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

Asymptotic expansion of a
nonlocal phase transition energy

Serena Dipierro Stefania Patrizi Enrico Valdinoci  and  Mary Vaughan The University of Western Australia, Department of Mathematics and Statistics, 35 Stirling HWY, Crawley WA 6009, Australia [email protected], [email protected], [email protected] The University of Texas at Austin, Department of Mathematics, 2515 Speedway, Austin, TX 78751, USA [email protected]
Abstract.

We study the asymptotic behavior of the fractional Allen–Cahn energy functional in bounded domains with prescribed Dirichlet boundary conditions.

When the fractional power s(0,12)s\in\left(0,\frac{1}{2}\right), we establish the first-order asymptotic development up to the boundary in the sense of Γ\Gamma-convergence. In particular, we prove that the first-order term is the nonlocal minimal surface functional. Also, we show that, in general, the second-order term is not properly defined and intermediate orders may have to be taken into account.

For s[12,1)s\in\left[\frac{1}{2},1\right), we focus on the one-dimensional case and we prove that the first-order term is the classical perimeter functional plus a penalization on the boundary. Towards this end, we establish existence of minimizers to a corresponding fractional energy in a half-line, which provides itself a new feature with respect to the existing literature.

Key words and phrases:
Gamma-convergence, nonlocal phase transitions, fractional Allen–Cahn equation
2010 Mathematics Subject Classification:
82B26, 35R11, 49J45, 35A01.

1. Introduction

1.1. Higher-order expansions and boundary effects

In this paper, we initiate the study of the asymptotic behavior as ε0\varepsilon\searrow 0 of the nonlocal phase coexistence model

(1.1) ε(u):=aε(n×n)(Ωc×Ωc)|u(x)u(y)|2|xy|n+2s𝑑x𝑑y+bεΩW(u(x))𝑑x,\mathcal{E}_{\varepsilon}(u):=a_{\varepsilon}\iint_{(\mathbb{R}^{n}\times\mathbb{R}^{n})\setminus(\Omega^{c}\times\Omega^{c})}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy+b_{\varepsilon}\int_{\Omega}W(u(x))\,dx,

where n1n\geqslant 1, Ωn\Omega\subset\mathbb{R}^{n} is a bounded domain, WW is a double-well potential (see (1.2)), ε>0\varepsilon>0 is a small phase parameter, and aεa_{\varepsilon}, bε>0b_{\varepsilon}>0 take into account the scaling properties of the energy functional in terms of the fractional parameter s(0,1)s\in(0,1) (see (1.3)). In particular, the effect of the potential energy, which favors the pure phases of the system, is balanced by the kinetic energy that takes into account the long-range particle interactions and is ferromagnetic.

The energy (1.1) was considered in [SV-gamma, SV-dens]. In the former, the authors prove that, under appropriate scaling, the energy Γ\Gamma-converges to the classical perimeter functional when s[12,1)s\in\left[\frac{1}{2},1\right) and to the nonlocal perimeter functional when s(0,12)s\in\left(0,\frac{1}{2}\right).

In this work, we initiate the analysis of the higher-order asymptotic behavior as ε0\varepsilon\searrow 0. This is important in the study of slow motion of interfaces in the corresponding evolutionary problem and showcases improved convergence rates for the minimum values. Different than the interior result in [SV-gamma], we also take into account the nonlocal Dirichlet boundary conditions by restricting the class of minimizers to those satisfying u=gu=g a.e. on Ωc\Omega^{c} for a given function gg (see (1.16)). In this sense, the novelties provided by our work cover two distinct but intertwined aspects: on the one hand, we analyze the effect of the boundary data on the limit energy functional; on the other hand, we provide a higher-order expansion of the limit functional with respect to the perturbation parameter ε\varepsilon.

Let us now precisely describe our setting mathematically and state our main results.

We let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain of class C2C^{2}. Roughly speaking, this domain is the “container” in which the phase transition takes place. Notice in (1.1) that the set

QΩ:=(n×n)(Ωc×Ωc)=(Ω×Ω)(Ω×Ωc)(Ωc×Ω)Q_{\Omega}:=(\mathbb{R}^{n}\times\mathbb{R}^{n})\setminus(\Omega^{c}\times\Omega^{c})=(\Omega\times\Omega)\cup(\Omega\times\Omega^{c})\cup(\Omega^{c}\times\Omega)

collects all the couples (x,y)n×n(x,y)\in\mathbb{R}^{n}\times\mathbb{R}^{n} for which at least one entry lies in Ω\Omega: that is, ε\mathcal{E}_{\varepsilon} takes into account all the possible long-range interactions which involve particles inside the container. As is customary, Ωc\Omega^{c} denotes the complement of Ω\Omega in n\mathbb{R}^{n}.

To write the total energy of the system in a convenient notation, given a measurable function u:nu:\mathbb{R}^{n}\to\mathbb{R} and a measurable set A2nA\subset\mathbb{R}^{2n}, we write

u(A):=A|u(x)u(y)|2|xy|n+2s𝑑y𝑑x.u(A):=\iint_{A}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}\,dy\,dx.

In particular, if A=A1×A2A=A_{1}\times A_{2} with A1A_{1} and A2A_{2} measurable subsets of n\mathbb{R}^{n}, we write

u(A1,A2):=u(A1×A2)=A1A2|u(x)u(y)|2|xy|n+2s𝑑y𝑑x.u(A_{1},A_{2}):=u(A_{1}\times A_{2})=\int_{A_{1}}\int_{A_{2}}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}\,dy\,dx.

We assume that WC2()W\in C^{2}(\mathbb{R}) is a double-well potential satisfying

(1.2) {W(1)=W(1)=0,W(1)=W(1)=0,W(r)>0 for any r(1,1),W′′(1),W′′(1)>0.\begin{cases}W(-1)=W(1)=0,\\ W^{\prime}(-1)=W^{\prime}(1)=0,\\ W(r)>0\quad\hbox{ for any }r\in(-1,1),\\ W^{\prime\prime}(-1),\leavevmode\nobreak\ W^{\prime\prime}(1)>0.\end{cases}

Following [SV-gamma, SV-dens], the scaling in the fractional Allen–Cahn energy (1.1) is given by

(1.3) aε:={ε if s(0,12),ε|lnε| if s=12,ε2s if s(12,1)andbε:={ε12s if s(0,12),1|lnε| if s=12,1 if s(12,1).a_{\varepsilon}:=\begin{cases}\varepsilon&{\mbox{ if }}s\in\left(0,\frac{1}{2}\right),\\ \displaystyle\frac{\varepsilon}{|\ln\varepsilon|}&{\mbox{ if }}s=\frac{1}{2},\\ \varepsilon^{2s}&{\mbox{ if }}s\in\left(\frac{1}{2},1\right)\end{cases}\qquad\hbox{and}\qquad b_{\varepsilon}:=\begin{cases}\varepsilon^{1-2s}&{\mbox{ if }}s\in\left(0,\frac{1}{2}\right),\\ \displaystyle\frac{1}{|\ln\varepsilon|}&{\mbox{ if }}s=\frac{1}{2},\\ 1&{\mbox{ if }}s\in\left(\frac{1}{2},1\right).\end{cases}

As s1s\nearrow 1, the functional in (1.1) recovers the classical Allen–Cahn phase transition energy.

Let g:n[1,1]g:\mathbb{R}^{n}\to[-1,1] be a given Lipschitz continuous function. We define

(1.4) mε:=min{ε(u)s.t.u:nis measurable,u=ga.e. inΩcandugHs(n)}.m_{\varepsilon}:=\min\left\{\mathcal{E}_{\varepsilon}(u)\leavevmode\nobreak\ \hbox{s.t.}\,\,\leavevmode\nobreak\ \begin{subarray}{c}\displaystyle u:\mathbb{R}^{n}\to\mathbb{R}\leavevmode\nobreak\ \hbox{is measurable},\\ \displaystyle u=g\leavevmode\nobreak\ \hbox{a.e.\leavevmode\nobreak\ in}\leavevmode\nobreak\ \Omega^{c}\leavevmode\nobreak\ \hbox{and}\leavevmode\nobreak\ u-g\in H^{s}(\mathbb{R}^{n})\end{subarray}\right\}.

We establish a first-order asymptotic expansion of mεm_{\varepsilon} for any fixed s(0,1)s\in(0,1).

For s(0,12)s\in(0,\frac{1}{2}), we prove a first-order expansion in any spatial dimension. Morally speaking, we show that mε/εm_{\varepsilon}/\varepsilon approximates the fractional perimeter of a nonlocal minimal surface in Ω\Omega with prescribed Dirichlet boundary data as ε0\varepsilon\searrow 0. See the seminal paper [CRS] for more on nonlocal minimal surfaces. The precise result goes as follows.

Theorem 1.1.

Let n1n\geqslant 1, s(0,12)s\in\left(0,\frac{1}{2}\right) and Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain of class C2C^{2}. It holds that

(1.5) mε=εm1+o(ε)m_{\varepsilon}=\varepsilon m_{1}+o(\varepsilon)

where

m1:=infus.t.u|Ω=χEχEc[u(Ω,Ω)+2Ω×Ωc|u(x)g(y)|2|xy|n+2s𝑑x𝑑y].m_{1}:=\inf_{\begin{subarray}{c}u\leavevmode\nobreak\ \text{s.t.}\\ u|_{\Omega}=\chi_{E}-\chi_{E^{c}}\end{subarray}}\left[u(\Omega,\Omega)+2\displaystyle\iint_{\Omega\times\Omega^{c}}\frac{|u(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy\right].

Here, the infimum is taken over measurable sets EnE\subset\mathbb{R}^{n}.

Here above and in what follows, χA\chi_{A} denotes the characteristic function associated to a measurable set AnA\subset\mathbb{R}^{n}.

Notably, for s(0,12)s\in(0,\frac{1}{2}) and in dimension n=1n=1, we can find a specific scenario in which mεm_{\varepsilon} does not have a meaningful second-order expansion and, in fact, intermediate fractional powers of ε\varepsilon could be needed for a meaningful expansion (see Theorem 1.9 and Corollary 1.10 below).

The case s[12,1)s\in\left[\frac{1}{2},1\right) is fundamentally different as the long-range interactions are lost in the limit as ε0\varepsilon\searrow 0. In this paper, we provide the first-order asymptotic expansion of mεm_{\varepsilon} in dimension n=1n=1. Since the proof is already quite involved, we leave the general case n2n\geqslant 2 and the higher-order asymptotic expansion of mεm_{\varepsilon} as future work.

We use the standard notation P(E,Ω)P(E,\Omega) to denote the (classical) perimeter of a set EE in an open set Ω\Omega\subset\mathbb{R}, see [Giusti].

For a function u:u:\mathbb{R}\to\mathbb{R} such that u|Ω=χEχEcu\big{|}_{\Omega}=\chi_{E}-\chi_{E^{c}} for some set EE\subset\mathbb{R} with P(E,Ω)<+P(E,\Omega)<+\infty, the trace of uu along Ω\partial\Omega, given by

(1.6) Tu(x):=limΩyxu(x) for all xΩ,Tu(x):=\lim_{\Omega\ni y\to x}u(x)\quad{\mbox{ for all }}x\in\partial\Omega,

is well-defined. As an abuse of notation in this setting, we often write u(x)u(x) in place of Tu(x)Tu(x), even when xΩx\in\partial\Omega.

When s[12,1)s\in\left[\frac{1}{2},1\right), we also take an extra assumption to control the behavior of the functions considered near the boundary. This is a technical assumption that we aim at relaxing in a forthcoming work, but, roughly speaking, it is used in our setting to prevent additional oscillations of minimizing heteroclinics.

This technical hypothesis goes as follows. Suppose that Ω=(x¯1,x¯2)\Omega=(\bar{x}_{1},\bar{x}_{2})\subset\mathbb{R}. Let κ[0,|Ω|2)\kappa\in\left[0,\frac{|\Omega|}{2}\right). Consider the class YκY_{\kappa} of measurable functions u:u:\mathbb{R}\to\mathbb{R} such that, for i=1,2i=1,2,

(1.7) ifg(x¯i)>0,thenu(x)g(x¯i)for allxΩBκ(x¯i);ifg(x¯i)<0,thenu(x)g(x¯i)for allxΩBκ(x¯i);ifg(x¯i)=0,then eitheru(x)g(x¯i)oru(x)g(x¯i)for allxΩBκ(x¯i).\begin{array}[]{rl}\hbox{if}\leavevmode\nobreak\ g(\bar{x}_{i})>0,&\hbox{then}\leavevmode\nobreak\ u(x)\geqslant g(\bar{x}_{i})\leavevmode\nobreak\ \hbox{for all}\leavevmode\nobreak\ x\in\Omega\cap B_{\kappa}(\bar{x}_{i});\\ \hbox{if}\leavevmode\nobreak\ g(\bar{x}_{i})<0,&\hbox{then}\leavevmode\nobreak\ u(x)\leqslant g(\bar{x}_{i})\leavevmode\nobreak\ \hbox{for all}\leavevmode\nobreak\ x\in\Omega\cap B_{\kappa}(\bar{x}_{i});\\ \hbox{if}\leavevmode\nobreak\ g(\bar{x}_{i})=0,&\hbox{then either}\leavevmode\nobreak\ u(x)\geqslant g(\bar{x}_{i})\leavevmode\nobreak\ \hbox{or}\leavevmode\nobreak\ u(x)\leqslant g(\bar{x}_{i})\leavevmode\nobreak\ \hbox{for all}\leavevmode\nobreak\ x\in\Omega\cap B_{\kappa}(\bar{x}_{i}).\end{array}

We point out that when κ=0\kappa=0, the set Y0Y_{0} consists of all the measurable functions u:u:\mathbb{R}\to\mathbb{R}.

Consider

(1.8) mεκ:=min{ε(u)s.t.uYκ,u=ga.e. inΩc,ugHs()}.m_{\varepsilon}^{\kappa}:=\min\Big{\{}\mathcal{E}_{\varepsilon}(u)\leavevmode\nobreak\ \hbox{s.t.}\;\;u\in Y_{\kappa},\;u=g\leavevmode\nobreak\ \hbox{a.e.\leavevmode\nobreak\ in}\leavevmode\nobreak\ \Omega^{c},\;u-g\in H^{s}(\mathbb{R})\Big{\}}.

In this setting, the first-order of the asymptotic expansion of the nonlocal phase coexistence functional goes as follows:

Theorem 1.2.

Let s[12,1)s\in\left[\frac{1}{2},1\right), Ω\Omega\subset\mathbb{R} be a bounded interval, and κ(0,|Ω|2)\kappa\in\left(0,\frac{|\Omega|}{2}\right). Suppose that |g|<1|g|<1 on Ω\partial\Omega.

It holds that

mεκ=εm1κ+o(ε){m_{\varepsilon}^{\kappa}}=\varepsilon{m_{1}^{\kappa}}+o(\varepsilon)

where

m1κ:=infuYκs.t.u|Ω=χEχEc[cPer(E,Ω)+ΩΨ(u(x),g(x))𝑑0(x)]{m_{1}^{\kappa}}:=\inf_{\begin{subarray}{c}{u\in Y_{\kappa}}\leavevmode\nobreak\ \text{s.t.}\\ u|_{\Omega}=\chi_{E}-\chi_{E^{c}}\end{subarray}}\left[c_{\star}\operatorname{Per}\,(E,\Omega)+\displaystyle\int_{\partial\Omega}\Psi(u(x),g(x))\,d{\mathcal{H}}^{0}(x)\right]

Here, the infimum is taken over measurable sets EE\subset\mathbb{R} such that Per(E,Ω)<+\operatorname{Per}\,(E,\Omega)<+\infty and the constant c>0c_{\star}>0 depends only on ss and WW.

Also, Ψ:{±1}×(1,1)(0,+)\Psi:\{\pm 1\}\times(-1,1)\to(0,+\infty) is a suitable function, whose explicit definition will be given in (7.3) relying on the energy of a suitable minimal layer solution.

Roughly speaking, for any γ(1,1)\gamma\in(-1,1) and s(12,1)s\in\left(\frac{1}{2},1\right), the penalization functional Ψ\Psi mentioned in Theorem 1.2 is given by

(1.9) Ψ(1,γ)=minwXγ[w(Q)+W(w(x))𝑑x],\Psi(-1,\gamma)=\min_{w\in X_{\gamma}}\left[w(Q_{\mathbb{R}^{-}})+\int_{\mathbb{R}^{-}}W(w(x))\,dx\right],

where XγX_{\gamma} is the class of functions wHs()w\in H^{s}(\mathbb{R}) such that 1wγ-1\leqslant w\leqslant\gamma, w()=1w(-\infty)=-1, and w=γw=\gamma in +\mathbb{R}^{+}, and similarly for Ψ(+1,γ)\Psi(+1,\gamma). That is, Ψ(u(x),g(x))\Psi(u(x),g(x)) for xΩx\in\partial\Omega is the minimum energy connecting the interior value u(x){±1}u(x)\in\{\pm 1\} (in the trace sense) to the boundary value γ=g(x)(1,1)\gamma=g(x)\in(-1,1).

Thus, in order to construct the penalization function Ψ\Psi, we need to construct the heteroclinic function connecting ±1\pm 1 to γ(1,1)\gamma\in(-1,1). This construction is also somewhat delicate and does not follow trivially from the constructions of the heteroclinic function connecting 1-1 to +1+1 put forth in [CABSI, PAL]. Indeed, in our setting, cutting methods and monotonicity considerations are not available. Moreover, since γ\gamma is not an equilibrium for the potential WW, a bespoke analysis is needed.

The case s=12s=\frac{1}{2} is even more delicate, due to the presence of infinite energy contributions. To overcome this difficulty, the minimal layer solution when s=12s=\frac{1}{2} will be obtained by a limit procedure from the cases s(12,1)s\in\left(\frac{1}{2},1\right).

We stress that the construction of this minimal heteroclinic connection is an important ingredient for the recovery sequence in the Γ\Gamma-convergence expansion that we propose and introduces a novelty with respect to the available methods in the nonlocal setting. Indeed, while the interior estimates of the Γ\Gamma-limits rely on the existing literature (see [SV-gamma]), the control of the boundary terms requires an ad-hoc analysis and an appropriate interpolation that involves this new heteroclinic connection.

We now describe in detail the mathematical setting related to this layer solution. Towards this end, for any AA\subseteq\mathbb{R}, we define

𝒢s(u,A):=u(QA)+AW(u(x))𝑑x\mathcal{G}_{s}(u,A):=u(Q_{A})+\int_{A}W(u(x))\,dx

and

𝒢s(u):={𝒢s(u,)ifs(12,1),lim infR+(𝒢s(u,BR)lnR)ifs=12,\mathcal{G}_{s}(u):=\begin{cases}\mathcal{G}_{s}(u,\mathbb{R}^{-})&\hbox{if}\leavevmode\nobreak\ s\in\left(\frac{1}{2},1\right),\\ \displaystyle\liminf_{R\to+\infty}\left(\frac{\mathcal{G}_{s}(u,B_{R}^{-})}{\ln R}\right)&\hbox{if}\leavevmode\nobreak\ s=\frac{1}{2},\end{cases}

where BR:=(R,0)B_{R}^{-}:=(-R,0).

Given γ(1,1)\gamma\in(-1,1), we define XγX_{\gamma} to be the closure, with respect to the HsH^{s}-seminorm, of the set

{uC()s.t.uγ,u(x)=γfor anyx>0,andthere exists xo<0 such thatu(x)=1for allx<xo}.\left\{u\in C^{\infty}(\mathbb{R})\leavevmode\nobreak\ \hbox{s.t.}\quad\begin{subarray}{c}\displaystyle u\leqslant\gamma,\\ \displaystyle u(x)=\gamma\leavevmode\nobreak\ \hbox{for any}\leavevmode\nobreak\ x>0,\leavevmode\nobreak\ \hbox{and}\\ \displaystyle\hbox{there exists\leavevmode\nobreak\ $x_{o}<0$ such that}\leavevmode\nobreak\ u(x)=-1\leavevmode\nobreak\ \hbox{for all}\leavevmode\nobreak\ x<x_{o}\end{subarray}\right\}.

Loosely speaking, XγX_{\gamma} is the family of functions in Hs()H^{s}(\mathbb{R}) that connect 1-1 at -\infty to γ\gamma in +\mathbb{R}^{+}.

With this notation, we can establish the existence of minimizers to the energy 𝒢s\mathcal{G}_{s} in the class XγX_{\gamma}:

Theorem 1.3.

Given s[12,1)s\in\left[\frac{1}{2},1\right) and γ(1,1)\gamma\in(-1,1), there exists a unique global minimizer w0Xγw_{0}\in X_{\gamma} of the energy 𝒢s\mathcal{G}_{s}.

Moreover, w0w_{0} is strictly increasing in \mathbb{R}^{-}, w0(x)(1,γ)w_{0}(x)\in(-1,\gamma) for all xx\in\mathbb{R}^{-}, w0Cs()Clocα()w_{0}\in C^{s}(\mathbb{R})\cap C^{\alpha}_{\text{loc}}(\mathbb{R}^{-}) for all α(0,1)\alpha\in(0,1), and w0w_{0} solves

(1.10) {(Δ)sw0+W(w0)=0in,w0=γin+,limxw0(x)=1.\begin{cases}(-\Delta)^{s}w_{0}+W^{\prime}(w_{0})=0&\hbox{in}\leavevmode\nobreak\ \mathbb{R}^{-},\\ w_{0}=\gamma&\hbox{in}\leavevmode\nobreak\ \mathbb{R}^{+},\\ \displaystyle\lim_{x\to-\infty}w_{0}(x)=-1.&\end{cases}

Furthermore, there exist constants CC, R1R\geqslant 1 such that

(1.11) 0w0(x)+1C|x|2sfor allx<R0\leqslant w_{0}(x)+1\leqslant\frac{C}{|x|^{2s}}\qquad\hbox{for all}\leavevmode\nobreak\ x<-R

and, for all α(0,1)\alpha\in(0,1),

(1.12) [w0]Cα((,R))CRα.[w_{0}]_{C^{\alpha}((-\infty,-R))}\leqslant\frac{C}{R^{\alpha}}.

If s(12,1)s\in\left(\frac{1}{2},1\right), then we additionally have w0Cloc2s()w_{0}\in C_{\text{loc}}^{2s}(\mathbb{R}^{-}) and there exist CC, R1R\geqslant 1 such that

(1.13) |w0(x)|C|x| for all x<Rand[w0]C2s1((,R))CR2s.|w_{0}^{\prime}(x)|\leqslant\frac{C}{|x|}\;\hbox{ for all }x<-R\qquad\hbox{and}\qquad[w_{0}]_{C^{2s-1}((-\infty,-R))}\leqslant\frac{C}{R^{2s}}.

A similar result holds when 1-1 is replaced by 11.

