelling of Natural Phenomena 15 (2020), 8.
  • [6] Golovaty, D., Novack, M., and Sternberg, P. A novel Landau-de Gennes model with quartic elastic terms. European Journal of Applied Mathematics (2020), 1–22.
  • [7] Golovaty, D., Novack, M., Sternberg, P., and Venkatraman, R. A model problem for nematic-isotropic transitions with highly disparate elastic constants. Archive for Rational Mechanics and Analysis (2020), 1–67.
  • [8] Golovaty, D., Sternberg, P., and Venkatraman, R. A Ginzburg-Landau type problem for highly anisotropic nematic liquid crystals. To appear in SIAM J. Math. Anal. (2018).
  • [9] Jerrard, R. L., and Sternberg, P. Critical points via Γ\Gamma-convergence: general theory and applications. J. Eur. Math. Soc. (JEMS) 11, 4 (2009), 705–753.
  • [10] Kohn, R. V., and Sternberg, P. Local minimisers and singular perturbations. Proc. Roy. Soc. Edinburgh Sect. A 111, 1-2 (1989), 69–84.
  • [11] Krisch, P., Heckmeier, M., and Tarumi, K. Design and synthesis of nematic liquid crystals with negative dialectric anisotropy. Liquid Crystals 26 (1999), 449–452.
  • [12] Majumdar, A., and Zarnescu, A. Landau-De Gennes theory of nematic liquid crystals: the Oseen-Frank limit and beyond. Arch. Ration. Mech. Anal. 196, 1 (2010), 227–280.
  • [13] Meiboom, S., Sethna, J. P., Anderson, P. W., and Brinkman, W. F. Theory of the blue phase of cholesteric liquid crystals. Phys. Rev. Lett. 46 (May 1981), 1216–1219.
  • [14] Modica, L. The gradient theory of phase transitions and the minimal interface criterion. Arch. Ration. Mech. Anal. 98, 2 (1987), 123–142.
  • [15] Modica, L., and Mortola, S. Un esempio di Γ\Gamma^{-}-convergenza. Boll. Un. Mat. Ital. B (5) 14, 1 (1977), 285–299.
  • [16] Mottram, N. J., and Newton, C. J. Introduction to Q-tensor theory. arXiv preprint arXiv:1409.3542 (2014).
  • [17] Outram, B. Long-pitch cholesterics. In Liquid Crystals, 2053-2563. IOP Publishing, 2018, pp. 5–1 to 5–18.
  • [18] Paterson, D. A., Gao, M., Kim, Y.-K., Jamali, A., Finley, K. L., Robles-Hernández, B., Diez-Berart, S., Salud, J., de la Fuente, M. R., Timimi, B. A., Zimmermann, H., Greco, C., Ferrarini, A., Storey, J. M. D., López, D. O., Lavrentovich, O. D., Luckhurst, G. R., and Imrie, C. T. Understanding the twist-bend nematic phase: the characterisation of 1-(4-cyanobiphenyl-4’-yloxy)-6-(4-cyanobiphenyl-4’-yl)hexane (CB6OCB) and comparison with CB7CB. Soft Matter 12 (2016), 6827–6840.
  • [19] Petrosyan, A., Shahgholian, H., and Uraltseva, N. Regularity of free boundaries in obstacle-type problems, vol. 136 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2012.
  • [20] Ravnik, M., Alexander, G. P., Yeomans, J. M., and Zumer, S. Mesoscopic modelling of colloids in chiral nematics. Faraday Discuss. 144 (2010), 159–169.
  • [21] Selinger, J. V. Interpretation of saddle-splay and the oseen-frank free energy in liquid crystals. Liquid Crystals Reviews 6, 2 (2018), 129–142.
  • [22] Smalyukh, I. I., and Lavrentovich, O. D. Defects, surface anchoring, and three-dimensional director fields in the lamellar structure of cholesteric liquid crystals as studied by fluorescence confocal polarizing microscopy. In Topology in condensed matter, vol. 150 of Springer Ser. Solid-State Sci. Springer, Berlin, 2006, pp. 205–250.
  • [23] Taylor, J. M. γ\gamma-convergence of a mean-field model of a chiral doped nematic liquid crystal to the oseen–frank description of cholesterics. Nonlinearity 33, 6 (2020), 3062.
  • [24] Virga, E. G. Variational theories for liquid crystals, vol. 8 of Applied Mathematics and Mathematical Computation. Chapman & Hall, London, 1994.