We faced several difficulties in proving the existence result in Theorem 1.3. First, if γ>0\gamma>0, then minimizers may instead prefer to connect 11 to γ\gamma in the sense that w0()=1w_{0}(-\infty)=1. For instance, due to the boundary condition w0γw_{0}\equiv\gamma in +\mathbb{R}^{+}, classical sliding methods as used in [PAL] are ineffectual in proving both monotonicity and uniqueness. To solve the issue, we impose the additional constraint wγw\leqslant\gamma in the class XγX_{\gamma}. Then, since the energy on the right-hand side of (1.9) is finite for s(12,1)s\in\left(\frac{1}{2},1\right), existence is proved using compactness. However, for s=12s=\frac{1}{2}, the energy is infinite and must be defined through an appropriate rescaling. In the latter case, we give a novel approach to proving existence by sending s12s\searrow\frac{1}{2}. For this, we first prove uniform asymptotics in ss near 12\frac{1}{2} for the heteroclinic orbits connecting 1-1 at -\infty to +1+1 at -\infty. See Sections 6 and 7 for full details.

Theorems 1.1 and 1.2 are a consequence of a general approach for asymptotic developments in the classical setting of Γ\Gamma-convergence which we now describe.

1.2. Asymptotic development

The notion of asymptotic development in the sense of Γ\Gamma-convergence was introduced by Anzellotti and Baldo in [Baldo1], see also [Baldo2]. We will present the definitions and details surrounding asymptotic development in our setting. For reference, let us first recall the definition of Γ\Gamma-convergence, see [Baldo1, Definition 1.1].

Definition 1.4.

For a topological space XX, let ε:X{±}\mathcal{F}_{\varepsilon}:X\to\mathbb{R}\cup\{\pm\infty\} be a family of functions with parameter ε>0\varepsilon>0. We say that :X{±}\mathcal{F}:X\to\mathbb{R}\cup\{\pm\infty\} is the Γ\Gamma-limit of ε\mathcal{F}_{\varepsilon} if the following hold.

Let εj0\varepsilon_{j}\searrow 0 as j+j\to+\infty.

  1. (1)

    For any sequence uju¯u_{j}\to\bar{u} in XX as j+j\to+\infty, it holds that

    lim infj+εj(uj)(u¯).\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})\geqslant\mathcal{F}(\bar{u}).
  2. (2)

    Given u¯X\bar{u}\in X, there exists a sequence uju¯u_{j}\to\bar{u} in XX as j+j\to+\infty such that

    lim supj+εj(uj)(u¯).\limsup_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})\leqslant\mathcal{F}(\bar{u}).

In this case, we write =Γlimε0ε\mathcal{F}=\Gamma-\lim_{\varepsilon\searrow 0}\mathcal{F}_{\varepsilon} in XX.

Consider the space XX of all the measurable functions u:nu:\mathbb{R}^{n}\to\mathbb{R} such that the restriction of uu to Ω\Omega belongs to L1(Ω)L^{1}(\Omega). We endow XX, as well as its subspaces, with the metric of L1(Ω)L^{1}(\Omega):

(1.14) a sequence of functions ujX converges to u¯Xif uju¯L1(Ω)0 as j+.\begin{split}&{\mbox{a sequence of functions\leavevmode\nobreak\ $u_{j}\in X$ converges to\leavevmode\nobreak\ $\bar{u}\in X$}}\\ &{\mbox{if\leavevmode\nobreak\ $\|u_{j}-\bar{u}\|_{L^{1}(\Omega)}\to 0$ as\leavevmode\nobreak\ $j\to+\infty$.}}\end{split}

Notice that the values outside Ω\Omega are neglected in this procedure.

We comprise the Dirichlet datum inside the functional by defining

(1.15) Xg:={uX s.t. u=g a.e. in Ωc and ugHs(n)}X_{g}:=\big{\{}{\mbox{$u\in X$ s.t.\leavevmode\nobreak\ $u=g$ a.e. in\leavevmode\nobreak\ $\Omega^{c}$ and\leavevmode\nobreak\ $u-g\in H^{s}(\mathbb{R}^{n})$}}\big{\}}

and

(1.16) ε(u):={ε(u) if uXg,+ if uXXg.\mathcal{F}_{\varepsilon}(u):=\begin{cases}\mathcal{E}_{\varepsilon}(u)&{\mbox{ if }}u\in X_{g},\\ +\infty&{\mbox{ if }}u\in X\setminus X_{g}.\end{cases}

Notice that XgX_{g} is the functional space used in the construction of mεm_{\varepsilon} in (1.4), so that

mε=minXgε.m_{\varepsilon}=\min_{X_{g}}\mathcal{F}_{\varepsilon}.

We also set ε(0):=ε\mathcal{F}^{(0)}_{\varepsilon}:=\mathcal{F}_{\varepsilon} and 𝒰1:=X{\mathcal{U}}_{-1}:=X.

Following [Baldo1, Baldo2] (see in particular [Baldo1, pages 109–110]), the asymptotic development of order kk\in\mathbb{N}, written as

ε=Γ(0)+ε(1)++εk(k)+o(εk),\mathcal{F}_{\varepsilon}=_{\Gamma}\mathcal{F}^{(0)}+\varepsilon\mathcal{F}^{(1)}+\dots+\varepsilon^{k}\mathcal{F}^{(k)}+o(\varepsilon^{k}),

holds in the sense of Γ\Gamma-convergence if

  1. (1)

    (0)=Γlimε0ε\mathcal{F}^{(0)}=\Gamma-\displaystyle\lim_{\varepsilon\searrow 0}\mathcal{F}_{\varepsilon} in 𝒰1{\mathcal{U}}_{-1};

  2. (2)

    for any j{0,,k1}j\in\{0,\dots,k-1\}, we have that (j+1)=Γlimε0ε(j+1)\mathcal{F}^{(j+1)}=\Gamma-\displaystyle\lim_{\varepsilon\searrow 0}\mathcal{F}_{\varepsilon}^{(j+1)} in 𝒰j{\mathcal{U}}_{j}, where

    𝒰j:={u𝒰j1 s.t. (j)(u)=mj}\displaystyle{\mathcal{U}}_{j}:=\{u\in{\mathcal{U}}_{j-1}{\mbox{ s.t. }}\mathcal{F}^{(j)}(u)=m_{j}\}
    with mj:=inf𝒰j1(j)\displaystyle m_{j}:=\inf_{{\mathcal{U}}_{j-1}}\mathcal{F}^{(j)}
    and ε(j+1):=ε(j)mjε.\displaystyle\mathcal{F}^{(j+1)}_{\varepsilon}:=\frac{\mathcal{F}^{(j)}_{\varepsilon}-m_{j}}{\varepsilon}.

One can show (see [Baldo1, pages 106 and 110]) that

{limits of minimizers ofε}𝒰k𝒰0𝒰1\{\text{limits of minimizers of}\leavevmode\nobreak\ \mathcal{E}_{\varepsilon}\}\subset\mathcal{U}_{k}\subset\dots\subset\mathcal{U}_{0}\subset\mathcal{U}_{-1}

and

(1.17) mε=m0+εm1++εkmk+o(εk).m_{\varepsilon}=m_{0}+\varepsilon m_{1}+\dots+\varepsilon^{k}m_{k}+o(\varepsilon^{k}).

Therefore, the asymptotic development provides a refined selection criteria for minimizers of ε\mathcal{E}_{\varepsilon}.

In this setting, we prove the following asymptotic behavior of ε\mathcal{E}_{\varepsilon} in (1.1) in the sense of Γ\Gamma-convergence.

Theorem 1.5.

Let n1n\geqslant 1, s(0,12)s\in\left(0,\frac{1}{2}\right) and Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain of class C2C^{2}. For all k2k\geqslant 2, it holds in the sense of Γ\Gamma-convergence that

ε=Γε(1)+o(ε)\mathcal{F}_{\varepsilon}=_{\Gamma}\varepsilon\mathcal{F}^{(1)}+o(\varepsilon)

where

(1)(u)={u(Ω,Ω)+2Ω×Ωc|u(x)g(y)|2|xy|n+2s𝑑x𝑑y if Xu=χEχEc a.e. in Ω, for some En,+ otherwise.\mathcal{F}^{(1)}(u)=\begin{cases}u(\Omega,\Omega)+2\displaystyle\iint_{\Omega\times\Omega^{c}}\frac{|u(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy&\begin{matrix}{\mbox{ if\leavevmode\nobreak\ $X\ni u=\chi_{E}-\chi_{E^{c}}$ a.e. in\leavevmode\nobreak\ $\Omega$,}}\\ {\mbox{ for some\leavevmode\nobreak\ $E\subset\mathbb{R}^{n}$,}}\end{matrix}\\ +\infty&{\mbox{ otherwise}}.\end{cases}

For s[12,1)s\in\left[\frac{1}{2},1\right), we recall (1.7) and define

(1.18) ε(u):={ε(u) if uXgYκ,+ if uX(XgYκ).\mathcal{F}_{\varepsilon}(u):=\begin{cases}\mathcal{E}_{\varepsilon}(u)&{\mbox{ if }}u\in X_{g}\cap Y_{\kappa},\\ +\infty&{\mbox{ if }}u\in X\setminus(X_{g}\cap Y_{\kappa}).\end{cases}

Here, XgYκX_{g}\cap Y_{\kappa} is used in the construction of mεκm_{\varepsilon}^{\kappa} in (1.8). In this setting, we prove the following first-order asymptotic of ε\mathcal{F}_{\varepsilon}.

Theorem 1.6.

Let s[12,1)s\in\left[\frac{1}{2},1\right)Ω\Omega\subset\mathbb{R} be a bounded interval, and κ(0,|Ω|2)\kappa\in\left(0,\frac{|\Omega|}{2}\right). Suppose that |g|<1|g|<1 on Ω\partial\Omega.

It holds in the sense of Γ\Gamma-convergence that

ε=Γ(0)+ε(1)+o(ε)\mathcal{F}_{\varepsilon}=_{\Gamma}\mathcal{F}^{(0)}+\varepsilon\mathcal{F}^{(1)}+o(\varepsilon)

where

(0)(u):=χ(12,1)(s)ΩW(u(x))𝑑xfor alluX{\mathcal{F}}^{(0)}(u):=\chi_{\left(\frac{1}{2},1\right)}(s)\int_{\Omega}W(u(x))dx\qquad\hbox{for all}\leavevmode\nobreak\ u\in X

and

(1)(u):={cPer(E,Ω)+ΩΨ(u(x),g(x))𝑑0(x) if XYκu=χEχEc a.e. in Ω, for some E s.t. Per(E,Ω)<+,+ otherwise.{\mathcal{F}}^{(1)}(u):=\begin{cases}c_{\star}\operatorname{Per}\,(E,\Omega)+\displaystyle\int_{\partial\Omega}\Psi(u(x),g(x))\,d{\mathcal{H}}^{0}(x)&\begin{matrix}{\mbox{ if\leavevmode\nobreak\ ${X\cap Y_{\kappa}}\ni u=\chi_{E}-\chi_{E^{c}}$ a.e. in\leavevmode\nobreak\ $\Omega$,}}\\ {\mbox{ for some\leavevmode\nobreak\ $E\subset\mathbb{R}$\leavevmode\nobreak\ \hbox{s.t.}\leavevmode\nobreak\ $\operatorname{Per}\,(E,\Omega)<+\infty$,}}\end{matrix}\\ +\infty&{\mbox{ otherwise.}}\end{cases}
Remark 1.7.

The limit (0)=Γlimε0ε\mathcal{F}^{(0)}=\Gamma-\lim_{\varepsilon\searrow 0}\mathcal{F}_{\varepsilon} in XX in Theorem 1.6 holds more generally for Ωn\Omega\subset\mathbb{R}^{n} in all dimensions n1n\geqslant 1, for all s(0,1)s\in(0,1), for κ=0\kappa=0, and allowing |g|=1|g|=1 on Ω\partial\Omega. See Lemma 2.1 for the precise statement.

Remark 1.8.

We will see in the proof that the lim sup\limsup-inequality in the first-order Γ\Gamma-convergence in Theorem 1.6 holds for κ=0\kappa=0, see Proposition 9.10.

In this framework and recalling (1.17), Theorems 1.1 and 1.2 are a consequence of Theorems 1.5 and 1.6, respectively.

For s(0,12)s\in\left(0,\frac{1}{2}\right), notice in Theorems 1.1 and 1.5 that the nonlocal energies ε\mathcal{F}_{\varepsilon} with exterior boundary conditions ugu\equiv g in Ωc\Omega^{c} give rise to a nonlocal energy with exterior boundary conditions. In contrast, for s[12,1)s\in\left[\frac{1}{2},1\right), we see in Theorems 1.2 and 1.6 that the limiting energy localizes both in the interior and at the boundary, in the sense that the penalization energy only sees gg on Ω\partial\Omega.

For s(0,12)s\in(0,\frac{1}{2}) and in a specific one-dimensional setting, we prove by direct calculation that ε\mathcal{F}_{\varepsilon} does not have a meaningful asymptotic development of order 22 or in fact any non-integer order μ+1>22s\mu+1>2-2s. We present the precise result here. An analogous statement is expected to hold in a more general setting and is left for future work.

Theorem 1.9.

Let n=1n=1, s(0,12)s\in\left(0,\frac{1}{2}\right), Ω:=(1,1)\Omega:=(-1,1), and g:=u¯:=χ(0,+)χ(,0)g:=\bar{u}:=\chi_{(0,+\infty)}-\chi_{(-\infty,0)}. Then, there exists a sequence vεXv_{\varepsilon}\in X such that vεu¯v_{\varepsilon}\to\bar{u} in XX as ε0\varepsilon\searrow 0, and

ε(1)(vε)m1=2s(12sω)12s2sς12sε12s.\mathcal{F}_{\varepsilon}^{(1)}(v_{\varepsilon})-m_{1}=-2s\left(\frac{1-2s}{\omega}\right)^{\frac{1-2s}{2s}}\varsigma^{\frac{1}{2s}}\varepsilon^{1-2s}.

In particular,

limε0ε(2)(vε)=\lim_{\varepsilon\searrow 0}\mathcal{F}_{\varepsilon}^{(2)}(v_{\varepsilon})=-\infty

and, for all μ>12s\mu>1-2s,

limε0ε(1)(vε)m1εμ=.\lim_{\varepsilon\searrow 0}\frac{\mathcal{F}_{\varepsilon}^{(1)}(v_{\varepsilon})-m_{1}}{\varepsilon^{\mu}}=-\infty.
Corollary 1.10.

In the setting of Theorem 1.9, we have that

(2)(u)={+if uX𝒰1,if u𝒰1.{\mathcal{F}}^{(2)}(u)=\begin{dcases}+\infty&{\mbox{if }}u\in X\setminus{\mathcal{U}}_{1},\\ -\infty&{\mbox{if }}u\in{\mathcal{U}}_{1}.\end{dcases}

Since the inaugural works [Baldo1, Baldo2] there have been several papers devoted to asymptotic development in the local setting. See for instance [Braides] for additional details. See also [MR759767, MR2971613, MR3385247], where a second-order Γ\Gamma-convergence expansion is produced for a classical, one-dimensional, Modica-Mortola energy functional, and [DalMaso, Leoni], for the higher-dimensional case.

To the best of our knowledge, we are the first to consider asymptotic development in the nonlocal setting. Our main obstacle to overcome is understanding the penalization function Ψ\Psi and its role at the boundary for s[12,1)s\in\left[\frac{1}{2},1\right).

1.3. Organization of the paper

The rest of the paper is organized as follows. First, in Section 2, we prove Γ\Gamma-convergence to (0)\mathcal{F}^{(0)} as described in Remark 2.2. Then, we set notation for first-order Γ\Gamma-convergence in Section 3. The proof of Theorem 1.5 and a discussion surrounding Theorem 1.9 for s(0,12)s\in\left(0,\frac{1}{2}\right) is in Section 4. After that, we will assume for the remainder of the paper that s[12,1)s\in\left[\frac{1}{2},1\right) and set notation in Section 5. Background and preliminaries on heteroclinic connections are provided in Section 6. Section 7 contains the proof of Theorem 1.3 and the construction of the penalization function Ψ\Psi. To prove Theorem 1.6 for s[12,1)s\in\left[\frac{1}{2},1\right), we establish the lim inf\liminf-inequality in Section 8 (see Proposition 8.3) and the lim sup\limsup-inequality in Section 9 (see Proposition 9.10). Lastly, we collect some auxiliary energy estimates in Appendix A.

2. Computation of (0)\mathcal{F}^{(0)}

In this section, we will establish the zero-th order term in Theorems 1.5 and 1.6 in a general setting. Let XgX_{g} be as in (1.15) and ε\mathcal{F}_{\varepsilon} as in (1.16) for all s(0,1)s\in(0,1) (i.e. κ=0\kappa=0 for s[12,1)s\in\left[\frac{1}{2},1\right)).

Lemma 2.1.

Let n1n\geqslant 1, s(0,1)s\in(0,1) and Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain of class C2C^{2}. It holds that (0)=Γlimε0ε\mathcal{F}^{(0)}=\displaystyle\Gamma-\lim_{\varepsilon\searrow 0}\mathcal{F}_{\varepsilon} where

(0)(u)=χ(1/2, 1)(s)ΩW(u(x))𝑑xfor alluX.\mathcal{F}^{(0)}(u)=\chi_{(1/2,\,1)}(s)\,\int_{\Omega}W(u(x))\,dx\qquad\hbox{for all}\leavevmode\nobreak\ u\in X.
Proof.

Take a sequence εj0\varepsilon_{j}\searrow 0.

We first show that if uju_{j} is a sequence in XX with uju¯u_{j}\to\bar{u} as j+j\to+\infty, then

(2.1) lim infj+εj(uj)χ(1/2, 1)(s)ΩW(u¯(x))𝑑x.\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})\geqslant\chi_{(1/2,\,1)}(s)\,\int_{\Omega}W(\bar{u}(x))\,dx.

We recall that the notion of convergence in XX is the one in (1.14).

Notice that if

lim infj+εj(uj)=+,\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})=+\infty,

the claim in (2.1) is obvious. Hence, we suppose that

lim infj+εj(uj)=L<+.\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})=L<+\infty.

In this case, we take a subsequence ujku_{j_{k}} realizing the above limit.

Moreover, we recall (1.1) and (1.16) and we observe that

εjk(ujk)εjk(ujk)bεjkΩW(ujk(x))𝑑x.\mathcal{F}_{\varepsilon_{j_{k}}}(u_{j_{k}})\geqslant\mathcal{E}_{\varepsilon_{j_{k}}}(u_{j_{k}})\geqslant b_{\varepsilon_{j_{k}}}\,\int_{\Omega}W(u_{j_{k}}(x))\,dx.

We also point out that, as ε0\varepsilon\searrow 0, we have that bε0b_{\varepsilon}\to 0 when s(0,12]s\in\left(0,\frac{1}{2}\right] and bε1b_{\varepsilon}\to 1 when s(12,1)s\in\left(\frac{1}{2},1\right), that is

(2.2) limε0bε=χ(1/2, 1)(s).\lim_{\varepsilon\searrow 0}b_{\varepsilon}=\chi_{(1/2,\,1)}(s).

Furthermore, by (1.14), we take a further subsequence, still denoted by ujku_{j_{k}}, that converges to u¯\bar{u} a.e. in Ω\Omega. Hence, by Fatou’s Lemma,

L=limk+εjk(ujk)limk+bεjkΩW(ujk(x))𝑑xχ(1/2, 1)(s)ΩW(u¯(x))𝑑x,L=\lim_{k\to+\infty}\mathcal{F}_{\varepsilon_{j_{k}}}(u_{j_{k}})\geqslant\lim_{k\to+\infty}b_{\varepsilon_{j_{k}}}\,\int_{\Omega}W(u_{j_{k}}(x))\,dx\geqslant\chi_{(1/2,\,1)}(s)\,\int_{\Omega}W(\bar{u}(x))\,dx,

which proves (2.1).

Now, to complete the proof of Lemma 2.1, we show that for every u¯X\bar{u}\in X there exists a sequence ujXu_{j}\in X which converges to u¯\bar{u} as j+j\to+\infty and such that

(2.3) lim supj+εj(uj)χ(1/2, 1)(s)ΩW(u¯(x))𝑑x.\limsup_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})\leqslant\chi_{(1/2,\,1)}(s)\,\int_{\Omega}W(\bar{u}(x))\,dx.

To construct such a recovery sequence, we perform several surgeries, such as truncations, mollifications and cutoffs (roughly speaking one smooths a bit the function u¯\bar{u} and glues it to gg smoothly near the boundary). The arguments are, in a sense, of elementary nature, but they require some delicate quantifications.

For this, we fix \ell\in\mathbb{N}. For any MM\in\mathbb{N}, we define

u¯M:=max{M,min{u¯,M}}.\bar{u}_{M}:=\max\big{\{}-M,\,\min\{\bar{u},\,M\}\big{\}}.

Notice that u¯Mu¯\bar{u}_{M}\to\bar{u} a.e. in Ω\Omega as M+M\to+\infty and |u¯M||u¯|L1(Ω)|\bar{u}_{M}|\leqslant|\bar{u}|\in L^{1}(\Omega). Therefore, by the Dominated Convergence Theorem, we have that

limM+u¯u¯ML1(Ω)=0.\lim_{M\to+\infty}\|\bar{u}-\bar{u}_{M}\|_{L^{1}(\Omega)}=0.

In particular, we can find MM_{\ell}\in\mathbb{N} such that

(2.4) u¯u¯ML1(Ω)1.\|\bar{u}-\bar{u}_{M_{\ell}}\|_{L^{1}(\Omega)}\leqslant\frac{1}{\ell}.

Now, for any δ>0\delta>0, we define Ωδ\Omega_{\delta} to be the set of all the points of Ω\Omega which are at distance larger than δ\delta from Ω\partial\Omega. We set u¯,δ:=u¯MχΩ4δ\bar{u}_{\ell,\delta}:=\bar{u}_{M_{\ell}}\chi_{\Omega_{4\delta}}.

Notice that

(2.5) |u¯,δ||u¯M|M,|\bar{u}_{\ell,\delta}|\leqslant|\bar{u}_{M_{\ell}}|\leqslant M_{\ell},

and, again by the Dominated Convergence Theorem, we have that

limδ0u¯Mu¯,δL1(Ω)=limδ0u¯Mu¯MχΩ4δL1(Ω)=0.\lim_{\delta\searrow 0}\|\bar{u}_{M_{\ell}}-\bar{u}_{\ell,\delta}\|_{L^{1}(\Omega)}=\lim_{\delta\searrow 0}\|\bar{u}_{M_{\ell}}-\bar{u}_{M_{\ell}}\chi_{\Omega_{4\delta}}\|_{L^{1}(\Omega)}=0.

Therefore, we can find δ>0\delta_{\ell}>0 sufficiently small such that

(2.6) u¯Mu¯,δL1(Ω)1\|\bar{u}_{M_{\ell}}-\bar{u}_{\ell,\delta_{\ell}}\|_{L^{1}(\Omega)}\leqslant\frac{1}{\ell}

and also that

(2.7) |ΩΩδ|1(M+supΩ|g|).\big{|}\Omega\setminus\Omega_{\delta_{\ell}}\big{|}\leqslant\frac{1}{\ell\,\left(M_{\ell}+\sup_{\Omega}|g|\right)}.

Now we perform a mollification argument. Let ϕC0(B1,[0,1])\phi\in C^{\infty}_{0}(B_{1},[0,1]). For any η>0\eta>0, we define

ϕη(x):=1ηnϕ(xη)andu~,η:=u¯,δϕη.\phi_{\eta}(x):=\frac{1}{\eta^{n}}\phi\left(\frac{x}{\eta}\right)\qquad{\mbox{and}}\qquad\tilde{u}_{\ell,\eta}:=\bar{u}_{\ell,\delta_{\ell}}*\phi_{\eta}.

By construction, u~,ηC(Ω)\tilde{u}_{\ell,\eta}\in C^{\infty}(\Omega) and

(2.8) limη0u¯,δu~,ηL1(Ω)=0.\lim_{\eta\searrow 0}\|\bar{u}_{\ell,\delta_{\ell}}-\tilde{u}_{\ell,\eta}\|_{L^{1}(\Omega)}=0.

In addition, by (2.5),

(2.9) u~,ηL(n)u¯,δL(n)M.\|\tilde{u}_{\ell,\eta}\|_{L^{\infty}(\mathbb{R}^{n})}\leqslant\|\bar{u}_{\ell,{\delta_{\ell}}}\|_{L^{\infty}(\mathbb{R}^{n})}\leqslant M_{\ell}.

We also claim that, if η(0,δ)\eta\in\left(0,{\delta_{\ell}}\right),

(2.10) u~,η\tilde{u}_{\ell,\eta} is supported in Ω2δ\Omega_{2\delta_{\ell}}.

To check this, suppose that u~,η(xo)0\tilde{u}_{\ell,\eta}(x_{o})\neq 0, that is

0nu¯,δ(xoy)ϕ(yη)𝑑y=Bηu¯,δ(xoy)ϕ(yη)𝑑y.0\neq\int_{\mathbb{R}^{n}}\bar{u}_{\ell,\delta_{\ell}}(x_{o}-y)\,\phi\left(\frac{y}{\eta}\right)\,dy=\int_{B_{\eta}}\bar{u}_{\ell,\delta_{\ell}}(x_{o}-y)\,\phi\left(\frac{y}{\eta}\right)\,dy.

This implies that there exists yoBηy_{o}\in B_{\eta} such that

0u¯,δ(xoyo)=u¯M(xoyo)χΩ4δ(xoyo).0\neq\bar{u}_{\ell,\delta_{\ell}}(x_{o}-y_{o})=\bar{u}_{M_{\ell}}(x_{o}-y_{o})\,\chi_{\Omega_{4\delta_{\ell}}}(x_{o}-y_{o}).

In particular, we have that xoyoΩ4δx_{o}-y_{o}\in\Omega_{4\delta_{\ell}}. Accordingly, for any pΩp\in\partial\Omega, we have that |xoyop|>4δ|x_{o}-y_{o}-p|>4\delta_{\ell}. So, we find that

|xop||xoyop||yo|>4δη>3δ,|x_{o}-p|\geqslant|x_{o}-y_{o}-p|-|y_{o}|>4\delta_{\ell}-\eta>3\delta_{\ell},

which says that xoΩ3δx_{o}\in\Omega_{3\delta_{\ell}}, thus establishing (2.10).

By (2.8) and (2.10), we can find η\eta_{\ell} sufficiently small, such that u~,ηC0(Ωδ)\tilde{u}_{\ell,\eta_{\ell}}\in C^{\infty}_{0}(\Omega_{\delta_{\ell}}) and

(2.11) u¯,δu~,ηL1(Ω)1.\|\bar{u}_{\ell,\delta_{\ell}}-\tilde{u}_{\ell,\eta_{\ell}}\|_{L^{1}(\Omega)}\leqslant\frac{1}{\ell}.

Now we perform a cutoff argument. Namely, we take τC0(Ωδ/2,[0,1])\tau_{\ell}\in C^{\infty}_{0}(\Omega_{\delta_{\ell}/2},\,[0,1]), with τ=1\tau_{\ell}=1 in Ωδ\Omega_{\delta_{\ell}} and we set

u:=τu~,η+(1τ)g.u^{*}_{\ell}:=\tau_{\ell}\,\tilde{u}_{\ell,\eta_{\ell}}+(1-\tau_{\ell})g.

By construction,

(2.12) uu^{*}_{\ell} is a Lipschitz function,

which coincides with gg outside Ω\Omega, and such that

ug=τ(u~,ηg)C00,1(Ω)Hs(n).u^{*}_{\ell}-g=\tau_{\ell}\,(\tilde{u}_{\ell,\eta_{\ell}}-g)\in C^{0,1}_{0}(\Omega)\subset H^{s}(\mathbb{R}^{n}).

In particular,

(2.13) uXg.u^{*}_{\ell}\in X_{g}.

Furthermore, by (2.9), we have that, in Ω\Omega,

|u|τ|u~,η|+(1τ)|g|M+supΩ|g|.|u^{*}_{\ell}|\leqslant\tau_{\ell}\,|\tilde{u}_{\ell,\eta_{\ell}}|+(1-\tau_{\ell})|g|\leqslant M_{\ell}+\sup_{\Omega}|g|.

Therefore, by (2.7),

u~,ηuL1(Ω)=(1τ)(gu~,η)L1(Ω)(M+supΩ|g|)1τL1(Ω)\displaystyle\|\tilde{u}_{\ell,\eta_{\ell}}-u^{*}_{\ell}\|_{L^{1}(\Omega)}=\|(1-\tau_{\ell})(g-\tilde{u}_{\ell,\eta_{\ell}})\|_{L^{1}(\Omega)}\leqslant\left(M_{\ell}+\sup_{\Omega}|g|\right)\,\|1-\tau_{\ell}\|_{L^{1}(\Omega)}
(M+supΩ|g|)|ΩΩδ|1.\displaystyle\qquad\qquad\qquad\leqslant\left(M_{\ell}+\sup_{\Omega}|g|\right)\,\big{|}\Omega\setminus\Omega_{\delta_{\ell}}\big{|}\leqslant\frac{1}{\ell}.

Combining this with (2.4), (2.6) and (2.11), we find that

(2.14) u¯uL1(Ω)u¯u¯ML1(Ω)+u¯Mu¯,δL1(Ω)+u¯,δu~,ηL1(Ω)+u~,ηuL1(Ω)4.\begin{split}&\|\bar{u}-u^{*}_{\ell}\|_{L^{1}(\Omega)}\\ \leqslant\;&\|\bar{u}-\bar{u}_{M_{\ell}}\|_{L^{1}(\Omega)}+\|\bar{u}_{M_{\ell}}-\bar{u}_{\ell,\delta_{\ell}}\|_{L^{1}(\Omega)}+\|\bar{u}_{\ell,\delta_{\ell}}-\tilde{u}_{\ell,\eta_{\ell}}\|_{L^{1}(\Omega)}+\|\tilde{u}_{\ell,\eta_{\ell}}-u^{*}_{\ell}\|_{L^{1}(\Omega)}\\ \leqslant\;&\frac{4}{\ell}.\end{split}

Now we remark that, for any s(0,1)s\in(0,1),

(2.15) limε0aε=0.\lim_{\varepsilon\searrow 0}a_{\varepsilon}=0.

On the other hand, for any fixed \ell, we have that u(QΩ)<+u^{*}_{\ell}(Q_{\Omega})<+\infty, thanks to (2.12). From these observations, we conclude that there exists jj_{\ell}\in\mathbb{N} such that, for any jjj\geqslant j_{\ell},

(2.16) u(QΩ)1aεj.u^{*}_{\ell}(Q_{\Omega})\leqslant\frac{1}{\sqrt{a_{\varepsilon_{j}}}}.

Now we define

ψ():=+maxi{1,,}ji.\psi(\ell):=\ell+\max_{i\in\{1,\dots,\ell\}}j_{i}.

We observe that ψ()j\psi(\ell)\geqslant j_{\ell} and that ψ\psi is strictly increasing. So, by a linear interpolation, we can extend ψ\psi to a piecewise linear function on [1,+)[1,+\infty), which is strictly increasing and therefore invertible on its image, with ψ1\psi^{-1} strictly increasing.

As a consequence, the inequality in (2.16) holds true for every jψ()j\geqslant\psi(\ell), that is, equivalently, for every ψ1(j)\ell\leqslant\psi^{-1}(j). Hence, in particular, if we set

uj:=uψ1(j),u_{j}:=u^{*}_{\psi^{-1}(j)},

we have that

(2.17) uj(QΩ)1aεj.u_{j}(Q_{\Omega})\leqslant\frac{1}{\sqrt{a_{\varepsilon_{j}}}}.

Now we observe that

(2.18) limj+ψ1(j)=+.\lim_{j\to+\infty}\psi^{-1}(j)=+\infty.

Indeed, suppose not. Then, since ψ1\psi^{-1} is increasing, the limit above exists and

λ=:limj+ψ1(j)=sup[1,+)ψ1.\mathbb{R}\ni\lambda=:\lim_{j\to+\infty}\psi^{-1}(j)=\sup_{[1,+\infty)}\psi^{-1}.

In particular, there exists joj_{o}\in\mathbb{N} such that for any jjoj\geqslant j_{o} we have that λψ1(j)1\lambda-\psi^{-1}(j)\leqslant 1. As a consequence, for any ii\in\mathbb{N},

ψ1(jo+i)ψ1(jo)λψ1(jo)1\psi^{-1}(j_{o}+i)-\psi^{-1}(j_{o})\leqslant\lambda-\psi^{-1}(j_{o})\leqslant 1

and so

jo+i=ψ(ψ1(jo+i))ψ(ψ1(jo)+1).j_{o}+i=\psi(\psi^{-1}(j_{o}+i))\leqslant\psi(\psi^{-1}(j_{o})+1).

Sending i+i\to+\infty, we obtain a contradiction and so (2.18) is proved.

Now, from (2.14) (used here with :=ψ1(j)\ell:=\psi^{-1}(j)) and (2.18), we infer that

limj+u¯ujL1(Ω)limj+4ψ1(j)=0.\lim_{j\to+\infty}\|\bar{u}-u_{j}\|_{L^{1}(\Omega)}\leqslant\lim_{j\to+\infty}\frac{4}{\psi^{-1}(j)}=0.

So, up to a subsequence, we have that

(2.19) uju¯u_{j}\to\bar{u} a.e. in Ω\Omega, as j+j\to+\infty.

Furthermore, using (2.13) and (2.17), we find that

εj(uj)\displaystyle\mathcal{F}_{\varepsilon_{j}}(u_{j}) =\displaystyle= εj(uj)\displaystyle\mathcal{E}_{\varepsilon_{j}}(u_{j})
=\displaystyle= aεjuj(QΩ)+bεjΩW(uj(x))𝑑x\displaystyle a_{\varepsilon_{j}}u_{j}(Q_{\Omega})+b_{\varepsilon_{j}}\,\int_{\Omega}W(u_{j}(x))\,dx
\displaystyle\leqslant aεj+bεjΩW(uj(x))𝑑x.\displaystyle\sqrt{a_{\varepsilon_{j}}}+b_{\varepsilon_{j}}\,\int_{\Omega}W(u_{j}(x))\,dx.

Hence, by (2.2) and (2.15),

lim supj+εj(uj)χ(1/2, 1)(s)lim supj+ΩW(uj(x))𝑑x.\limsup_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})\leqslant\chi_{(1/2,\,1)}(s)\,\limsup_{j\to+\infty}\int_{\Omega}W(u_{j}(x))\,dx.

Therefore, recalling (2.19) and the Dominated Convergence Theorem,

lim supj+εj(uj)χ(1/2, 1)(s)ΩW(u¯(x))𝑑x.\limsup_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}(u_{j})\leqslant\chi_{(1/2,\,1)}(s)\,\int_{\Omega}W(\bar{u}(x))\,dx.

This proves (2.3).

By combining (2.1) and (2.3), we obtain that

Γlimε0ε(u)=χ(1/2, 1)(s)ΩW(u(x))𝑑x\Gamma-\displaystyle\lim_{\varepsilon\searrow 0}\mathcal{F}_{\varepsilon}(u)=\chi_{(1/2,\,1)}(s)\,\int_{\Omega}W(u(x))\,dx

in XX, which completes the proof of Lemma 2.1. ∎

Remark 2.2.

The proof of the zero-th order convergence in Theorem 1.6 for Ω\Omega\subset\mathbb{R} and κ(0,|Ω|2)\kappa\in\left(0,\frac{|\Omega|}{2}\right) follows exactly as the proof of Lemma 2.1.

3. Set up for first-order expansion

To determine the first-order term (1)\mathcal{F}^{(1)} in Theorems 1.5 and 1.6, we first describe the functionals ε(1)\mathcal{F}_{\varepsilon}^{(1)}. By Lemma 2.1, we clearly have that m0=0m_{0}=0 if s(0,12]s\in\left(0,\frac{1}{2}\right]. For s(12,1)s\in\left(\frac{1}{2},1\right),

m0=infuX(0)(u)=infuXΩW(u(x))𝑑x=0m_{0}=\inf_{u\in X}\mathcal{F}^{(0)}(u)=\inf_{u\in X}\int_{\Omega}W(u(x))\,dx=0

since we can always choose |u||u| constantly equal to 11. Therefore, the set of minimizers is

𝒰0\displaystyle{\mathcal{U}}_{0} ={uX s.t. (0)(u)=0}\displaystyle=\{u\in X{\mbox{ s.t. }}\mathcal{F}^{(0)}(u)=0\}
={Xifs(0,12],{uX s.t. u=χEχEc a.e. in Ω, for some En}ifs(12,1)\displaystyle=\begin{cases}X&\hbox{if}\leavevmode\nobreak\ s\in\left(0,\frac{1}{2}\right],\\ \{u\in X{\mbox{ s.t. }}u=\chi_{E}-\chi_{E^{c}}{\mbox{ a.e. in }}\Omega,{\mbox{ for some }}E\subset\mathbb{R}^{n}\}&\hbox{if}\leavevmode\nobreak\ s\in\left(\frac{1}{2},1\right)\end{cases}

and the functionals ε(1)\mathcal{F}_{\varepsilon}^{(1)} are

(3.1) ε(1)(u)=ε(0)(u)0ε={ε1ε(u) if uXg,+ if uXXg,={a~εu(QΩ)+b~εΩW(u(x))𝑑x if uXg,+ if uXXg,\begin{split}&\mathcal{F}^{(1)}_{\varepsilon}(u)=\frac{\mathcal{F}^{(0)}_{\varepsilon}(u)-0}{\varepsilon}=\begin{cases}\displaystyle\varepsilon^{-1}\mathcal{E}_{\varepsilon}(u)&{\mbox{ if }}u\in X_{g},\\ +\infty&{\mbox{ if }}u\in X\setminus X_{g},\end{cases}\\ &\qquad\qquad=\begin{cases}\displaystyle\tilde{a}_{\varepsilon}\,u(Q_{\Omega})+\tilde{b}_{\varepsilon}\int_{\Omega}W(u(x))\,dx&{\mbox{ if }}u\in X_{g},\\ +\infty&{\mbox{ if }}u\in X\setminus X_{g},\end{cases}\end{split}

where

(3.2) a~ε:={1 if s(0,12),|lnε|1 if s=12,ε2s1 if s(12,1)andb~ε:={ε2s if s(0,12),(ε|lnε|)1 if s=12,ε1 if s(12,1).\tilde{a}_{\varepsilon}:=\begin{cases}1&{\mbox{ if }}s\in\left(0,\frac{1}{2}\right),\\ |\ln\varepsilon|^{-1}&{\mbox{ if }}s=\frac{1}{2},\\ \varepsilon^{2s-1}&{\mbox{ if }}s\in\left(\frac{1}{2},1\right)\end{cases}\qquad\hbox{and}\qquad\tilde{b}_{\varepsilon}:=\begin{cases}\varepsilon^{-2s}&{\mbox{ if }}s\in\left(0,\frac{1}{2}\right),\\ (\varepsilon|\ln\varepsilon|)^{-1}&{\mbox{ if }}s=\frac{1}{2},\\ \varepsilon^{-1}&{\mbox{ if }}s\in\left(\frac{1}{2},1\right).\end{cases}

When s[12,1)s\in\left[\frac{1}{2},1\right), we replace XgX_{g} by XgYκX_{g}\cap Y_{\kappa} in (3.1).

4. Asymptotic development for s(0,12)s\in\left(0,\frac{1}{2}\right) and proof of Theorem 1.5

This section is devoted to the proofs of Theorem 1.5 and Theorem 1.9. We assume throughout that s(0,12)s\in\left(0,\frac{1}{2}\right) and n1n\geqslant 1 (unless otherwise stated) are fixed.

4.1. Computation of (1)\mathcal{F}^{(1)}

Here, we establish the first-order term (1)\mathcal{F}^{(1)} in Theorem 1.5.

Lemma 4.1.

Let s(0,12)s\in\left(0,\frac{1}{2}\right). It holds that (1)=Γlimε0ε(1)\displaystyle\mathcal{F}^{(1)}=\Gamma-\lim_{\varepsilon\searrow 0}\mathcal{F}^{(1)}_{\varepsilon} where (1)\mathcal{F}^{(1)} is given by

(1)(u)={u(Ω,Ω)+2Ω×Ωc|u(x)g(y)|2|xy|n+2s𝑑x𝑑y if Xu=χEχEc a.e. in Ω, for some En,+ otherwise\mathcal{F}^{(1)}(u)=\begin{cases}u(\Omega,\Omega)+2\displaystyle\iint_{\Omega\times\Omega^{c}}\frac{|u(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy&\begin{matrix}{\mbox{ if $X\ni u=\chi_{E}-\chi_{E^{c}}$ a.e. in $\Omega$,}}\\ {\mbox{ for some\leavevmode\nobreak\ $E\subset\mathbb{R}^{n}$,}}\end{matrix}\\ +\infty&{\mbox{ otherwise}}\end{cases}
Proof.

We take a sequence εj0\varepsilon_{j}\searrow 0.

First, we show that if uju_{j} is a sequence in XX with uju¯u_{j}\to\bar{u} in XX as j+j\to+\infty, then

(4.1) lim infj+εj(1)(uj)(1)(u¯).\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}^{(1)}(u_{j})\geqslant\mathcal{F}^{(1)}(\bar{u}).

To this aim, we may assume that

(4.2) ujXg,u_{j}\in X_{g},

otherwise εj(1)(uj)=+\mathcal{F}_{\varepsilon_{j}}^{(1)}(u_{j})=+\infty (recall (3.1)) and we are done.

In addition, we may suppose that

(4.3) |u¯|=1|\bar{u}|=1 a.e. in Ω\Omega,

since, if not,

ΩW(u¯(x))𝑑x>0,\displaystyle\int_{\Omega}W(\bar{u}(x))\,dx>0,

and so, by Fatou’s Lemma,

lim infj+εj(1)(uj)lim infj+1εj2sΩW(uj(x))𝑑x=+.\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}^{(1)}(u_{j})\geqslant\liminf_{j\to+\infty}\frac{1}{\varepsilon_{j}^{2s}}\,\displaystyle\int_{\Omega}W(u_{j}(x))\,dx=+\infty.

Accordingly, by (4.3) we know that u¯|Ω=χEχEc\bar{u}\big{|}_{\Omega}=\chi_{E}-\chi_{E^{c}} for some EnE\subset\mathbb{R}^{n}. By (4.2) and Fatou’s Lemma,

lim infj+εj(1)(uj)lim infj+uj(QΩ)=lim infj+[uj(Ω,Ω)+2Ω×Ωc|uj(x)g(y)|2|xy|n+2s𝑑x𝑑y]\displaystyle\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}^{(1)}(u_{j})\geqslant\liminf_{j\to+\infty}u_{j}(Q_{\Omega})=\liminf_{j\to+\infty}\left[u_{j}(\Omega,\Omega)+2\iint_{\Omega\times\Omega^{c}}\frac{|u_{j}(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy\right]
u¯(Ω,Ω)+2Ω×Ωc|u¯(x)g(y)|2|xy|n+2s𝑑x𝑑y=(1)(u¯).\displaystyle\qquad\qquad\geqslant\bar{u}(\Omega,\Omega)+2\iint_{\Omega\times\Omega^{c}}\frac{|\bar{u}(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy=\mathcal{F}^{(1)}(\bar{u}).

This proves (4.1).

Now we show that for every u¯X\bar{u}\in X there exists a sequence ujXu_{j}\in X which converges to u¯\bar{u} as j+j\to+\infty and such that

(4.4) lim supj+εj(1)(uj)(1)(u¯).\limsup_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}^{(1)}(u_{j})\leqslant\mathcal{F}^{(1)}(\bar{u}).

For this, we may suppose that u¯=χEχEc\bar{u}=\chi_{E}-\chi_{E^{c}} for some EnE\subset\mathbb{R}^{n}. Otherwise (1)(u¯)=+\mathcal{F}^{(1)}(\bar{u})=+\infty, and we are done. In particular, we have that

ΩW(u¯(x))𝑑x=0.\int_{\Omega}W(\bar{u}(x))\,dx=0.

So, we define

uj(x):={u¯(x) if xΩ,g(x) if xΩc.u_{j}(x):=\begin{cases}\bar{u}(x)&{\mbox{ if }}x\in\Omega,\\ g(x)&{\mbox{ if }}x\in\Omega^{c}.\end{cases}

Then, ujXgu_{j}\in X_{g} and we have that

lim supj+εj(1)(uj)=lim supj+[uj(QΩ)+1εj2sΩW(uj(x))𝑑x]\displaystyle\limsup_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}^{(1)}(u_{j})=\limsup_{j\to+\infty}\left[u_{j}(Q_{\Omega})+\frac{1}{\varepsilon_{j}^{2s}}\,\displaystyle\int_{\Omega}W(u_{j}(x))\,dx\right]
=lim supj+[u¯(Ω,Ω)+2Ω×Ωc|u¯(x)g(y)|2|xy|n+2s𝑑x𝑑y+0]=(1)(u¯).\displaystyle\qquad\qquad=\limsup_{j\to+\infty}\left[\bar{u}(\Omega,\Omega)+2\iint_{\Omega\times\Omega^{c}}\frac{|\bar{u}(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy+0\right]=\mathcal{F}^{(1)}(\bar{u}).

This proves (4.4).

The desired result then follows from (4.1) and (4.4). ∎

4.2. Proof of Theorem 1.5

Proof of Theorem 1.5.

The result follows from Lemmata 2.1 and 4.1. ∎

4.3. Computation for (2)\mathcal{F}^{(2)}

The rest of this section is devoted to the second-order asymptotic development when s(0,12)s\in(0,\frac{1}{2}). In light of Lemma 4.1, we have that

m1=infu𝒰0(1)(u)=infuXs.t.u|Ω=χEχEc[u(Ω,Ω)+2Ω×Ωc|u(x)g(y)|2|xy|n+2s𝑑x𝑑y].m_{1}=\inf_{u\in\mathcal{U}_{0}}\mathcal{F}^{(1)}(u)=\inf_{\begin{subarray}{c}u\in X\leavevmode\nobreak\ \text{s.t.}\\ u|_{\Omega}=\chi_{E}-\chi_{E^{c}}\end{subarray}}\left[u(\Omega,\Omega)+2\displaystyle\iint_{\Omega\times\Omega^{c}}\frac{|u(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy\right].

The set of minimizers is denoted by

𝒰1\displaystyle\mathcal{U}_{1} ={u𝒰0s.t.(1)(u)=m1}\displaystyle=\{u\in\mathcal{U}_{0}\leavevmode\nobreak\ \hbox{s.t.}\leavevmode\nobreak\ \mathcal{F}^{(1)}(u)=m_{1}\}
={uXs.t.u|Ω=χEχEcandu(Ω,Ω)+2Ω×Ωc|u(x)g(y)|2|xy|n+2s𝑑x𝑑y=m1}\displaystyle=\left\{u\in X\leavevmode\nobreak\ \hbox{s.t.}\leavevmode\nobreak\ u|_{\Omega}=\chi_{E}-\chi_{E^{c}}\leavevmode\nobreak\ \hbox{and}\leavevmode\nobreak\ u(\Omega,\Omega)+2\displaystyle\iint_{\Omega\times\Omega^{c}}\frac{|u(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy=m_{1}\right\}

and the functionals ε(2)\mathcal{F}_{\varepsilon}^{(2)} are given by

ε(2)(u)\displaystyle\mathcal{F}_{\varepsilon}^{(2)}(u) =ε(1)(u)m1ε\displaystyle=\frac{\mathcal{F}_{\varepsilon}^{(1)}(u)-m_{1}}{\varepsilon}
={1ε(u(Ω,Ω)+2u(Ω,Ωc)m1)+1ε1+2sΩW(u(x))𝑑xifuXg,+ifuXXg.\displaystyle=\begin{cases}\displaystyle\frac{1}{\varepsilon}\big{(}u(\Omega,\Omega)+2u(\Omega,\Omega^{c})-m_{1}\big{)}+\frac{1}{\varepsilon^{1+2s}}\int_{\Omega}W(u(x))\,dx&\hbox{if}\leavevmode\nobreak\ u\in X_{g},\\ +\infty&\hbox{if}\leavevmode\nobreak\ u\in X\setminus X_{g}.\end{cases}

Now we notice that ε(2)\mathcal{F}_{\varepsilon}^{(2)} Γ\Gamma-converges to ++\infty in X𝒰1X\setminus\mathcal{U}_{1}.

Lemma 4.2.

Assume s(0,12)s\in(0,\frac{1}{2}), and fix u¯X𝒰1\bar{u}\in X\setminus\mathcal{U}_{1}. Let uju_{j} be such that uju¯u_{j}\to\bar{u} in XX and εj0\varepsilon_{j}\searrow 0 as j+j\to+\infty.

Then

lim infj+εj(2)(uj)=+.\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}^{(2)}(u_{j})=+\infty.
Proof.

As in (4.2), we may assume that ujXgu_{j}\in X_{g}. Suppose first that u¯|Ω=χEχEc\bar{u}\big{|}_{\Omega}=\chi_{E}-\chi_{E^{c}} for some measurable EnE\subset\mathbb{R}^{n}. By Fatou’s Lemma and the definition of m1m_{1},

lim infj+(uj(Ω,Ω)+2Ω×Ωc|uj(x)g(y)|2|xy|n+2s𝑑x𝑑ym1)\displaystyle\liminf_{j\to+\infty}\bigg{(}u_{j}(\Omega,\Omega)+2\iint_{\Omega\times\Omega^{c}}\frac{|u_{j}(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy-m_{1}\bigg{)}
u¯(Ω,Ω)+2Ω×Ωc|u¯(x)g(y)|2|xy|n+2s𝑑x𝑑ym1\displaystyle\qquad\qquad\geqslant\bar{u}(\Omega,\Omega)+2\iint_{\Omega\times\Omega^{c}}\frac{|\bar{u}(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy-m_{1}
=(1)(u¯)m1>0.\displaystyle\qquad\qquad=\mathcal{F}^{(1)}(\bar{u})-m_{1}>0.

Consequently,

lim infj+εj(2)(uj)lim infj+1εj(uj(Ω,Ω)+2Ω×Ωc|uj(x)g(y)|2|xy|n+2s𝑑x𝑑ym1)=+.\liminf_{j\to+\infty}\mathcal{F}_{\varepsilon_{j}}^{(2)}(u_{j})\geqslant\liminf_{j\to+\infty}\frac{1}{\varepsilon_{j}}\bigg{(}u_{j}(\Omega,\Omega)+2\iint_{\Omega\times\Omega^{c}}\frac{|u_{j}(x)-g(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy-m_{1}\bigg{)}=+\infty.

Now, suppose that |{xΩs.t.u¯(x)±1}|>0|\{x\in\Omega\leavevmode\nobreak\ \hbox{s.t.}\leavevmode\nobreak\ \bar{u}(x)\neq\pm 1\}|>0. In this case,

ΩW(u¯(x))𝑑x>0.\int_{\Omega}W(\bar{u}(x))\,dx>0.

By Fatou’s Lemma,

lim infj(1εj2sΩW(uj¯(x))𝑑xm1)=+,\liminf_{j\to\infty}\left(\frac{1}{\varepsilon_{j}^{2s}}\int_{\Omega}W(\bar{u_{j}}(x))\,dx-m_{1}\right)=+\infty,

and consequently,

lim infjεj(2)(uj)lim infj1εj(1εj2sΩW(uj(x))𝑑xm1)=+.\liminf_{j\to\infty}\mathcal{F}_{\varepsilon_{j}}^{(2)}(u_{j})\geqslant\liminf_{j\to\infty}\frac{1}{\varepsilon_{j}}\left(\frac{1}{\varepsilon_{j}^{2s}}\int_{\Omega}W(u_{j}(x))\,dx-m_{1}\right)=+\infty.

This completes the proof. ∎

We now consider Γ\Gamma-convergence of ε(2)\mathcal{F}_{\varepsilon}^{(2)} in 𝒰1\mathcal{U}_{1} in dimension n=1n=1. In particular, we establish Theorem 1.9. The proof is broken up into several small observations.

Lemma 4.3.

Let n=1n=1 and s(0,12)s\in\left(0,\frac{1}{2}\right). Let u¯:=χ(0,+)χ(,0)\bar{u}:=\chi_{(0,+\infty)}-\chi_{(-\infty,0)} and δ(0,1]\delta\in(0,1]. Let also uδu_{\delta} be the minimizer of u(Q(1,1))u(Q_{(-1,1)}) among measurable functions u:u:\mathbb{R}\to\mathbb{R} with u=u¯u=\bar{u} in (δ,δ)\mathbb{R}\setminus(-\delta,\delta).

Then,

uδ(x)=u1(xδ)u_{\delta}(x)=u_{1}\left(\frac{x}{\delta}\right)

and

uδ(Q(1,1))=u¯(Q(1,1))ςδ12s,u_{\delta}(Q_{(-1,1)})=\bar{u}(Q_{(-1,1)})-\varsigma{\delta^{1-2s}},

where

ς:=u¯(Q(1,1))u1(Q(1,1)).\varsigma:=\bar{u}(Q_{(-1,1)})-u_{1}(Q_{(-1,1)}).
Proof.

Given v:v:\mathbb{R}\to\mathbb{R} with v=u¯v=\bar{u} in (δ,δ)\mathbb{R}\setminus(-\delta,\delta), we define w(x):=v(δx)w(x):=v(\delta x) and

ϕ(x,y):=|v(x)v(y)|2|u¯(x)u¯(y)|2|xy|1+2s.\phi(x,y):=\frac{|v(x)-v(y)|^{2}-|\bar{u}(x)-\bar{u}(y)|^{2}}{|x-y|^{1+2s}}.

Since ϕ\phi vanishes identically in ((δ,δ))2((1,1))2(\mathbb{R}\setminus(-\delta,\delta))^{2}\supseteq(\mathbb{R}\setminus(-1,1))^{2} and u¯(x)=u¯(δx)\bar{u}(x)=\bar{u}(\delta x) for all xx\in\mathbb{R}, we have that

v(Q(1,1))u¯(Q(1,1))=Q(1,1)ϕ(x,y)𝑑x𝑑y=×ϕ(x,y)𝑑x𝑑y\displaystyle v(Q_{(-1,1)})-\bar{u}(Q_{(-1,1)})=\iint_{Q_{(-1,1)}}\phi(x,y)\,dx\,dy=\iint_{\mathbb{R}\times\mathbb{R}}\phi(x,y)\,dx\,dy
=δ2×ϕ(δx,δy)𝑑x𝑑y=δ2Q(1,1)ϕ(δx,δy)𝑑x𝑑y\displaystyle\quad={\delta^{2}}\iint_{\mathbb{R}\times\mathbb{R}}\phi(\delta x,\delta y)\,dx\,dy={\delta^{2}}\iint_{Q_{(-1,1)}}\phi(\delta x,\delta y)\,dx\,dy
=δ2Q(1,1)|w(x)w(y)|2|u¯(x)u¯(y)|2|δxδy|1+2s𝑑x𝑑y\displaystyle\quad={\delta^{2}}\iint_{Q_{(-1,1)}}\frac{|w(x)-w(y)|^{2}-|\bar{u}(x)-\bar{u}(y)|^{2}}{|\delta x-\delta y|^{1+2s}}\,dx\,dy
=δ12s(w(Q(1,1))u¯(Q(1,1))).\displaystyle\quad={\delta^{1-2s}}\big{(}w(Q_{(-1,1)})-\bar{u}(Q_{(-1,1)})\big{)}.

From this, we obtain the desired result (we stress that u1u_{1} means uδu_{\delta} with δ:=1\delta:=1). ∎

Lemma 4.4.

In the notation of Lemma 4.3, we have that ς>0\varsigma>0.

Proof.

Due to the minimality of u1u_{1}, we know that ς0\varsigma\geqslant 0. Also, a minimizer u1u_{1} satisfies the Euler-Lagrange equation (Δ)su1=0(-\Delta)^{s}u_{1}=0 in Ω=(1,1)\Omega=(-1,1) and therefore it is necessarily continuous at the origin, ruling out the possibility for u¯\bar{u} to attain the minimum. ∎

Lemma 4.5.

In the notation of Lemma 4.3, we have that u¯\bar{u} is a minimizer of u(Q(1,1))u(Q_{(-1,1)}) among measurable functions u:{1,1}u:\mathbb{R}\to\{-1,1\} with u=1u=1 in [1,+)[1,+\infty) and u=1u=-1 in (,1](-\infty,-1].

Proof.

We know that a minimizer can be taken to be monotone (see [Alberti, Theorem 2.11]) and odd symmetric (see [MR3596708, Lemma A.1]). Since the minimizer is constrained to take values in {1,1}\{-1,1\}, the desired result follows. ∎

Corollary 4.6.

Let n=1n=1, s(0,12)s\in\left(0,\frac{1}{2}\right), Ω:=(1,1)\Omega:=(-1,1), g:=χ(0,+)χ(,0)g:=\chi_{(0,+\infty)}-\chi_{(-\infty,0)}, and ε\varepsilon, δ(0,1]\delta\in(0,1].

Let uδu_{\delta} be as in Lemma 4.3.

Then,

(4.5) uδ(Q(1,1))m1+1ε2s11W(uδ(x))𝑑x=ςδ12s+ωδε2s,u_{\delta}(Q_{{(-1,1)}})-m_{1}+\frac{1}{\varepsilon^{2s}}\int_{-1}^{1}W(u_{\delta}(x))\,dx=-\varsigma{\delta^{1-2s}}+\frac{\omega\delta}{\varepsilon^{2s}},

where ς>0\varsigma>0 is as in Lemmata 4.3 and 4.4, and

ω:=11W(u1(x))𝑑x>0.\omega:=\int_{-1}^{1}W(u_{1}(x))\,dx>0.
Proof.

It follows from Lemma 4.5 that m1=u¯(Q(1,1))=g(Q(1,1))m_{1}=\bar{u}(Q_{(-1,1)})=g(Q_{(-1,1)}) and consequently, by Lemma 4.3,

uδ(Q(1,1))m1+1ε2s11W(uδ(x))𝑑x=ςδ12s+1ε2sδδW(uδ(x))𝑑x\displaystyle u_{\delta}(Q_{(-1,1)})-m_{1}+\frac{1}{\varepsilon^{2s}}\int_{-1}^{1}W(u_{\delta}(x))\,dx=-\varsigma{\delta^{1-2s}}+\frac{1}{\varepsilon^{2s}}\int_{-\delta}^{\delta}W(u_{\delta}(x))\,dx
=ςδ12s+1ε2sδδW(u1(xδ))𝑑x=ςδ12s+δε2s11W(u1(x))𝑑x.\displaystyle\qquad=-\varsigma{\delta^{1-2s}}+\frac{1}{\varepsilon^{2s}}\int_{-\delta}^{\delta}W\left(u_{1}\left(\frac{x}{\delta}\right)\right)\,dx=-\varsigma{\delta^{1-2s}}+\frac{\delta}{\varepsilon^{2s}}\int_{-1}^{1}W(u_{1}(x))\,dx.

Since u1(0)=0u_{1}(0)=0 and u1u_{1} is continuous at the origin (see the proof of Lemma 4.4), the desired result follows. ∎

Corollary 4.7.

In the notation of Corollary 4.6, the minimum for δ(0,1]\delta\in(0,1] of the quantity in (4.5) is attained when δ=((12s)ςω)12sε\delta=\left(\frac{(1-2s)\varsigma}{\omega}\right)^{\frac{1}{2s}}\varepsilon and equals 2s(12sω)12s2sς12sε12s-2s\left(\frac{1-2s}{\omega}\right)^{\frac{1-2s}{2s}}\varsigma^{\frac{1}{2s}}\varepsilon^{1-2s}.

Proof.

We consider the function

[0,1)δf(δ):=ςδ12s+ωδε2s[0,1)\ni\delta\mapsto f(\delta):=-\varsigma{\delta^{1-2s}}+\frac{\omega\delta}{\varepsilon^{2s}}

and we observe that f(δ)=(12s)ςδ2s+ωε2sf^{\prime}(\delta)=-(1-2s)\varsigma{\delta^{-2s}}+\frac{\omega}{\varepsilon^{2s}} which is positive if and only if δ>((12s)ςω)12sε\delta>\left(\frac{(1-2s)\varsigma}{\omega}\right)^{\frac{1}{2s}}\varepsilon and the desired result follows. ∎

Proof of Theorem 1.9.

Use Corollary 4.7 with vε:=uδv_{\varepsilon}:=u_{\delta} for δ:=((12s)ςω)12sε\delta:=\left(\frac{(1-2s)\varsigma}{\omega}\right)^{\frac{1}{2s}}\varepsilon. ∎

Proof of Corollary 1.10.

This is a consequence of Theorem 1.9 and Lemma 4.2. ∎

5. Notation for s[12,1)s\in\left[\frac{1}{2},1\right)

Throughout the remainder of the paper, we assume that s[12,1)s\in\left[\frac{1}{2},1\right).

We use the following notation for measurable sets AA, Ω\Omega\subset\mathbb{R}:

(5.1) ε(1)(u,A)\displaystyle\mathcal{F}_{\varepsilon}^{(1)}(u,A) :=a~εu(QA)+b~εAW(u(x))𝑑x,\displaystyle:=\tilde{a}_{\varepsilon}u(Q_{A})+\tilde{b}_{\varepsilon}\int_{A}W(u(x))\,dx,
Iε(u,A,Ω)\displaystyle I_{\varepsilon}(u,A,\Omega) :=a~ε[u(AΩ,AΩ)+2u(AΩ,A(AΩ))]+b~εAΩW(u(x))𝑑x,\displaystyle:=\tilde{a}_{\varepsilon}\Big{[}u(A\cap\Omega,A\cap\Omega)+2u(A\cap\Omega,A\setminus(A\cap\Omega))\Big{]}+\tilde{b}_{\varepsilon}\int_{A\cap\Omega}W(u(x))\,dx,
𝒢s(u,A)\displaystyle\mathcal{G}_{s}(u,A) :=u(QA)+AW(u(x))𝑑x.\displaystyle:=u(Q_{A})+\int_{A}W(u(x))\,dx.

We recall that a~ε\tilde{a}_{\varepsilon} and b~ε\tilde{b}_{\varepsilon} are defined in (3.2). We point out that in contrast to ε(1)(u,A)\mathcal{F}_{\varepsilon}^{(1)}(u,A), the energy Iε(u,A,Ω)I_{\varepsilon}(u,A,\Omega) only depends on the values of uu in AA.

Also, let AA, BB\subset\mathbb{R} and assume that vv, w:w:\mathbb{R}\to\mathbb{R} are such that v=wv=w in (AB)c(A\cap B)^{c}. It is a straightforward computation to show that

(5.2) ε(1)(w,A)ε(1)(v,A)=ε(1)(w,B)ε(1)(v,B).\mathcal{F}_{\varepsilon}^{(1)}(w,A)-\mathcal{F}_{\varepsilon}^{(1)}(v,A)=\mathcal{F}_{\varepsilon}^{(1)}(w,B)-\mathcal{F}_{\varepsilon}^{(1)}(v,B).

Lastly, we note how the energies in (5.1) scale. First, if uρ(x):=u(ρx)u_{\rho}(x):=u(\rho x) for ρ>0\rho>0, then it is easy to check that

(5.3) ε(1)(uρ,A)={ρε(1)(u,ρA)ifs(12,1),|ln(ρε)||lnε|ρε(1)(u,ρA)ifs=12as long asρε1,\mathcal{F}_{\varepsilon}^{(1)}(u_{\rho},A)=\begin{cases}\displaystyle\mathcal{F}_{\rho\varepsilon}^{(1)}(u,\rho A)&\hbox{if}\leavevmode\nobreak\ s\in\left(\frac{1}{2},1\right),\\ \\ \displaystyle\frac{|\ln(\rho\varepsilon)|}{|\ln\varepsilon|}\mathcal{F}_{\rho\varepsilon}^{(1)}(u,\rho A)&\hbox{if}\leavevmode\nobreak\ s=\frac{1}{2}\quad\hbox{as long as}\leavevmode\nobreak\ \rho\not=\varepsilon^{-1},\end{cases}

and similarly for Iε(uρ,A,Ω)I_{\varepsilon}(u_{\rho},A,\Omega). In the special case uε(x):=u(x/ε)u_{\varepsilon}(x):=u(x/\varepsilon) (i.e. ρ=ε1\rho=\varepsilon^{-1}), we have

(5.4) ε(1)(uε,A)={𝒢s(u,A/ε)ifs(12,1),|lnε|1𝒢12(u,A/ε)ifs=12.\mathcal{F}_{\varepsilon}^{(1)}(u_{\varepsilon},A)=\begin{cases}\mathcal{G}_{s}(u,A/\varepsilon)&\hbox{if}\leavevmode\nobreak\ s\in\left(\frac{1}{2},1\right),\\ \displaystyle|\ln\varepsilon|^{-1}\mathcal{G}_{\frac{1}{2}}(u,A/\varepsilon)&\hbox{if}\leavevmode\nobreak\ s=\frac{1}{2}.\end{cases}

6. Heteroclinic connections

In this section, we give background and preliminaries on heteroclinic connections that connect 1-1 at -\infty to +1+1 at ++\infty. As proved in [SV-gamma], these are used to construct the recovery sequence in the interior (i.e. in compact subsets ΩΩ\Omega^{\prime}\Subset\Omega) for Γ\Gamma-convergence. We will also use the heteroclinic connections as barrier when studying connections at the boundary.

Recalling (5.1), we set

(6.1) 𝒢~s(u):={𝒢s(u,)ifs(12,1),limR+𝒢s(u,BR)lnRifs=12.\tilde{\mathcal{G}}_{s}(u):=\begin{cases}{\mathcal{G}}_{s}(u,\mathbb{R})&\hbox{if}\leavevmode\nobreak\ s\in\left(\frac{1}{2},1\right),\\ \displaystyle\lim_{R\to+\infty}\frac{{\mathcal{G}}_{s}(u,B_{R})}{\ln R}&\hbox{if}\leavevmode\nobreak\ s=\frac{1}{2}.\end{cases}

Existence and uniqueness (up to translations) of minimizers of 𝒢~s\tilde{\mathcal{G}}_{s} over the class of HsH^{s}-functions that connect 1-1 at -\infty and +1+1 at ++\infty was established in [PAL, CABSI]. After fixing the value of the minimizer at the origin, the authors also show that the unique minimizer u0:(1,1)u_{0}:\mathbb{R}\to(-1,1) is in the class C2()C^{2}(\mathbb{R}) and satisfies

(6.2) {(Δ)su0(x)+W(u0(x))=0for allx,limx±u0(x)=±1,u0(0)=0,u>0.\begin{cases}(-\Delta)^{s}u_{0}(x)+W^{\prime}(u_{0}(x))=0&\;\hbox{for all}\leavevmode\nobreak\ x\in\mathbb{R},\\ \displaystyle\lim_{x\to\pm\infty}u_{0}(x)=\pm 1,&\\ u_{0}(0)=0,&\\ u^{\prime}>0.\end{cases}

Moreover, by [PAL, Theorem 2], for any s[12,1)s\in\left[\frac{1}{2},1\right), there exist constants CC, R1R\geqslant 1 such that

(6.3) |u0(x)sgn(x)|C|x|2sand0<u0(x)C|x|1+2sfor all|x|R.|u_{0}(x)-\operatorname{sgn}(x)|\leqslant\frac{C}{|x|^{2s}}\qquad\hbox{and}\qquad 0<u_{0}^{\prime}(x)\leqslant\frac{C}{|x|^{1+2s}}\qquad\hbox{for all}\leavevmode\nobreak\ |x|\geqslant R.

6.1. Limiting behavior near s=12s=\frac{1}{2}

We will need the following results on the limiting behavior of the minimizers u0u_{0} as s12s\searrow\frac{1}{2}.

For clarity, let u0su_{0}^{s} denote the minimizer of 𝒢~s\tilde{\mathcal{G}}_{s} given in (6.2). Since |u0s|1|u_{0}^{s}|\leqslant 1, as a consequence of [CS, Theorem 26], there exist α(0,1)\alpha\in(0,1) and C>0C>0 such that, for all s[12,1)s\in\left[\frac{1}{2},1\right),

(6.4) u0sCα()C.\|u_{0}^{s}\|_{C^{\alpha}(\mathbb{R})}\leqslant C.

Fix now s0s_{0} such that

(6.5) 12<s0<α+12.\frac{1}{2}<s_{0}<\frac{\alpha+1}{2}.

We see that the asymptotic behavior of u0su_{0}^{s} in (6.3) is uniform in s[12,s0]s\in\left[\frac{1}{2},s_{0}\right].

Lemma 6.1.

There exist constants CC, R1R\geqslant 1 such that, for all s[12,s0]s\in\left[\frac{1}{2},s_{0}\right],

|u0s(x)sgn(x)|C|x|2sfor all|x|R.|u_{0}^{s}(x)-\operatorname{sgn}(x)|\leqslant\frac{C}{|x|^{2s}}\qquad\hbox{for all}\leavevmode\nobreak\ |x|\geqslant R.

The proof relies on the barrier constructed in [SV-dens]. More precisely, one can check that the following version, in which all constants are uniformly bounded in ss in compact subsets of (0,1)(0,1), holds.

Lemma 6.2 (Lemma 3.1 in [SV-dens]).

Let n1n\geqslant 1. Fix δ>0\delta>0 and 0<s1<s2<10<s_{1}<s_{2}<1.

There exists C=C(δ,n,s1,s2)1C=C(\delta,n,s_{1},s_{2})\geqslant 1 such that, for any RCR\geqslant C and all s[s1,s2]s\in[s_{1},s_{2}], there exists a rotationally symmetric function

ωsC(;[1+CR2s,1])\omega_{s}\in C(\mathbb{R};[-1+CR^{-2s},1])

with

ωs=1in(BR)c\omega_{s}=1\quad\hbox{in}\leavevmode\nobreak\ (B_{R})^{c}

such that

(Δ)sωs(x)=nωs(y)ωs(x)|xy|1+2s𝑑yδ(1+ωs(x))-(-\Delta)^{s}\omega_{s}(x)=\int_{\mathbb{R}^{n}}\frac{\omega_{s}(y)-\omega_{s}(x)}{|x-y|^{1+2s}}\,dy\leqslant\delta\big{(}1+\omega_{s}(x)\big{)}

and

(6.6) 1C(R+1|x|)2s1+ωs(x)C(R+1|x|)2s\frac{1}{C}(R+1-|x|)^{-2s}\leqslant 1+\omega_{s}(x)\leqslant C(R+1-|x|)^{-2s}

for all xBRx\in B_{R}.

Several times in the paper, we will use for b>a>0b>a>0 that, for s(12,1)s\in\left(\frac{1}{2},1\right),

(6.7) a12sb12s2s1lims12a12sb12s2s1=lnblna=lnba.\frac{a^{1-2s}-b^{1-2s}}{2s-1}\leqslant\lim_{s\searrow\frac{1}{2}}\frac{a^{1-2s}-b^{1-2s}}{2s-1}=\ln b-\ln a=\ln\frac{b}{a}.
Proof of Lemma 6.1.

We will prove that there exist CC, R1R\geqslant 1 such that, for all s[12,s0)s\in\left[\frac{1}{2},s_{0}\right),

(6.8) 0u0s(x)+1C|x|2sforx<R.0\leqslant u_{0}^{s}(x)+1\leqslant\frac{C}{|x|^{2s}}\qquad\hbox{for}\leavevmode\nobreak\ x<-R.

The uniform decay at ++\infty is proved similarly.

Fix δ>0\delta>0 small and s[12,s0]s\in\left[\frac{1}{2},s_{0}\right]. Let rs,δ>0r_{s,\delta}>0 be such that u0s(rs,δ)=1+δu_{0}^{s}(-r_{s,\delta})=-1+\delta.

We claim that

(6.9) rs,δr_{s,\delta} can be bounded from above by some rδr_{\delta}, independent of s[12,s0]s\in\left[\frac{1}{2},s_{0}\right].

In the following, CC denotes a positive constant, independent of s[12,s0]s\in\left[\frac{1}{2},s_{0}\right].

We may assume that rs,δ4r_{s,\delta}\geqslant 4, otherwise we are done. For ease in notation, we write r=rs,δr=r_{s,\delta} and consider the sets A:=[r,0]A:=[-r,0] and A:=[r+1,1]A^{\prime}:=[-r+1,-1]. Let ζC(;[0,1])\zeta\in C(\mathbb{R};[0,1]) be such that

ζ=1inAc,ζ=0inA,andζC0,1(A).\zeta=1\leavevmode\nobreak\ \hbox{in}\leavevmode\nobreak\ A^{c},\quad\zeta=0\leavevmode\nobreak\ \hbox{in}\leavevmode\nobreak\ A^{\prime},\quad\hbox{and}\quad\zeta\in C^{0,1}(A).

Note that the Lipschitz constant of ζ\zeta in AA can be taken independently of ss. Define the function

u(x):=ζ(x)u0s(x)(1ζ(x)).u_{*}(x):=\zeta(x)u_{0}^{s}(x)-(1-\zeta(x)).

Since u=u0su_{*}=u_{0}^{s} in AcA^{c} and u0su_{0}^{s} is a minimizer of 𝒢s{\mathcal{G}}_{s}, it holds that

(6.10) 0𝒢s(u0s,A)𝒢s(u,A).0\geqslant{\mathcal{G}}_{s}(u_{0}^{s},A)-{\mathcal{G}}_{s}(u_{*},A).

Now, we notice that 1+δu0s0-1+\delta\leqslant u_{0}^{s}\leqslant 0 in AA, and therefore there exists C1=C1(δ,W)>0C_{1}=C_{1}(\delta,W)>0 such that

AW(u0s(x))𝑑xinft[1+δ,0]W(t)|A|=C1rs,δ.\int_{A}W(u_{0}^{s}(x))\,dx\geqslant\inf_{t\in[-1+\delta,0]}W(t)|A|=C_{1}r_{s,\delta}.

On the other hand, since W(u(x))=W(1)=0W(u_{*}(x))=W(-1)=0 in AA^{\prime}, there exists C2=C2(δ,W)>0C_{2}=C_{2}(\delta,W)>0 such that

AW(u(x))𝑑x=AAW(u(x))𝑑x2supt[1,0]W(t)=C2.\int_{A}W(u_{*}(x))\,dx=\int_{A\setminus A^{\prime}}W(u_{*}(x))\,dx\leqslant 2\sup_{t\in[-1,0]}W(t)=C_{2}.

Therefore, from (6.10), we have that

(6.11) 0u0s(QA)u(QA)+C1rs,δC2.0\geqslant u_{0}^{s}(Q_{A})-u_{*}(Q_{A})+C_{1}r_{s,\delta}-C_{2}.

We next show that there exists C3=C3(δ)>0C_{3}=C_{3}(\delta)>0 such that

(6.12) u(QA)u0s(QA)C3(1+lnrs,δ).u_{*}(Q_{A})-u_{0}^{s}(Q_{A})\leqslant C_{3}(1+\ln r_{s,\delta}).

For this, for any xx, yy\in\mathbb{R}, we use the fact that

|u(x)u(y)|\displaystyle|u_{*}(x)-u_{*}(y)| =|ζ(x)u0s(x)ζ(y)us(y)+ζ(x)ζ(y)|\displaystyle=|\zeta(x)u_{0}^{s}(x)-\zeta(y)u^{s}(y)+\zeta(x)-\zeta(y)|
=|ζ(x)u0s(x)ζ(x)u0s(y)+ζ(x)u0s(y)ζ(y)us(y)+ζ(x)ζ(y)|\displaystyle=|\zeta(x)u_{0}^{s}(x)-\zeta(x)u_{0}^{s}(y)+\zeta(x)u_{0}^{s}(y)-\zeta(y)u^{s}(y)+\zeta(x)-\zeta(y)|
|ζ(x)||u0s(x)u0s(y)|+|u0s(y)+1||ζ(x)ζ(y)|\displaystyle\leqslant|\zeta(x)|\,|u_{0}^{s}(x)-u_{0}^{s}(y)|+|u_{0}^{s}(y)+1|\,|\zeta(x)-\zeta(y)|
|u0s(x)u0s(y)|+2|ζ(x)ζ(y)|\displaystyle\leqslant|u_{0}^{s}(x)-u_{0}^{s}(y)|+2|\zeta(x)-\zeta(y)|

to note that

|u(x)u(y)|2|u0s(x)u0s(y)|24|u0s(x)u0s(y)||ζ(x)ζ(y)|+4|ζ(x)ζ(y)|2.|u_{*}(x)-u_{*}(y)|^{2}-|u_{0}^{s}(x)-u_{0}^{s}(y)|^{2}\leqslant 4|u_{0}^{s}(x)-u_{0}^{s}(y)|\,|\zeta(x)-\zeta(y)|+4|\zeta(x)-\zeta(y)|^{2}.

From this and (6.4), and since ζ\zeta is Lipschitz in AA, for any A′′AA^{\prime\prime}\subseteq A,

AA\displaystyle\int_{A\setminus A^{\prime}} A′′{|xy|<1}|u(x)u(y)|2|xy|1+2s𝑑y𝑑xAAA′′{|xy|<1}|u0s(x)u0s(y)|2|xy|1+2s𝑑y𝑑x\displaystyle\int_{A^{\prime\prime}\cap\{|x-y|<1\}}\frac{|u_{*}(x)-u_{*}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx-\int_{A\setminus A^{\prime}}\int_{A^{\prime\prime}\cap\{|x-y|<1\}}\frac{|u_{0}^{s}(x)-u_{0}^{s}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx
4AAA′′{|xy|<1}|u0s(x)u0s(y)||ζ(x)ζ(y)|+|ζ(x)ζ(y)|2|xy|1+2s𝑑y𝑑x\displaystyle\leqslant 4\int_{A\setminus A^{\prime}}\int_{A^{\prime\prime}\cap\{|x-y|<1\}}\frac{|u_{0}^{s}(x)-u_{0}^{s}(y)|\,|\zeta(x)-\zeta(y)|+|\zeta(x)-\zeta(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx
CAA{|xy|<1}|xy|α2s𝑑y𝑑x.\displaystyle\leqslant C\int_{A\setminus A^{\prime}}\int_{\{|x-y|<1\}}|x-y|^{\alpha-2s}\,dy\,dx.

Recalling (6.5), it follows for all s[12,s0]s\in\left[\frac{1}{2},s_{0}\right] that

(6.13) AAA′′{|xy|<1}|u(x)u(y)|2|xy|1+2s𝑑y𝑑xAAA′′{|xy|<1}|u0s(x)u0s(y)|2|xy|1+2s𝑑y𝑑xCα2s0+1|AA|C.\begin{split}\int_{A\setminus A^{\prime}}&\int_{A^{\prime\prime}\cap\{|x-y|<1\}}\frac{|u_{*}(x)-u_{*}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx-\int_{A\setminus A^{\prime}}\int_{A^{\prime\prime}\cap\{|x-y|<1\}}\frac{|u_{0}^{s}(x)-u_{0}^{s}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &\leqslant\frac{C}{\alpha-2s_{0}+1}|A\setminus A^{\prime}|\leqslant C.\end{split}

Also,

(6.14) AA{|xy|>1}|u(x)u(y)|2|xy|1+2s𝑑y𝑑xCAA{|xy|>1}dxdy|xy|1+2sC2s|AA|C.\begin{split}&\int_{A\setminus A^{\prime}}\int_{\{|x-y|>1\}}\frac{|u_{*}(x)-u_{*}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\leqslant C\int_{A\setminus A^{\prime}}\int_{\{|x-y|>1\}}\frac{dx\,dy}{|x-y|^{1+2s}}\\ &\qquad\qquad\leqslant\frac{C}{2s}|A\setminus A^{\prime}|\leqslant C.\end{split}

As a consequence of (6.13) and (6.14), for any A′′AA^{\prime\prime}\subseteq A,

(6.15) u(AA,A′′)u0s(AA,A′′)C.u_{*}(A\setminus A^{\prime},A^{\prime\prime})-u_{0}^{s}(A\setminus A^{\prime},A^{\prime\prime})\leqslant C.

Moreover, since u1u^{*}\equiv-1 in AA^{\prime}, we have that

u(A,A)u0s(A,A)=0u0s(A,A)0,u_{*}(A^{\prime},A^{\prime})-u_{0}^{s}(A^{\prime},A^{\prime})=0-u_{0}^{s}(A^{\prime},A^{\prime})\leqslant 0,

and thus

u(A,A)u0s(A,A)\displaystyle u_{*}(A,A)-u_{0}^{s}(A,A)
2[u(AA,A)u0s(AA,A)]+[u(AA,AA)u0s(AA,AA)].\displaystyle\qquad\leqslant 2\big{[}u_{*}(A\setminus A^{\prime},A^{\prime})-u_{0}^{s}(A\setminus A^{\prime},A^{\prime})\big{]}+\big{[}u_{*}(A\setminus A^{\prime},A\setminus A^{\prime})-u_{0}^{s}(A\setminus A^{\prime},A\setminus A^{\prime})\big{]}.

Combining this with (6.15) (used with A′′:=AA^{\prime\prime}:=A^{\prime} and A′′:=AAA^{\prime\prime}:=A\setminus A^{\prime}),

(6.16) u(A,A)u0s(A,A)C.u_{*}(A,A)-u_{0}^{s}(A,A)\leqslant C.

Next, we check the interactions in A×AcA\times A^{c} by first writing

u(A,Ac)\displaystyle u_{*}(A,A^{c}) u0s(A,Ac)\displaystyle-u_{0}^{s}(A,A^{c})
[u(A,Ac)u0s(A,Ac)]+[u(AA,Ac)u0s(AA,Ac)].\displaystyle\leqslant\big{[}u_{*}(A^{\prime},A^{c})-u_{0}^{s}(A^{\prime},A^{c})\big{]}+\big{[}u_{*}(A\setminus A^{\prime},A^{c})-u_{0}^{s}(A\setminus A^{\prime},A^{c})\big{]}.

Since |u|1|u_{*}|\leqslant 1 and with (6.7), we estimate, for s(12,s0)s\in\left(\frac{1}{2},s_{0}\right),

u(A,Ac)u0s(A,Ac)AAc|u(x)u(y)|2|xy|1+2s𝑑y𝑑x\displaystyle u_{*}(A^{\prime},A^{c})-u_{0}^{s}(A^{\prime},A^{c})\leqslant\int_{A^{\prime}}\int_{A^{c}}\frac{|u^{*}(x)-u^{*}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx
4rs,δ+11[rs,δ(xy)12s𝑑y+0+(yx)12s𝑑y]𝑑x\displaystyle\qquad\qquad\leqslant 4\int_{-r_{s,\delta}+1}^{-1}\left[\int_{-\infty}^{-r_{s,\delta}}(x-y)^{-1-2s}\,dy+\int_{0}^{+\infty}(y-x)^{-1-2s}\,dy\right]\,dx
=4s(2s1)[1(rs,δ1)12s]Clnrs,δ.\displaystyle\qquad\qquad=\frac{4}{s(2s-1)}\big{[}1-(r_{s,\delta}-1)^{1-2s}\big{]}\leqslant C\ln r_{s,\delta}.

Taking s=12s=\frac{1}{2} gives the same estimate.

On the other hand, we use (6.14) and estimate as in (6.13) for all s[12,s0]s\in\left[\frac{1}{2},s_{0}\right] to find

u(AA,Ac)u0s(AA,Ac)\displaystyle u_{*}(A\setminus A^{\prime},A^{c})-u_{0}^{s}(A\setminus A^{\prime},A^{c})
AA{|xy|<1}|u(x)u(y)|2|u0s(x)u0s(y)|2|xy|1+2s𝑑y𝑑x+CC.\displaystyle\qquad\leqslant\int_{A\setminus A^{\prime}}\int_{\{|x-y|<1\}}\frac{|u_{*}(x)-u_{*}(y)|^{2}-|u_{0}^{s}(x)-u_{0}^{s}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx+C\leqslant C.

Combining the previous two displays gives

(6.17) u(A,Ac)u0s(A,Ac)C(1+lnrs,δ).u_{*}(A,A^{c})-u_{0}^{s}(A,A^{c})\leqslant C(1+\ln r_{s,\delta}).

Thus, the estimate in (6.12) follows from (6.16) and (6.17).

By (6.11) and (6.12), there exists C>0C>0 such that, for all s[12,s0]s\in\left[\frac{1}{2},s_{0}\right],

rs,δC(1+lnrs,δ).r_{s,\delta}\leqslant C(1+\ln r_{s,\delta}).

Consequently, rs,δr_{s,\delta} is bounded uniformly in s[12,s0]s\in\left[\frac{1}{2},s_{0}\right], thus giving the claim in (6.9).

We are now ready to show (6.8). The proof follows the lines of [PAL, Proposition 3] except that we have to carefully track the dependence on s[12,s0]s\in\left[\frac{1}{2},s_{0}\right]. For the sake of the reader, we sketch the idea.

First, note that there exists some δ>0\delta>0 such that

W(t2)W(t1)+δ(t2t1)for 1t1t21+δ.W^{\prime}(t_{2})\geqslant W^{\prime}(t_{1})+\delta(t_{2}-t_{1})\qquad\hbox{for }-1\leqslant t_{1}\leqslant t_{2}\leqslant-1+\delta.

For this particular δ>0\delta>0 and for s[12,s0]s\in\left[\frac{1}{2},s_{0}\right], let ωs\omega_{s} be the barrier in Lemma 6.2 and take RCR\geqslant C. As in the proof of [PAL, Proposition 3] (note that all the constants in [PAL, Corollary 4 and Lemma 9] can be made uniform in ss in compact subsets [s1,s2](0,1)[s_{1},s_{2}]\subset(0,1)), there exist k¯\bar{k}, x¯\bar{x}\in\mathbb{R} (possibly depending on ss) such that

ωs(x¯k¯)=u0s(x¯)>1+δandωs(xk¯)u0s(x) for all x.\omega_{s}(\bar{x}-\bar{k})=u_{0}^{s}(\bar{x})>-1+\delta\qquad\hbox{and}\qquad\omega_{s}(x-\bar{k})\geqslant u_{0}^{s}(x)\;\hbox{ for all }x\in\mathbb{R}.

Moreover, one can show that there exists C>0C^{\prime}>0 such that, for all s[12,s0]s\in\left[\frac{1}{2},s_{0}\right],

x¯k¯[RC,R].\bar{x}-\bar{k}\in[R-C^{\prime},R].

Recall from above that rs,δr_{s,\delta} is such that u0s(rs,δ)=1+δu_{0}^{s}(-r_{s,\delta})=-1+\delta and that there exists some rδr_{\delta} such that rδrs,δr_{\delta}\geqslant r_{s,\delta} for all s[12,s0]s\in\left[\frac{1}{2},s_{0}\right]. Since u0su_{0}^{s} is strictly increasing, it must be that rδx¯-r_{\delta}\leqslant\bar{x} and

u0s(xrδ)u0s(x+x¯)for allx.u_{0}^{s}(x-r_{\delta})\leqslant u_{0}^{s}(x+\bar{x})\quad\hbox{for all}\leavevmode\nobreak\ x\in\mathbb{R}.

Now take y[R2,R]y\in\left[\frac{R}{2},R\right]. As in [PAL], one can check that

x¯yk¯[R2,R2]for alls[12,s0],\bar{x}-y-\bar{k}\in\left[-\frac{R}{2},\frac{R}{2}\right]\quad\hbox{for all}\leavevmode\nobreak\ s\in\left[\frac{1}{2},s_{0}\right],

so, with (6.6), we have that

1+ωs(x¯yk¯)C(R/2)2s4Cy2s.1+\omega_{s}(\bar{x}-y-\bar{k})\leqslant C(R/2)^{-2s}\leqslant 4Cy^{-2s}.

Hence, it holds that

u0s(rδy)u0s(x¯y)ωs(x¯yk¯)1+4Cy2s.u_{0}^{s}(-r_{\delta}-y)\leqslant u_{0}^{s}(\bar{x}-y)\leqslant\omega_{s}(\bar{x}-y-\bar{k})\leqslant-1+4Cy^{-2s}.

Since rδr_{\delta} is a constant and RR can be made arbitrarily large, (6.8) holds. ∎

Corollary 6.3.

There exists a subsequence u0sku_{0}^{s_{k}} such that

limsk12u0sk=u012locally uniformly in.\lim_{s_{k}\searrow\frac{1}{2}}u_{0}^{s_{k}}=u_{0}^{\frac{1}{2}}\quad\hbox{locally uniformly in}\leavevmode\nobreak\ \mathbb{R}.
Proof.

By (6.4), there exist a subsequence and a measurable function vv such that

lims12u0s(x)=v(x)locally uniformly in.\lim_{s\searrow\frac{1}{2}}u_{0}^{s}(x)=v(x)\quad\quad\hbox{locally uniformly in}\leavevmode\nobreak\ \mathbb{R}.

Note that v:[1,1]v:\mathbb{R}\to[-1,1] is non-decreasing, v(0)=0v(0)=0, and vv solves

(Δ)12v(x)+W(v(x))=0for allx.(-\Delta)^{\frac{1}{2}}v(x)+W^{\prime}(v(x))=0\quad\hbox{for all}\leavevmode\nobreak\ x\in\mathbb{R}.

We claim that v=u012v=u_{0}^{\frac{1}{2}}. Indeed, by Lemma 6.1, there exist CC, R1R\geqslant 1 such that, for all |x|R|x|\geqslant R,

|v(x)sgn(x)|=lims12|u0s(x)sgn(x)|lims12C|x|2s=C|x|.|v(x)-\operatorname{sgn}(x)|=\lim_{s\searrow\frac{1}{2}}|u_{0}^{s}(x)-\operatorname{sgn}(x)|\leqslant\lim_{s\searrow\frac{1}{2}}\frac{C}{|x|^{2s}}=\frac{C}{|x|}.

In particular,

limx±v(x)=±1.\lim_{x\to\pm\infty}v(x)=\pm 1.

We now have that vv satisfies (6.2) for s=12s=\frac{1}{2}. By uniqueness (see [PAL]), we then have that v=u012v=u_{0}^{\frac{1}{2}}, as desired. ∎

7. Connections at the boundary and proof of Theorem 1.3

In this section, we construct the penalization function Ψ\Psi in (1.9) and prove Theorem 1.3.

We recall the setting of XγX_{\gamma} stated on page 1.2 and define the functional

(7.1) 𝒢s(u):={𝒢s(u,)ifs(12,1),lim infR+(𝒢s(u,BR)lnR)ifs=12,\mathcal{G}_{s}(u):=\begin{cases}\mathcal{G}_{s}(u,\mathbb{R}^{-})&\hbox{if}\leavevmode\nobreak\ s\in\left(\frac{1}{2},1\right),\\ \displaystyle\liminf_{R\to+\infty}\left(\frac{\mathcal{G}_{s}(u,B_{R}^{-})}{\ln R}\right)&\hbox{if}\leavevmode\nobreak\ s=\frac{1}{2},\end{cases}

where 𝒢s(u,)\mathcal{G}_{s}(u,\cdot) is given in (5.1). Note the contrast with (6.1).

Let w0(x;+1,γ)w_{0}(x;+1,\gamma) denote the minimizer of 𝒢s\mathcal{G}_{s} in the class XγX^{\gamma} (whose definition is obtained replacing 1-1 with 11 in the definition of XγX_{\gamma}) and

(7.2) w0(x;1,γ)w_{0}(x;-1,\gamma) denote the minimizer in the class XγX_{\gamma} described in Theorem 1.3.

The penalization function Ψ:{±1}×(1,1)(0,+)\Psi:\{\pm 1\}\times(-1,1)\to(0,+\infty) is then defined as

(7.3) Ψ(±1,γ):=𝒢s(w0(;±1,γ)).\Psi(\pm 1,\gamma):=\mathcal{G}_{s}(w_{0}(\cdot;\pm 1,\gamma)).

We are left to prove Theorem 1.3 for w0(x):=w0(x;1,γ)w_{0}(x):=w_{0}(x;-1,\gamma) for a fixed γ(1,1)\gamma\in(-1,1). The proof is split into several lemmata. We first establish properties of non-decreasing solutions to the PDE in (1.10). Then, we prove existence of minimizers for s(12,1)s\in\left(\frac{1}{2},1\right) and show that they satisfy (1.10). By sending s12s\searrow\frac{1}{2}, we prove existence of minimizers for s=12s=\frac{1}{2}. Next, we collectively show asymptotic behavior at -\infty for all s[12,1)s\in\left[\frac{1}{2},1\right) and finally prove uniqueness.

For clarity, let us state the definition of local and global minimizers in our setting.

Definition 7.1.

We say that a function ww is a local minimizer of 𝒢s\mathcal{G}_{s} in BRB_{R}^{-} if w=γw=\gamma in +\mathbb{R}^{+} and

𝒢s(w,BR)𝒢s(w+ϕ,BR)for any measurableϕsuch thatsuppϕBRandw+ϕγ.\mathcal{G}_{s}(w,B_{R}^{-})\leqslant\mathcal{G}_{s}(w+\phi,B_{R}^{-})\qquad\begin{subarray}{c}\hbox{for any measurable}\leavevmode\nobreak\ \displaystyle\phi\leavevmode\nobreak\ \hbox{such that}\\ \displaystyle\operatorname{supp}\phi\subset B_{R}^{-}\leavevmode\nobreak\ \hbox{and}\leavevmode\nobreak\ w+\phi\leqslant\gamma.\end{subarray}

We say that ww is a global minimizer of 𝒢s\mathcal{G}_{s} if ww is a local minimizer of 𝒢s\mathcal{G}_{s} in BRB_{R}^{-} for all R>0R>0 and if 𝒢s(w)<+\mathcal{G}_{s}(w)<+\infty.

7.1. Properties of solutions

For later reference, we start with the following properties of solutions to the PDE in (1.10).

Lemma 7.2.

Let s[12,1)s\in\left[\frac{1}{2},1\right) and γ(1,1)\gamma\in(-1,1). Let w0w_{0} be a non-decreasing solution to (1.10) such that 1w0(x)<γ-1\leqslant w_{0}(x)<\gamma for all xx\in\mathbb{R}^{-}.

It holds that w0Cs()Clocα()w_{0}\in C^{s}(\mathbb{R})\cap C^{\alpha}_{\text{loc}}(\mathbb{R}^{-}) for any α(0,1)\alpha\in(0,1), w0w_{0} is strictly increasing in \mathbb{R}^{-}, and w0>1w_{0}>-1.

If s(12,1)s\in\left(\frac{1}{2},1\right), then additionally w0Cloc2s()w_{0}\in C^{2s}_{\text{loc}}(\mathbb{R}^{-}).

Proof.

The interior regularity w0Clocα()w_{0}\in C^{\alpha}_{\text{loc}}(\mathbb{R}^{-}) (and Cloc2s()C^{2s}_{\text{loc}}(\mathbb{R}^{-}) for s12s\not=\frac{1}{2}) follows from [ROSerra, Theorem 1.1] and the global regularity w0Cs()w_{0}\in C^{s}(\mathbb{R}) follows from [RosOtonBook, Proposition 2.6.15].

Now, we show that w0>1w_{0}>-1. For this, suppose by contradiction that there exist a point x0x_{0} such that w0(x0)=1w_{0}(x_{0})=-1. Since γ>1\gamma>-1, it must be that x0<0x_{0}<0, so we have

0=W(w0(x0))=(Δ)sw0(x0)=w0(y)+1|x0y|1+2s𝑑y.0=W^{\prime}(w_{0}(x_{0}))=-(-\Delta)^{s}w_{0}(x_{0})=\int_{\mathbb{R}}\frac{w_{0}(y)+1}{|x_{0}-y|^{1+2s}}\,dy.

Since w0+10w_{0}+1\geqslant 0, it follows that w01w_{0}\equiv-1 in \mathbb{R}, a contradiction.

Lastly, we show that w0w_{0} is strictly increasing in \mathbb{R}^{-}. Suppose, by way of contradiction, that there exist a<b<0a<b<0 such that w0(a)=w0(b)w_{0}(a)=w_{0}(b). Observe that

(Δ)sw0(a)(Δ)sw0(b)=W(w0(b))W(w0(a))=0.(-\Delta)^{s}w_{0}(a)-(-\Delta)^{s}w_{0}(b)=W^{\prime}(w_{0}(b))-W^{\prime}(w_{0}(a))=0.

Therefore,

0=w0(a)w0(a+y)|y|1+2s𝑑yw0(b)w0(b+y)|y|1+2s𝑑y=w0(b+y)w0(a+y)|y|1+2s𝑑y.0=\int_{\mathbb{R}}\frac{w_{0}(a)-w_{0}(a+y)}{|y|^{1+2s}}\,dy-\int_{\mathbb{R}}\frac{w_{0}(b)-w_{0}(b+y)}{|y|^{1+2s}}\,dy=\int_{\mathbb{R}}\frac{w_{0}(b+y)-w_{0}(a+y)}{|y|^{1+2s}}\,dy.

Since w0w_{0} is non-decreasing and b>ab>a, the integrand is nonnegative. Consequently, w0(b+y)=w0(a+y)w_{0}(b+y)=w_{0}(a+y) for all yy\in\mathbb{R}, which is false (for example take y=by=-b). Therefore, w0w_{0} is strictly increasing in \mathbb{R}^{-}. ∎

7.2. Existence of solutions for s(12,1)s\in\left(\frac{1}{2},1\right)

Here, we prove existence of global minimizers w0w_{0} in Theorem 1.3 for s(12,1)s\in\left(\frac{1}{2},1\right) and then show that, after translating, the heteroclinic function u0u_{0} in Section 6 is above w0w_{0} and touches w0w_{0} at the origin (and therefore it can be used as a barrier).

Proposition 7.3.

Let s(12,1)s\in\left(\frac{1}{2},1\right) and γ(1,1)\gamma\in(-1,1).

There exists a global minimizer w0Xγw_{0}\in X_{\gamma} of 𝒢s\mathcal{G}_{s} such that w0Cs()Cloc2s()w_{0}\in C^{s}(\mathbb{R})\cap C^{2s}_{\text{loc}}(\mathbb{R}^{-}), w0w_{0} is strictly increasing in \mathbb{R}^{-}, 1<w0(x)<γ-1<w_{0}(x)<\gamma for all xx\in\mathbb{R}^{-}, and w0w_{0} solves (1.10).

Proof.

By Lemma A.2, there exists a function hXγh\in X_{\gamma} such that 𝒢s(h)<C(1+12s1)\mathcal{G}_{s}(h)<C\left(1+\frac{1}{2s-1}\right) for some C>0C>0. Consequently,

(7.4) 0infwXγ𝒢s(w)<+.0\leqslant\inf_{w\in X_{\gamma}}\mathcal{G}_{s}(w)<+\infty.

Consider the set of functions

Xγ:={uXγs.t.1wγ,𝒢s(w)<+,andwis non-decreasing}.X_{\gamma}^{\star}:=\left\{u\in X_{\gamma}\leavevmode\nobreak\ \hbox{s.t.}\leavevmode\nobreak\ \begin{subarray}{c}\displaystyle-1\leqslant w\leqslant\gamma,\leavevmode\nobreak\ \mathcal{G}_{s}(w)<+\infty,\leavevmode\nobreak\ \hbox{and}\\ \displaystyle w\leavevmode\nobreak\ \hbox{is non-decreasing}\end{subarray}\right\}.

We claim that

(7.5) infwXγ𝒢s(w)=infwXγ𝒢s(w).\inf_{w\in X_{\gamma}}\mathcal{G}_{s}(w)=\inf_{w\in X_{\gamma}^{\star}}\mathcal{G}_{s}(w).

Since XγXγX_{\gamma}^{\star}\subset X_{\gamma}, we have that infwXγ𝒢s(w)infwXγ𝒢s(w)\inf_{w\in X_{\gamma}}\mathcal{G}_{s}(w)\leqslant\inf_{w\in X_{\gamma}^{\star}}\mathcal{G}_{s}(w). For the reverse inequality, first note that 𝒢s\mathcal{G}_{s} is decreasing under truncations at 1-1, so it is not restrictive to minimize among all functions wXγw\in X_{\gamma} satisfying 1wγ-1\leqslant w\leqslant\gamma such that 𝒢s(w)<+\mathcal{G}_{s}(w)<+\infty.

It remains to check that

(7.6) 𝒢s\mathcal{G}_{s} is decreasing under monotone rearrangements.

This is a consequence of [Alberti, Theorem 2.11] (see also [AlmgrenLiem, Theorem 9.2] for symmetric decreasing rearrangements). Indeed, consider the superlevel sets of ww

A(w):={x s.t. w(x)}, with (1,γ].A_{\ell}(w):=\{x\in\mathbb{R}\;{\mbox{ s.t. }}\;w(x)\geqslant\ell\},\quad{\mbox{ with }}\ell\in(-1,\gamma].

We claim that

(7.7) for wXγ satisfying 𝒢s(w)<+the symmetric difference A(w)+ has finite measure for all (1,γ].\begin{split}&{\mbox{for\leavevmode\nobreak\ $w\in X_{\gamma}$ satisfying\leavevmode\nobreak\ $\mathcal{G}_{s}(w)<+\infty$}}\\ &{\mbox{the symmetric difference\leavevmode\nobreak\ $A_{\ell}(w)\triangle\mathbb{R}^{+}$ has finite measure for all\leavevmode\nobreak\ $\ell\in(-1,\gamma]$.}}\end{split}

For this, note that

A(w)+=(+A(w))(A(w)+)=A(w).A_{\ell}(w)\triangle\mathbb{R}^{+}=(\mathbb{R}^{+}\setminus A_{\ell}(w))\cup(A_{\ell}(w)\setminus\mathbb{R}^{+})=A_{\ell}(w)\cap\mathbb{R}^{-}.

If |A(w)|=+|A_{\ell}(w)\cap\mathbb{R}^{-}|=+\infty, then

𝒢s(w)A(w)W(w(x))𝑑xinfr[,γ]W(r)|A(w)|=+\mathcal{G}_{s}(w)\geqslant\int_{A_{\ell}(w)\cap\mathbb{R}^{-}}W(w(x))\,dx\geqslant\inf_{r\in[\ell,\gamma]}W(r)\,|A_{\ell}(w)\cap\mathbb{R}^{-}|=+\infty

which gives a contradiction. This establishes (7.7).

We can now apply [Alberti, Theorem 2.11] to find that the increasing rearrangement ww^{*} of ww satisfies 𝒢s(w)𝒢s(w)\mathcal{G}_{s}(w^{*})\leqslant\mathcal{G}_{s}(w), and therefore (7.6) holds true.

Consequently, infwXγ𝒢s(w)infwXγ𝒢s(w)\inf_{w\in X_{\gamma}^{\star}}\mathcal{G}_{s}(w)\leqslant\inf_{w\in X_{\gamma}}\mathcal{G}_{s}(w), and so the proof of (7.5) is complete.

In light of (7.5), we now let {wk}Xγ\{w_{k}\}\subset X_{\gamma}^{\star} be a minimizing sequence for 𝒢s\mathcal{G}_{s}, namely

limk+𝒢s(wk)=infwXγ𝒢s(w).\lim_{k\to+\infty}\mathcal{G}_{s}(w_{k})=\inf_{w\in X_{\gamma}}\mathcal{G}_{s}(w).

As a consequence of [Hitchhikers, Theorem 7.1], up to a subsequence, there exists a measurable function w0w_{0} such that wkw0w_{k}\to w_{0} almost everywhere as k+k\to+\infty. By construction, w0=γw_{0}=\gamma in +\mathbb{R}^{+}, 1w0γ-1\leqslant w_{0}\leqslant\gamma, and w0w_{0} is non-decreasing. Also, by Fatou’s Lemma,

(7.8) 𝒢s(w0)lim infk+𝒢s(wk)=infwXγ𝒢s(w)<+,\mathcal{G}_{s}(w_{0})\leqslant\liminf_{k\to+\infty}\mathcal{G}_{s}(w_{k})=\inf_{w\in X_{\gamma}}\mathcal{G}_{s}(w)<+\infty,

which shows that w0w_{0} is a (global) minimizer of 𝒢s\mathcal{G}_{s}.

Furthermore, since w0w_{0} is non-decreasing and bounded, the limit

limxw0(x)=:a[1,γ]\lim_{x\to-\infty}w_{0}(x)=:a\in[-1,\gamma]

exists. If a1a\not=-1, then

W(w0)𝑑x=+\int_{\mathbb{R}^{-}}W(w_{0})\,dx=+\infty

which contradicts (7.8). Therefore, a=1a=-1 and we have that w0Xγw_{0}\in X_{\gamma}.

Moreover, by Lemma A.1, w0<γw_{0}<\gamma in \mathbb{R}^{-}. With this and the minimizing property, w0w_{0} satisfies (1.10). The remaining properties follow from Lemma 7.2. ∎

We will now use the solution u0u_{0} to (6.2) as a barrier for w0w_{0}. Towards this end, let xγx_{\gamma}\in\mathbb{R} be the unique point at which u0(xγ)=γu_{0}(x_{\gamma})=\gamma and set

uγ(x):=u0(x+xγ).u_{\gamma}(x):=u_{0}(x+x_{\gamma}).

Via the sliding method (see [PAL]), we show the following.

Lemma 7.4.

Let s(12,1)s\in\left(\frac{1}{2},1\right) and γ(1,1)\gamma\in(-1,1).

It holds that

(7.9) uγ(x)w0(x) for all xanduγ(0)=w0(0)=γ.u_{\gamma}(x)\geqslant w_{0}(x)\;\hbox{ for all }x\in\mathbb{R}\qquad\hbox{and}\qquad u_{\gamma}(0)=w_{0}(0)=\gamma.
Proof.

First, since 1uγ,w0γ-1\leqslant u_{\gamma},w_{0}\leqslant\gamma in \mathbb{R}^{-} and γ=w0uγ1\gamma=w_{0}\leqslant u_{\gamma}\leqslant 1 in +\mathbb{R}^{+}, we have that, for all ε>0\varepsilon>0, there exists kεk_{\varepsilon}\in\mathbb{R} such that

uγ(xk)+ε>w0(x)for allkkεand allx.u_{\gamma}(x-k)+\varepsilon>w_{0}(x)\quad\hbox{for all}\leavevmode\nobreak\ k\geqslant k_{\varepsilon}\leavevmode\nobreak\ \hbox{and all}\leavevmode\nobreak\ x\in\mathbb{R}.

Take kεk_{\varepsilon} as large as possible so that

(7.10) uγ(xkε)+εw0(x)for allxu_{\gamma}(x-k_{\varepsilon})+\varepsilon\geqslant w_{0}(x)\quad\hbox{for all}\leavevmode\nobreak\ x\in\mathbb{R}

and, for all jj\in\mathbb{N}, there exist ηj,ε>0\eta_{j,\varepsilon}>0 and xj,εx_{j,\varepsilon}\in\mathbb{R} such that limj+ηj,ε=0\lim_{j\to+\infty}\eta_{j,\varepsilon}=0 and

(7.11) uγ(xj,ε(kε+ηj,ε))+ε<w0(xj,ε).u_{\gamma}(x_{j,\varepsilon}-(k_{\varepsilon}+\eta_{j,\varepsilon}))+\varepsilon<w_{0}(x_{j,\varepsilon}).

We claim that

(7.12) xj,εx_{j,\varepsilon} is a bounded sequence in jj.

Indeed, if there existed a subsequence such that xj,ε+x_{j,\varepsilon}\to+\infty, then from (7.11) and the continuity of w0w_{0} and uγu_{\gamma},

1+ε=limj+uγ(xj,ε(kε+ηj,ε))+εlimj+w0(xj,ε)=γ,1+\varepsilon=\lim_{j\to+\infty}u_{\gamma}(x_{j,\varepsilon}-(k_{\varepsilon}+\eta_{j,\varepsilon}))+\varepsilon\leqslant\lim_{j\to+\infty}w_{0}(x_{j,\varepsilon})=\gamma,

a contradiction. If instead there existed a subsequence such that xj,εx_{j,\varepsilon}\to-\infty, then, similarly,

1+ε=limj+uγ(xj,ε(kε+ηj,ε))+εlimj+w0(xj,ε)=1,-1+\varepsilon=\lim_{j\to+\infty}u_{\gamma}(x_{j,\varepsilon}-(k_{\varepsilon}+\eta_{j,\varepsilon}))+\varepsilon\leqslant\lim_{j\to+\infty}w_{0}(x_{j,\varepsilon})=-1,

a contradiction. Hence, (7.12) is proved.

Thus, thanks to (7.12), up to a subsequence, there exists a point xεx_{\varepsilon}\in\mathbb{R} such that

limj+xj,ε=xε.\lim_{j\to+\infty}x_{j,\varepsilon}=x_{\varepsilon}.

Therefore, taking the limit in jj in (7.11) and recalling (7.10), we have that

(7.13) uγ(xεkε)+ε=w0(xε).u_{\gamma}(x_{\varepsilon}-k_{\varepsilon})+\varepsilon=w_{0}(x_{\varepsilon}).

Now, if xε>0x_{\varepsilon}>0 for some ε>0\varepsilon>0, by the strict monotonicity of uγu_{\gamma}, (7.10) and (7.13),

uγ(0kε)+ε<uγ(xεkε)+ε=w0(xε)=w0(0)uγ(0kε)+ε,u_{\gamma}(0-k_{\varepsilon})+\varepsilon<u_{\gamma}(x_{\varepsilon}-k_{\varepsilon})+\varepsilon=w_{0}(x_{\varepsilon})=w_{0}(0)\leqslant u_{\gamma}(0-k_{\varepsilon})+\varepsilon,

a contradiction. Therefore, xε0x_{\varepsilon}\leqslant 0 for all ε>0\varepsilon>0.

Now, suppose that there exists a subsequence such that xε=0x_{\varepsilon}=0 for all ε\varepsilon. By (7.13) and since w0(xε)=γw_{0}(x_{\varepsilon})=\gamma, we have that, for ε>0\varepsilon>0 sufficiently small,

uγ(kε)=γε(1,1).u_{\gamma}(-k_{\varepsilon})=\gamma-\varepsilon\in(-1,1).

Since uγ1:(1,1)u_{\gamma}^{-1}:(-1,1)\to\mathbb{R} exists and is continuous, we have that

limε0kε=limε0uγ1(γε)=uγ1(γ)=0.\lim_{\varepsilon\searrow 0}k_{\varepsilon}=-\lim_{\varepsilon\searrow 0}u_{\gamma}^{-1}(\gamma-\varepsilon)=-u_{\gamma}^{-1}(\gamma)=0.

Therefore, taking the limit in (7.10) and (7.13) as ε0\varepsilon\searrow 0 gives (7.9) in this case.

Consider now the scenario in which there exists ε0>0\varepsilon_{0}>0 such that xε<0x_{\varepsilon}<0 for all ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}). Define the function

uγε(x):=uγ(xkε)+ε.u_{\gamma}^{\varepsilon}(x):=u_{\gamma}(x-k_{\varepsilon})+\varepsilon.

Note that, in light of (7.10) and (7.13), uγε(x)w0(x)u_{\gamma}^{\varepsilon}(x)\geqslant w_{0}(x) for all xx\in\mathbb{R} and uγε(xε)=w0(xε)u_{\gamma}^{\varepsilon}(x_{\varepsilon})=w_{0}(x_{\varepsilon}). Also, for all xx\in\mathbb{R},

(Δ)suγε(x)=(Δ)suγ(xkε)=W(uγ(xkε))=W(uγε(x)ε).\displaystyle-(-\Delta)^{s}u_{\gamma}^{\varepsilon}(x)=-(-\Delta)^{s}u_{\gamma}(x-k_{\varepsilon})=W^{\prime}(u_{\gamma}(x-k_{\varepsilon}))=W^{\prime}(u_{\gamma}^{\varepsilon}(x)-\varepsilon).

Consequently, since we have the equation for w0w_{0} at xε<0x_{\varepsilon}<0,

(7.14) 0(uγεw0)(y)|xεy|1+2s𝑑y=(Δ)s(uγεw0)(xε)=W(uγε(xε)ε)W(w0(xε))=W(w0(xε)ε)W(w0(xε)).\begin{split}&0\leqslant\int_{\mathbb{R}}\frac{(u_{\gamma}^{\varepsilon}-w_{0})(y)}{|x_{\varepsilon}-y|^{1+2s}}\,dy=-(-\Delta)^{s}(u_{\gamma}^{\varepsilon}-w_{0})(x_{\varepsilon})\\ &\qquad=W^{\prime}(u_{\gamma}^{\varepsilon}(x_{\varepsilon})-\varepsilon)-W^{\prime}(w_{0}(x_{\varepsilon}))=W^{\prime}(w_{0}(x_{\varepsilon})-\varepsilon)-W^{\prime}(w_{0}(x_{\varepsilon})).\end{split}

We point out that

(7.15) xεx_{\varepsilon} is also bounded from below.

Indeed, if there existed a subsequence such that xεx_{\varepsilon}\to-\infty, then

(7.16) limxw0(xε)=1.\lim_{x\to-\infty}w_{0}(x_{\varepsilon})=-1.

Since W′′(1)>0W^{\prime\prime}(-1)>0, there exists δ>0\delta>0 such that

(7.17) W(t)W(r)+δ(tr)for all1rt1+δ.W^{\prime}(t)\geqslant W^{\prime}(r)+\delta(t-r)\qquad\hbox{for all}\leavevmode\nobreak\ -1\leqslant r\leqslant t\leqslant-1+\delta.

By (7.16), there exists ε1>0\varepsilon_{1}>0 such that, for all ε(0,ε1)\varepsilon\in(0,\varepsilon_{1}), it holds that 1w0(xε)ε<w0(xε)1+δ-1\leqslant w_{0}(x_{\varepsilon})-\varepsilon<w_{0}(x_{\varepsilon})\leqslant-1+\delta. Therefore, we can employ (7.17) and find that

(7.18) W(w0(xε))W(w0(xε)ε)+δε>W(w0(xε)ε),W^{\prime}(w_{0}(x_{\varepsilon}))\geqslant W^{\prime}(w_{0}(x_{\varepsilon})-\varepsilon)+\delta\varepsilon>W^{\prime}(w_{0}(x_{\varepsilon})-\varepsilon),

contradicting (7.14) and thus proving (7.15).

Thanks to (7.15), up to a subsequence, there exists a point x00x_{0}\leqslant 0 such that

limε0xε=x0.\lim_{\varepsilon\searrow 0}x_{\varepsilon}=x_{0}.

We claim that kεk_{\varepsilon} is a bounded sequence in ε\varepsilon. Indeed, if there existed a subsequence such that kεk_{\varepsilon}\to-\infty, then from (7.13)

1=limε0uγ(xεkε)+ε=limε0w0(xε)=w0(x0)γ,1=\lim_{\varepsilon\searrow 0}u_{\gamma}(x_{\varepsilon}-k_{\varepsilon})+\varepsilon=\lim_{\varepsilon\searrow 0}w_{0}(x_{\varepsilon})=w_{0}(x_{0})\leqslant\gamma,

a contradiction. If instead there existed a subsequence such that kε+k_{\varepsilon}\to+\infty, then

1=limε0uγ(xεkε)+ε=limε0w0(xε)=w0(x0)>1,-1=\lim_{\varepsilon\searrow 0}u_{\gamma}(x_{\varepsilon}-k_{\varepsilon})+\varepsilon=\lim_{\varepsilon\searrow 0}w_{0}(x_{\varepsilon})=w_{0}(x_{0})>-1,

a contradiction.

Therefore kεk_{\varepsilon} is bounded, and so, up to a subsequence, there exists k0k_{0}\in\mathbb{R} such that

limε0kε=k0.\lim_{\varepsilon\searrow 0}k_{\varepsilon}=k_{0}.

We thus send ε0\varepsilon\searrow 0 in (7.10) and (7.13) to obtain

(7.19) uγ(xk0)w0(x) for all xanduγ(x0k0)=w0(x0).u_{\gamma}(x-k_{0})\geqslant w_{0}(x)\;\hbox{ for all }x\in\mathbb{R}\qquad\hbox{and}\qquad u_{\gamma}(x_{0}-k_{0})=w_{0}(x_{0}).

Accordingly, the desired result in (7.9) is proved if we show that

(7.20) x0=k0=0.x_{0}=k_{0}=0.

Hence, we now focus on the proof of (7.20).

Suppose by contradiction that x0<0x_{0}<0. Sending ε0\varepsilon\searrow 0 in (7.14) gives

uγ(yk0)w0(y)|x0y|1+2s𝑑y=0.\int_{\mathbb{R}}\frac{u_{\gamma}(y-k_{0})-w_{0}(y)}{|x_{0}-y|^{1+2s}}\,dy=0.

From (7.19), we know that the integrand is nonnegative. Therefore, it must be that uγ(xk0)=w0(x)u_{\gamma}(x-k_{0})=w_{0}(x) for all xx\in\mathbb{R}, which is false for x>max{0,k0}x>\max\{0,k_{0}\}. Therefore, we have that x0=0x_{0}=0.

Moreover, since uγu_{\gamma} is invertible and uγ(k0)=w0(0)=γu_{\gamma}(-k_{0})=w_{0}(0)=\gamma, we have that k0=uγ1(γ)=0-k_{0}=u_{\gamma}^{-1}(\gamma)=0, which completes the proof of (7.20), as desired. ∎

7.3. Existence of solutions for s=12s=\frac{1}{2}

Now, we establish existence of global minimizers in Theorem 1.3 for s=12s=\frac{1}{2}. For this, we will show that a subsequence in s(12,1)s\in\left(\frac{1}{2},1\right) of the minimizers given by Proposition 7.3 converges as s12s\searrow\frac{1}{2} to a minimizer for s=12s=\frac{1}{2}.

For s(12,1)s\in\left(\frac{1}{2},1\right), let wsw_{s} denote the global minimizer of 𝒢s\mathcal{G}_{s} from Proposition 7.3.

Lemma 7.5.

Let γ(1,1)\gamma\in(-1,1). There exist α(0,1)\alpha\in(0,1) and C>0C>0 such that, for all s(12,1)s\in\left(\frac{1}{2},1\right),

(7.21) wsCα()C.\|w_{s}\|_{C^{\alpha}(\mathbb{R})}\leqslant C.

Furthermore, there exists a subsequence wskw_{s_{k}} that converges locally uniformly as sk12s_{k}\searrow\frac{1}{2} to a non-decreasing function w0:[1,γ]w_{0}:\mathbb{R}\to[-1,\gamma] such that w0γw_{0}\equiv\gamma in \mathbb{R}^{-} and (7.9) holds for s=12s=\frac{1}{2}.

Proof.

Since 1wsγ-1\leqslant w_{s}\leqslant\gamma, to prove (7.21), we only need to check the uniform bound on the CαC^{\alpha}-seminorm. By Proposition 7.3 and [ServadeiValdinoci, Theorem 1], wsHs()w_{s}\in H^{s}(\mathbb{R}) is also a viscosity solution to

{(Δ)sws+W(ws)=0in,ws=γin+,1<ws<γin.\begin{cases}(-\Delta)^{s}w_{s}+W^{\prime}(w_{s})=0&\hbox{in}\leavevmode\nobreak\ \mathbb{R}^{-},\\ w_{s}=\gamma&\hbox{in}\leavevmode\nobreak\ \mathbb{R}^{+},\\ -1<w_{s}<\gamma&\hbox{in}\leavevmode\nobreak\ \mathbb{R}^{-}.\end{cases}

Note that there exists C0>0C_{0}>0 such that, for all s(12,1)s\in\left(\frac{1}{2},1\right),

(7.22) |ws(x)|1+|x|2𝑑x11+|x|2𝑑x<+and|W(ws)|C0.\int_{\mathbb{R}}\frac{|w_{s}(x)|}{1+|x|^{2}}\,dx\leqslant\int_{\mathbb{R}}\frac{1}{1+|x|^{2}}\,dx<+\infty\qquad\hbox{and}\qquad|W^{\prime}(w_{s})|\leqslant C_{0}.

Consider a ball B1(x0)B_{1}(x_{0})\Subset\mathbb{R}^{-}. By [CS, Theorem 26] with (7.22), there exist α1(0,1)\alpha_{1}\in(0,1) and C>0C>0 such that, for every s(12,1)s\in\left(\frac{1}{2},1\right),

[ws]Cα1(B1/2(x0))C.[w_{s}]_{C^{\alpha_{1}}(B_{1/2}(x_{0}))}\leqslant C.

On the other hand, by [ROSduke, Proposition 1.1], there exist α2(0,1)\alpha_{2}\in(0,1) and C>0C>0 such that, for every s(12,1)s\in\left(\frac{1}{2},1\right),

[ws|x|s]Cα2(B1/2(0)¯)C.\left[\frac{w_{s}}{|x|^{s}}\right]_{C^{\alpha_{2}}(\overline{B_{1/2}^{-}(0)})}\leqslant C.

The previous two displays yield (7.21).

By (7.21), there exists a function w0w_{0} such that, up to a subsequence,

lims12ws(x)=w0(x)locally uniformly in.\lim_{s\searrow\frac{1}{2}}w_{s}(x)=w_{0}(x)\quad\hbox{locally uniformly in}\leavevmode\nobreak\ \mathbb{R}.

Moreover, w0:[1,γ]w_{0}:\mathbb{R}\to[-1,\gamma], w0γw_{0}\equiv\gamma in \mathbb{R}^{-}, and w0w_{0} is non-decreasing.

For all s[12,1)s\in\left[\frac{1}{2},1\right), let u0su_{0}^{s} denote the corresponding solution to (6.2). Let xγsx_{\gamma}^{s} be such that u0s(xγs)=γu_{0}^{s}(x_{\gamma}^{s})=\gamma and set uγs(x):=u0s(x+xγs)u_{\gamma}^{s}(x):=u_{0}^{s}(x+x_{\gamma}^{s}). In particular, uγsu_{\gamma}^{s} solves

(7.23) {(Δ)suγs(x)+W(uγs(x))=0for allxlimx±uγs(x)=±1,uγs(0)=γ,(uγs)>0.\begin{cases}(-\Delta)^{s}u_{\gamma}^{s}(x)+W^{\prime}(u_{\gamma}^{s}(x))=0&\;\hbox{for all}\leavevmode\nobreak\ x\in\mathbb{R}\\ \displaystyle\lim_{x\to\pm\infty}u_{\gamma}^{s}(x)=\pm 1,&\\ u_{\gamma}^{s}(0)=\gamma,&\\ (u_{\gamma}^{s})^{\prime}>0.&\end{cases}

As in Corollary 6.3, we can show that, up to taking another subsequence,

(7.24) lims12uγs(x)=uγ12(x)locally uniformly in.\lim_{s\searrow\frac{1}{2}}u_{\gamma}^{s}(x)=u_{\gamma}^{\frac{1}{2}}(x)\quad\hbox{locally uniformly in}\leavevmode\nobreak\ \mathbb{R}.

Note here that

(7.25) the sequence xγsx_{\gamma}^{s} is bounded uniformly in s[12,1]s\in\left[\frac{1}{2},1\right].

Indeed, if there existed a subsequence such that xγs±x_{\gamma}^{s}\to\pm\infty as s12s\searrow\frac{1}{2}, then, by Lemma 6.1,

lims12|u0s(xγs)±1|lims12C|xγs|2slims12C|xγs|=0.\lim_{s\searrow\frac{1}{2}}|u_{0}^{s}(x_{\gamma}^{s})\pm 1|\leqslant\lim_{s\searrow\frac{1}{2}}\frac{C}{|x_{\gamma}^{s}|^{2s}}\leqslant\lim_{s\searrow\frac{1}{2}}\frac{C}{|x_{\gamma}^{s}|}=0.

On the other hand, since u0s(xγs)=γu_{0}^{s}(x_{\gamma}^{s})=\gamma for all s(12,1)s\in\left(\frac{1}{2},1\right),

lims12|u0s(xγs)±1|=lims12|γ±1|=|γ±1|>0,\lim_{s\searrow\frac{1}{2}}|u_{0}^{s}(x_{\gamma}^{s})\pm 1|=\lim_{s\searrow\frac{1}{2}}|\gamma\pm 1|=|\gamma\pm 1|>0,

thus giving the desired contradiction and proving (7.25).

From (7.25) we deduce that xγsx¯x_{\gamma}^{s}\to\bar{x} as s12s\searrow\frac{1}{2}. Moreover, we claim that

(7.26) u012(x¯)=γ.u_{0}^{\frac{1}{2}}(\bar{x})=\gamma.

To check this, we observe that

|u012(x¯)γ|\displaystyle|u_{0}^{\frac{1}{2}}(\bar{x})-\gamma| \displaystyle\leqslant |u012(x¯)u012(xγs)|+|u012(xγs)γ|\displaystyle|u_{0}^{\frac{1}{2}}(\bar{x})-u_{0}^{\frac{1}{2}}(x_{\gamma}^{s})|+|u_{0}^{\frac{1}{2}}(x_{\gamma}^{s})-\gamma|
=\displaystyle= |u012(x¯)u012(xγs)|+|u012(xγs)u0s(xγs)|.\displaystyle|u_{0}^{\frac{1}{2}}(\bar{x})-u_{0}^{\frac{1}{2}}(x_{\gamma}^{s})|+|u_{0}^{\frac{1}{2}}(x_{\gamma}^{s})-u_{0}^{s}(x_{\gamma}^{s})|.

Hence, taking the limit as s12s\searrow\frac{1}{2} and using the continuity of u012u_{0}^{\frac{1}{2}} and the uniform convergence in (7.24), we obtain (7.26).

It remains to check that (7.9) holds for s=12s=\frac{1}{2}. To this end, we recall that, for s(12,1)s\in\left(\frac{1}{2},1\right), Lemma 7.4 can be written as

(7.27) uγs(x)ws(x) for all xanduγs(0)=ws(0)=γ.u_{\gamma}^{s}(x)\geqslant w_{s}(x)\;\hbox{ for all }x\in\mathbb{R}\qquad\hbox{and}\qquad u_{\gamma}^{s}(0)=w_{s}(0)=\gamma.

Moreover, for all xx\in\mathbb{R},

|uγs(x)u012(x+x¯)||u0s(x+xγs)u012(x+xγs)|+|u012(x+xγs)u012(x+x¯)|.\displaystyle|u_{\gamma}^{s}(x)-u_{0}^{\frac{1}{2}}(x+\bar{x})|\leqslant|u_{0}^{s}(x+x_{\gamma}^{s})-u_{0}^{\frac{1}{2}}(x+x_{\gamma}^{s})|+|u_{0}^{\frac{1}{2}}(x+x_{\gamma}^{s})-u_{0}^{\frac{1}{2}}(x+\bar{x})|.

Thus, using the uniform convergence in (7.24) and the continuity of u012u_{0}^{\frac{1}{2}}, we obtain that, for all xx\in\mathbb{R},

lims12uγs(x)=u012(x+x¯).\lim_{s\searrow\frac{1}{2}}u_{\gamma}^{s}(x)=u_{0}^{\frac{1}{2}}(x+\bar{x}).

Recalling (7.26), this gives that, for all xx\in\mathbb{R},

lims12uγs(x)=uγ12(x).\lim_{s\searrow\frac{1}{2}}u_{\gamma}^{s}(x)=u_{\gamma}^{\frac{1}{2}}(x).

We can therefore pass to the limit the inequalities in (7.27) and obtain that

uγ12(x)w0(x) for all xanduγ12(0)=w0(0)=γ.u_{\gamma}^{\frac{1}{2}}(x)\geqslant w_{0}(x)\;\hbox{ for all }x\in\mathbb{R}\qquad\hbox{and}\qquad u_{\gamma}^{\frac{1}{2}}(0)=w_{0}(0)=\gamma.\qed
Proposition 7.6.

The function w0Xγw_{0}\in X_{\gamma} in Lemma 7.5 is a global minimizer of 𝒢12\mathcal{G}_{\frac{1}{2}}.

Furthermore, w0Cs()Clocα()w_{0}\in C^{s}(\mathbb{R})\cap C^{\alpha}_{\text{loc}}(\mathbb{R}^{-}) for all α(0,1)\alpha\in(0,1), w0w_{0} is strictly increasing in \mathbb{R}^{-}, 1<w0(x)<γ-1<w_{0}(x)<\gamma for all xx\in\mathbb{R}^{-}, and w0w_{0} solves (1.10) for s=12s=\frac{1}{2}.

Proof.

We first check that

(7.28) w0w_{0} is a local minimizer of 𝒢12\mathcal{G}_{\frac{1}{2}}.

For this, fix R>0R>0 and let ϕ\phi be a measurable function with suppϕBR\operatorname{supp}\phi\subset B_{R}^{-} and such that w0+ϕγw_{0}+\phi\leqslant\gamma. We will show that

(7.29) 𝒢12(w0+ϕ,BR)𝒢12(w0,BR)0.{\mathcal{G}}_{\frac{1}{2}}(w_{0}+\phi,B_{R}^{-})-{\mathcal{G}}_{\frac{1}{2}}(w_{0},B_{R}^{-})\geqslant 0.

It is enough to consider the case in which ϕ\phi is smooth with suppϕBR\operatorname{supp}\phi\Subset B_{R}^{-}. Since wsw_{s} is a local minimizer of 𝒢s\mathcal{G}_{s} for s(12,1)s\in\left(\frac{1}{2},1\right),

(7.30) 𝒢s(ws+ϕ,BR)𝒢s(ws,BR)0.{\mathcal{G}}_{s}(w_{s}+\phi,B_{R}^{-})-{\mathcal{G}}_{s}(w_{s},B_{R}^{-})\geqslant 0.

By writing out the expression for the energies, we see that (7.30) is equivalent to

(7.31) 0BRBR|(ws+ϕ)(x)(ws+ϕ)(y)|2|xy|1+2s𝑑y𝑑x+2(BR)cBR|(ws+ϕ)(x)(ws+ϕ)(y)|2|xy|1+2s𝑑y𝑑xBRBR|ws(x)ws(y)|2|xy|1+2s𝑑y𝑑x2(BR)cBR|ws(x)ws(y)|2|xy|1+2s𝑑y𝑑x+BR[W(ws(x)+ϕ(x))W(ws(x))]𝑑x=2BRBR(ws(x)ws(y))(ϕ(x)ϕ(y))|xy|1+2s𝑑y𝑑x+4(BR)cBR(ws(x)ws(y))(ϕ(x)ϕ(y))|xy|1+2s𝑑y𝑑x+[BRBR|ϕ(x)ϕ(y)|2|xy|1+2s𝑑y𝑑x+2(BR)cBR|ϕ(x)|2|xy|1+2s𝑑y𝑑x]+BR[W(ws(x)+ϕ(x))W(ws(x))]𝑑x.\begin{split}0&\leqslant\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{|(w_{s}+\phi)(x)-(w_{s}+\phi)(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &\quad+2\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{|(w_{s}+\phi)(x)-(w_{s}+\phi)(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &\quad-\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{|w_{s}(x)-w_{s}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx-2\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{|w_{s}(x)-w_{s}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &\quad+\int_{B_{R}^{-}}\big{[}W(w_{s}(x)+\phi(x))-W(w_{s}(x))\big{]}\,dx\\ &=2\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))}{|x-y|^{1+2s}}\,dy\,dx\\ &\quad+4\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))}{|x-y|^{1+2s}}\,dy\,dx\\ &\quad+\left[\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{|\phi(x)-\phi(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx+2\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{|\phi(x)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\right]\\ &\quad+\int_{B_{R}^{-}}\big{[}W(w_{s}(x)+\phi(x))-W(w_{s}(x))\big{]}\,dx.\end{split}

By the Dominated Convergence Theorem, since ϕ\phi is smooth and bounded,

lims12\displaystyle\lim_{s\searrow\frac{1}{2}} [BRBR|ϕ(x)ϕ(y)|2|xy|1+2s𝑑y𝑑x+2(BR)cBR|ϕ(x)|2|xy|1+2s𝑑y𝑑x]\displaystyle\left[\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{|\phi(x)-\phi(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx+2\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{|\phi(x)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\right]
=BRBR|ϕ(x)ϕ(y)|2|xy|2𝑑y𝑑x+2(BR)cBR|ϕ(x)|2|xy|2𝑑y𝑑x\displaystyle=\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{|\phi(x)-\phi(y)|^{2}}{|x-y|^{2}}\,dy\,dx+2\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{|\phi(x)|^{2}}{|x-y|^{2}}\,dy\,dx

and since |ws|1|w_{s}|\leqslant 1 and, for ss close to 12\frac{1}{2}, |ws+ϕ|1|w_{s}+\phi|\leqslant 1,

lims12BR[W(ws(x)+ϕ(x))W(ws(x))]𝑑x=BR[W(w0(x)+ϕ(x))W(w0(x))]𝑑x.\lim_{s\searrow\frac{1}{2}}\int_{B_{R}^{-}}\big{[}W(w_{s}(x)+\phi(x))-W(w_{s}(x))\big{]}\,dx=\int_{B_{R}^{-}}\big{[}W(w_{0}(x)+\phi(x))-W(w_{0}(x))\big{]}\,dx.

Moreover, thanks to (7.21), there exists some C>0C>0, independent of ss, such that, for ss close to 12\frac{1}{2},

BR\displaystyle\int_{B_{R}^{-}} BR|(ws(x)ws(y))(ϕ(x)ϕ(y))||xy|1+2s𝑑y𝑑x\displaystyle\int_{B_{R}^{-}}\frac{|(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))|}{|x-y|^{1+2s}}\,dy\,dx
=BRBR{|xy|>1}|(ws(x)ws(y))(ϕ(x)ϕ(y))||xy|1+2s𝑑y𝑑x\displaystyle=\int_{B_{R}^{-}}\int_{B_{R}^{-}\cap\{|x-y|>1\}}\frac{|(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))|}{|x-y|^{1+2s}}\,dy\,dx
+BRBR{|xy|<1}|(ws(x)ws(y))(ϕ(x)ϕ(y))||xy|1+2s𝑑y𝑑x\displaystyle\quad+\int_{B_{R}^{-}}\int_{B_{R}^{-}\cap\{|x-y|<1\}}\frac{|(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))|}{|x-y|^{1+2s}}\,dy\,dx
BRBR{|xy|>1}C|xy|1+2s𝑑y𝑑x+BRBR{|xy|<1}C|xy|α+1|xy|1+2s𝑑y𝑑xC\displaystyle\leqslant\int_{B_{R}^{-}}\int_{B_{R}^{-}\cap\{|x-y|>1\}}\frac{C}{|x-y|^{1+2s}}\,dy\,dx+\int_{B_{R}^{-}}\int_{B_{R}^{-}\cap\{|x-y|<1\}}\frac{C|x-y|^{\alpha+1}}{|x-y|^{1+2s}}\,dy\,dx\leqslant C

and

(BR)c\displaystyle\int_{(B_{R}^{-})^{c}} BR|(ws(x)ws(y))(ϕ(x)ϕ(y))||xy|1+2s𝑑y𝑑x\displaystyle\int_{B_{R}^{-}}\frac{|(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))|}{|x-y|^{1+2s}}\,dy\,dx
=(BR)cBR{|xy|>1}|(ws(x)ws(y))(ϕ(x)ϕ(y))||xy|1+2s𝑑y𝑑x\displaystyle=\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}\cap\{|x-y|>1\}}\frac{|(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))|}{|x-y|^{1+2s}}\,dy\,dx
+(BR)cBR{|xy|<1}|(ws(x)ws(y))(ϕ(x)ϕ(y))||xy|1+2s𝑑y𝑑x\displaystyle\quad+\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}\cap\{|x-y|<1\}}\frac{|(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))|}{|x-y|^{1+2s}}\,dy\,dx
=C(BR)cBR{|xy|>1}C|xy|1+2s𝑑y𝑑x\displaystyle=C\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}\cap\{|x-y|>1\}}\frac{C}{|x-y|^{1+2s}}dy\,dx
+(BR)cBR{|xy|<1}C|xy|α+1|xy|1+2s𝑑y𝑑xC.\displaystyle\quad+\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}\cap\{|x-y|<1\}}\frac{C|x-y|^{\alpha+1}}{|x-y|^{1+2s}}\,dy\,dx\leqslant C.

Therefore, by the Dominated Convergence Theorem,

lims12\displaystyle\lim_{s\searrow\frac{1}{2}} BRBR(ws(x)ws(y))(ϕ(x)ϕ(y))|xy|1+2s𝑑y𝑑x\displaystyle\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))}{|x-y|^{1+2s}}\,dy\,dx
=BRBR(w0(x)w0(y))(ϕ(x)ϕ(y))|xy|2𝑑y𝑑x\displaystyle=\int_{B_{R}^{-}}\int_{B_{R}^{-}}\frac{(w_{0}(x)-w_{0}(y))(\phi(x)-\phi(y))}{|x-y|^{2}}\,dy\,dx

and

lims12\displaystyle\lim_{s\searrow\frac{1}{2}} (BR)cBR(ws(x)ws(y))(ϕ(x)ϕ(y))|xy|1+2s𝑑y𝑑x\displaystyle\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{(w_{s}(x)-w_{s}(y))(\phi(x)-\phi(y))}{|x-y|^{1+2s}}\,dy\,dx
=(BR)cBR(w0(x)w0(y))(ϕ(x)ϕ(y))|xy|2𝑑y𝑑x.\displaystyle=\int_{(B_{R}^{-})^{c}}\int_{B_{R}^{-}}\frac{(w_{0}(x)-w_{0}(y))(\phi(x)-\phi(y))}{|x-y|^{2}}\,dy\,dx.

Collecting the above limits, we take the limit as s12s\searrow\frac{1}{2} of (7.30) written as (7.31) and obtain that (7.29) holds. Namely, w0w_{0} is a local minimizer, thus completing the proof of (7.28).

We next show that

(7.32) w0w_{0} is a global minimizer of 𝒢12\mathcal{G}_{\frac{1}{2}}.

In light of (7.28), in order to establish (7.32) we only need to check that

𝒢12(w0)<+.\mathcal{G}_{\frac{1}{2}}(w_{0})<+\infty.

For this, let hh be as in Lemma A.2. Fix s(12,s0]s\in\left(\frac{1}{2},s_{0}\right] (with s0s_{0} as given in (6.5)). Notice that the function hh belongs to XγX_{\gamma} (recall the definition of XγX_{\gamma} on page 1.2) and therefore, since wsw_{s} is a minimizer of 𝒢s\mathcal{G}_{s} in XγX_{\gamma}, we have that

(7.33) 𝒢s(ws)𝒢s(h)<+.\mathcal{G}_{s}(w_{s})\leqslant\mathcal{G}_{s}(h)<+\infty.

Let now R>3R>3. We point out that if uu is such that 𝒢s(u)<+\mathcal{G}_{s}(u)<+\infty, then one can check that

(7.34) 𝒢s(u)𝒢s(u,[R,0])=RR|u(x)u(y)|2|xy|1+2s𝑑y𝑑x+2R0+|u(x)u(y)|2|xy|1+2s𝑑y𝑑x+RW(u)𝑑x.\begin{split}&\mathcal{G}_{s}(u)-\mathcal{G}_{s}(u,[-R,0])\\ &=\int_{-\infty}^{-R}\int_{-\infty}^{-R}\frac{|u(x)-u(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx+2\int_{-\infty}^{-R}\int_{0}^{+\infty}\frac{|u(x)-u(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &\qquad\qquad+\int_{-\infty}^{-R}W(u)\,dx.\end{split}

Using this with u:=hu:=h, since h1h\equiv-1 in (,R)(-\infty,-R) and hγh\equiv\gamma in +\mathbb{R}^{+}, we have that

𝒢s(h)𝒢s(h,[R,0])=2R0+(1+γ)2(yx)1+2s𝑑y𝑑x.\mathcal{G}_{s}(h)-\mathcal{G}_{s}(h,[-R,0])=2\int_{-\infty}^{-R}\int_{0}^{+\infty}\frac{(1+\gamma)^{2}}{(y-x)^{1+2s}}\,dy\,dx.

Therefore, exploiting (7.34) with u:=wsu:=w_{s} and recalling also (7.33),

𝒢s(ws,[R,0])\displaystyle\mathcal{G}_{s}(w_{s},[-R,0]) 𝒢s(ws)2R0+|ws(x)ws(y)|2|xy|1+2s𝑑y𝑑x\displaystyle\leqslant\mathcal{G}_{s}(w_{s})-2\int_{-\infty}^{-R}\int_{0}^{+\infty}\frac{|w_{s}(x)-w_{s}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx
𝒢s(h)2R0+(γws(x))2(yx)1+2s𝑑y𝑑x\displaystyle\leqslant\mathcal{G}_{s}(h)-2\int_{-\infty}^{-R}\int_{0}^{+\infty}\frac{(\gamma-w_{s}(x))^{2}}{(y-x)^{1+2s}}\,dy\,dx
=𝒢s(h,[R,0])+2R0+(γ+1)2(γws(x))2(yx)1+2s𝑑y𝑑x.\displaystyle=\mathcal{G}_{s}(h,[-R,0])+2\int_{-\infty}^{-R}\int_{0}^{+\infty}\frac{(\gamma+1)^{2}-(\gamma-w_{s}(x))^{2}}{(y-x)^{1+2s}}\,dy\,dx.

Now, for RR sufficiently large, Lemmata 7.4 and 6.1 imply that

R0+|(γ+1)2(γws(x))2|(yx)1+2s𝑑y𝑑x=12sR|(γ+1)2(γws(x))2||x|2s𝑑x\displaystyle\int_{-\infty}^{-R}\int_{0}^{+\infty}\frac{|(\gamma+1)^{2}-(\gamma-w_{s}(x))^{2}|}{(y-x)^{1+2s}}\,dy\,dx=\frac{1}{2s}\int_{-\infty}^{-R}\frac{|(\gamma+1)^{2}-(\gamma-w_{s}(x))^{2}|}{|x|^{2s}}\,dx
CR1+ws(x)|x|2s𝑑xCR1+uγs(x)|x|2s𝑑x\displaystyle\qquad\qquad\leqslant C\int_{-\infty}^{-R}\frac{1+w_{s}(x)}{|x|^{2s}}\,dx\leqslant C\int_{-\infty}^{-R}\frac{1+u_{\gamma}^{s}(x)}{|x|^{2s}}\,dx
CRdx|x|4sCRdx|x|2=CRC,\displaystyle\qquad\qquad\leqslant C\int_{-\infty}^{-R}\frac{dx}{|x|^{4s}}\leqslant C\int_{-\infty}^{-R}\frac{dx}{|x|^{2}}=\frac{C}{R}\leqslant C,

where CC is independent of s(12,s0]s\in\left(\frac{1}{2},s_{0}\right].

With this and Lemma A.2, we have that

lims12[𝒢s(h,[R,0])+2R0+(γ+1)2(γws(x))2(yx)1+2s𝑑y𝑑x]C(1+lnR).\lim_{s\searrow\frac{1}{2}}\left[\mathcal{G}_{s}(h,[-R,0])+2\int_{-\infty}^{-R}\int_{0}^{+\infty}\frac{(\gamma+1)^{2}-(\gamma-w_{s}(x))^{2}}{(y-x)^{1+2s}}\,dy\,dx\right]\leqslant C(1+\ln R).

Therefore, with Fatou’s Lemma,

𝒢12(w0,[R,0])lims12𝒢s(ws,[R,0])C(1+lnR),\mathcal{G}_{\frac{1}{2}}(w_{0},[-R,0])\leqslant\lim_{s\searrow\frac{1}{2}}\mathcal{G}_{s}(w_{s},[-R,0])\leqslant C(1+\ln R),

and we have that

𝒢12(w0)=limR+𝒢12(w0,[R,0])lnRC,\mathcal{G}_{\frac{1}{2}}(w_{0})=\lim_{R\to+\infty}\frac{\mathcal{G}_{\frac{1}{2}}(w_{0},[-R,0])}{\ln R}\leqslant C,

as desired. Hence, w0w_{0} is a global minimizer of 𝒢12\mathcal{G}_{\frac{1}{2}} and (7.32) is thereby established.

Finally, we check the remaining properties in the statement of Proposition 7.6. Recall uγ12u_{\gamma}^{\frac{1}{2}} in (7.23) and that, by Lemma 7.5, the estimate (7.9) holds. Since uγ12u_{\gamma}^{\frac{1}{2}} is strictly decreasing, this implies that w0<γw_{0}<\gamma in \mathbb{R}^{-} and also

0limx(w0(x)+1)limx(uγ12(x)+1)=0.0\leqslant\lim_{x\to-\infty}(w_{0}(x)+1)\leqslant\lim_{x\to-\infty}(u_{\gamma}^{\frac{1}{2}}(x)+1)=0.

With this and the minimization property, w0Xγw_{0}\in X_{\gamma} solves (1.10) for s=12s=\frac{1}{2}. The remaining properties follow from Lemma 7.2. ∎

7.4. Asymptotics

We now prove the asymptotic behavior at -\infty of global minimizers.

Lemma 7.7.

Let s[12,1)s\in\left[\frac{1}{2},1\right) and γ(1,1)\gamma\in(-1,1). The global minimizers in Propositions 7.3 and 7.6 satisfy the estimates (1.11) and (1.12) and, if s(12,1)s\in\left(\frac{1}{2},1\right), satisfy (1.13).

Proof.

We begin by proving (1.11). By (6.3) together with (7.9) (which holds also for s=12s=\frac{1}{2}, recall Lemma 7.5), there exist C~\tilde{C}, R~1\tilde{R}\geqslant 1 such that

w0(x)+1uγ(x)+1C|x+xγ|2s for all x+xγR~.w_{0}(x)+1\leqslant u_{\gamma}(x)+1\leqslant\frac{C}{|x+x_{\gamma}|^{2s}}\quad\hbox{ for all }x+x_{\gamma}\leqslant-\tilde{R}.

In particular, (1.11) holds with some CC, R1R\geqslant 1 depending on γ\gamma.

Let s(12,1)s\in\left(\frac{1}{2},1\right). We prove the first estimate in (1.13) for w0w_{0}^{\prime}. For this, let CC, RR be given by (1.11), fix x0<2Rx_{0}<-2R and set K:=|x0|K:=|x_{0}|. Let v(x):=w0(KxK)v(x):=w_{0}(Kx-K). One can readily check that v:(1,γ]v:\mathbb{R}\to(-1,\gamma] satisfies

{(Δ)sv(x)=K2sW(v(x))for x<1,v(x)=γfor x>1,0<v(x)+1<CK2s|x1|2sfor x<1RK.\begin{cases}-(-\Delta)^{s}v(x)=K^{2s}W^{\prime}(v(x))&\hbox{for }x<1,\\ v(x)=\gamma&\hbox{for }x>1,\\ \displaystyle 0<v(x)+1<\frac{C}{K^{2s}|x-1|^{2s}}&\hbox{for }x<1-\frac{R}{K}.\end{cases}

By the interior regularity theory for the fractional Laplacian, see for example [ROSerra],

(7.35) vC2s(B1/4)C(K2sW(v)L(B1/2)+vL()).\|v\|_{C^{2s}(B_{1/4})}\leqslant C\Big{(}K^{2s}\|W^{\prime}(v)\|_{L^{\infty}(B_{1/2})}+\|v\|_{L^{\infty}(\mathbb{R})}\Big{)}.

Since 0<RK<120<\frac{R}{K}<\frac{1}{2}, we have that

(7.36) supB1/2|W(v)|CsupB1/2|v+1|CK2ssupB1/2|x1|2sCK2s.\sup_{B_{1/2}}|W^{\prime}(v)|\leqslant C\sup_{B_{1/2}}|v+1|\leqslant\frac{C}{K^{2s}}\sup_{B_{1/2}}|x-1|^{-2s}\leqslant\frac{C}{K^{2s}}.

Moreover,

(7.37) vC2s(B1/4)vL(B1/4)=Kw0L(BK/4(K))Kw0(K).\|v\|_{C^{2s}(B_{1/4})}\geqslant\|v^{\prime}\|_{L^{\infty}(B_{1/4})}=K\|w_{0}^{\prime}\|_{L^{\infty}(B_{K/4}(-K))}\geqslant Kw_{0}^{\prime}(-K).

Combining (7.35), (7.36), and (7.37), and using that |v|1|v|\leqslant 1, up to renaming CC,

Kw0(K)C.Kw_{0}^{\prime}(-K)\leqslant C.

In particular,

(7.38) w0(x0)C|x0| for any x0<2Rw_{0}^{\prime}(x_{0})\leqslant\frac{C}{|x_{0}|}\quad\hbox{ for any }x_{0}<-2R

and therefore the first estimate in (1.13) holds for w0w_{0}^{\prime} with 2R2R in place of RR.

We now prove (1.12). Let s[12,1)s\in\left[\frac{1}{2},1\right) and α(0,1)\alpha\in(0,1). Take x0x_{0}, KK, and vv as in the previous paragraph. By interior regularity (see again [ROSerra]) and estimating as above,

[v]Cα(B1/4)\displaystyle[v]_{C^{\alpha}(B_{1/4})} C(K2sW(v)L(B1/2)+vL())C.\displaystyle\leqslant C\Big{(}K^{2s}\|W^{\prime}(v)\|_{L^{\infty}(B_{1/2})}+\|v\|_{L^{\infty}(\mathbb{R})}\Big{)}\leqslant C.

Rescaling back, we get that

[w0]Cα(BK/4(K))CKα.[w_{0}]_{C^{\alpha}(B_{K/4}(-K))}\leqslant\frac{C}{K^{\alpha}}.

Recalling the definition of KK, we obtain that

[w0]Cα(B|x0|/4(x0))C|x0|αCRα for all x0<R[w_{0}]_{C^{\alpha}(B_{|x_{0}|/4}(x_{0}))}\leqslant\frac{C}{|x_{0}|^{\alpha}}\leqslant\frac{C}{R^{\alpha}}\quad\hbox{ for all }x_{0}<-R

and the estimate in (1.12) follows.

By a similar argument, we can use (7.35) to show the second estimate in (1.13) when s(12,1)s\in\left(\frac{1}{2},1\right). ∎

7.5. Uniqueness

Finally, we show that minimizers in XγX_{\gamma} are unique.

Lemma 7.8.

Let s[12,1)s\in\left[\frac{1}{2},1\right) and γ(1,1)\gamma\in(-1,1). The global minimizer w0Xγw_{0}\in X_{\gamma} of 𝒢s\mathcal{G}_{s} is unique.

Proof.

Let w0w_{0} be the minimizer found in Propositions 7.3 and 7.6, and let wXγw\in X_{\gamma} be another minimizer. Since cutting at level 1-1 decreases the energy, we have that 1wγ-1\leqslant w\leqslant\gamma in \mathbb{R}^{-}.

Set v:=ww0v:=w-w_{0}. Since ww and w0w_{0} solve

{(Δ)sw+W(w)=(Δ)sw0+W(w0)=0 in ,w=w0=γ in +,\begin{cases}(-\Delta)^{s}w+W^{\prime}(w)=(-\Delta)^{s}w_{0}+W^{\prime}(w_{0})=0&\hbox{ in }\mathbb{R}^{-},\\ w=w_{0}=\gamma&\hbox{ in }\mathbb{R}^{+},\end{cases}

we have that

{(Δ)sv=f in ,v=0 in +,\begin{cases}(-\Delta)^{s}v=f&\hbox{ in }\mathbb{R}^{-},\\ v=0&\hbox{ in }\mathbb{R}^{+},\end{cases}

where f:=W(w0)W(w)L()f:=W^{\prime}(w_{0})-W^{\prime}(w)\in L^{\infty}(\mathbb{R}^{-}) since |w|,|w0|1|w|,|w_{0}|\leqslant 1.

Since ww, w0XγHs()w_{0}\in X_{\gamma}\subset H^{s}(\mathbb{R}), we have that vHs()v\in H^{s}(\mathbb{R}). Moreover, since |v|2|v|\leqslant 2,

|v(y)|(1+|y|)1+2s𝑑y2(1+|y|)1+2s𝑑y<+.\int_{\mathbb{R}}\frac{|v(y)|}{(1+|y|)^{1+2s}}\,dy\leqslant\int_{\mathbb{R}}\frac{2}{(1+|y|)^{1+2s}}\,dy<+\infty.

Fix R>0R>0. By [RosOtonWeidner, Corollary 6.3], there exists C>0C>0, depending only on ss, such that

[v]Cs(BR)CRs.[v]_{C^{s}(B_{R}^{-})}\leqslant CR^{-s}.

Sending R+R\to+\infty, we find that [v]Cs()=0[v]_{C^{s}(\mathbb{R}^{-})}=0 which implies that vv is constant in \mathbb{R}^{-}. Therefore,

w=w0+C in andw=w0=γ in +w=w_{0}+C\;\hbox{ in }\mathbb{R}^{-}\qquad{\mbox{and}}\qquad w=w_{0}=\gamma\;\hbox{ in }\mathbb{R}^{+}

for some CC\in\mathbb{R}. Since ww, w0Cs()w_{0}\in C^{s}(\mathbb{R}), it must be that C=0C=0. Consequently, ww0w\equiv w_{0} and thus the minimizer is unique. ∎

7.6. Proof of Theorem 1.3

With the work done so far, we can now complete the proof of Theorem 1.3.

Proof of Theorem 1.3.

The theorem follows from Proposition 7.3 for s(12,1)s\in\left(\frac{1}{2},1\right) and Proposition 7.6 for s=12s=\frac{1}{2} together with Lemmata 7.7 and 7.8. ∎

8. Bounding the energy from below for s[12,1)s\in\left[\frac{1}{2},1\right)

Here, we establish the Γlim inf\Gamma-\liminf inequality for the first-order convergence in Theorem 1.6.

8.1. Interpolation near the boundary

We begin by proving an adaptation of [SV-gamma, Proposition 4.1] near Ω\partial\Omega. For x¯Ω\bar{x}\in\partial\Omega and ρo>0\rho_{o}>0, consider Bρo(x¯)B_{\rho_{o}}(\bar{x})\subset\mathbb{R}. Set

D:=Bρo(x¯)Ω,D:=B_{\rho_{o}}(\bar{x})\cap\Omega,

and, for small t>0t>0, define

Dt:={xD s.t. dDΩ(x)>t}D_{t}:=\big{\{}x\in D\;{\mbox{ s.t. }}\;d_{\partial D\setminus\partial\Omega}(x)>t\big{\}}

where dDΩ(x)d_{\partial D\setminus\partial\Omega}(x) denotes the distance from a point xx to DΩ=D{x¯}\partial D\setminus\partial\Omega=\partial D\setminus\{\bar{x}\}.

Recall also the notation (5.1).

Proposition 8.1.

Fix δ>0\delta>0 small and γ(1,1)\gamma\in(-1,1). Let εk0+\varepsilon_{k}\searrow 0^{+} as k+k\to+\infty, and let uku_{k} be a sequence in L1(Ω)L^{1}(\Omega) and wkw_{k} be a sequence in L1(Bρo(x¯)Ω)L^{1}(B_{\rho_{o}}(\bar{x})\cup\Omega) such that

{ukwk0ask+ in L1(DDδ),uk=g in Ωc,wk=g(0) in Bρo(x¯)D,|uk|1andwkγinukγinD.\begin{cases}u_{k}-w_{k}\to 0\quad\hbox{as}\leavevmode\nobreak\ k\to+\infty&\hbox{ in }L^{1}(D\setminus D_{\delta}),\\ u_{k}=g&\hbox{ in }\Omega^{c},\\ w_{k}=g(0)&\hbox{ in }B_{\rho_{o}}(\bar{x})\setminus D,\\ |u_{k}|\leqslant 1\leavevmode\nobreak\ \hbox{and}\leavevmode\nobreak\ w_{k}\leqslant\gamma&\hbox{in}\leavevmode\nobreak\ \mathbb{R}\\ u_{k}\leqslant\gamma&\hbox{in}\leavevmode\nobreak\ D.\end{cases}

Then, there exists a sequence vkγv_{k}\leqslant\gamma such that

vk(x)={uk(x)ifxDδ,wk(x)ifxDcv_{k}(x)=\begin{cases}u_{k}(x)&\hbox{if}\leavevmode\nobreak\ x\in D_{\delta},\\ w_{k}(x)&\hbox{if}\leavevmode\nobreak\ x\in D^{c}\end{cases}

and

lim supk+εk(1)(vk,Ω)lim supk+[εk(1)(wk,Ω)εk(1)(wk,Dδ)+Iεk(uk,Bρo(x¯),Ω)].\limsup_{k\to+\infty}\mathcal{F}_{\varepsilon_{k}}^{(1)}(v_{k},\Omega)\leqslant\limsup_{k\to+\infty}\Big{[}\mathcal{F}_{\varepsilon_{k}}^{(1)}(w_{k},\Omega)-\mathcal{F}_{\varepsilon_{k}}^{(1)}(w_{k},D_{\delta})+I_{\varepsilon_{k}}(u_{k},B_{\rho_{o}}(\bar{x}),\Omega)\Big{]}.
Proof.

For ease in the proof, we assume that

x¯=0,ρo=1,D=B1,so thatDt=B1t,\bar{x}=0,\quad\rho_{o}=1,\quad D=B_{1}^{-},\quad\hbox{so that}\quad D_{t}=B_{1-t}^{-},

and take s(12,1)s\in\left(\frac{1}{2},1\right). The general setting follows along the same lines, and the proof for s=12s=\frac{1}{2} is similar (see [SV-gamma, Proposition 4.1]). Furthermore, we drop the notation for the subscript kk.

Also, we may assume that there exists some C0>0C_{0}>0 such that

(8.1) ε(1)(w,Ω)ε(1)(w,Dδ)+Iε(u,D,Ω)C0,\mathcal{F}_{\varepsilon}^{(1)}(w,\Omega)-\mathcal{F}_{\varepsilon}^{(1)}(w,D_{\delta})+I_{\varepsilon}(u,D,\Omega)\leqslant C_{0},

otherwise there is nothing to show. Since

(8.2) w(QΩ)w(QDδ)=w(ΩDδ,ΩDδ)+2w(ΩDδ,Ωc)w(ΩDδ,(Dδ)c),\begin{split}w(Q_{\Omega})-w(Q_{D_{\delta}})&=w(\Omega\setminus D_{\delta},\Omega\setminus D_{\delta})+2w(\Omega\setminus D_{\delta},\Omega^{c})\\ &\geqslant w(\Omega\setminus D_{\delta},(D_{\delta})^{c}),\end{split}

we obtain from (8.1) that

(8.3) w(ΩDδ,(Dδ)c)+u(DDδ,D)C0ε12s.w(\Omega\setminus D_{\delta},(D_{\delta})^{c})+u(D\setminus D_{\delta},D)\leqslant C_{0}\varepsilon^{1-2s}.

The rest of the proof is broken into four steps. First, we will partition the set DDδD\setminus D_{\delta} into a finite sequence of intervals so that one of the intervals satisfies an estimate similar to (8.3) with right-hand side small for all ε\varepsilon. Then, we perform a second partition to find a subinterval such that the difference between uu and ww is sufficiently small there. Next, we use the first two steps to appropriately partition \mathbb{R} and construct the desired function vv. Lastly, we estimate errors.

Step 1. (Partition DDδD\setminus D_{\delta}). Fix σ>0\sigma>0 small. For M=M(σ)>1M=M(\sigma)>1, to be determined, set

δ~:=δM.\tilde{\delta}:=\frac{\delta}{M}.

We partition DDδ=(1,1+δ)D\setminus D_{\delta}=(-1,-1+\delta) into MM disjoint intervals Djδ~D(j+1)δ~D_{j\tilde{\delta}}\setminus D_{(j+1)\tilde{\delta}}, so that

DDδ=j=0M1Djδ~D(j+1)δ~.D\setminus D_{\delta}=\bigcup_{j=0}^{M-1}D_{j\tilde{\delta}}\setminus D_{(j+1)\tilde{\delta}}.

From (8.3), and since DΩD\subset\Omega, we have that

C0ε12sj=0M1[w(Djδ~D(j+1)δ~,(Dδ)c)+u(Djδ~D(j+1)δ~,D)].C_{0}\varepsilon^{1-2s}\geqslant\sum_{j=0}^{M-1}\left[w(D_{j\tilde{\delta}}\setminus D_{(j+1)\tilde{\delta}},(D_{\delta})^{c})+u(D_{j\tilde{\delta}}\setminus D_{(j+1)\tilde{\delta}},D)\right].

For M=M(σ)M=M(\sigma) sufficiently large, there exists some 0jM10\leqslant j\leqslant M-1 such that

w(Djδ~D(j+1)δ~,(Dδ)c)+u(Djδ~D(j+1)δ~,D)σε12s.w(D_{j\tilde{\delta}}\setminus D_{(j+1)\tilde{\delta}},(D_{\delta})^{c})+u(D_{j\tilde{\delta}}\setminus D_{(j+1)\tilde{\delta}},D)\leqslant\sigma\varepsilon^{1-2s}.

Fix such a jj and set

D~:=Djδ~\widetilde{D}:=D_{j\tilde{\delta}}

Since D~D\widetilde{D}\subset D, we have that

(8.4) w(D~D~δ~,(Dδ)c)+u(D~D~δ~,D~)σε12s.w(\widetilde{D}\setminus\widetilde{D}_{\tilde{\delta}},(D_{\delta})^{c})+u(\widetilde{D}\setminus\widetilde{D}_{\tilde{\delta}},\widetilde{D})\leqslant\sigma\varepsilon^{1-2s}.

Step 2. (Partition D~D~δ~\widetilde{D}\setminus\widetilde{D}_{\tilde{\delta}}). Let NN\in\mathbb{N} be the integer part of δ~/(2ε)\tilde{\delta}/(2\varepsilon) and let x~\tilde{x} be the left endpoint of D~\widetilde{D}. For 0iN10\leqslant i\leqslant N-1, we define

Ai:={xD~s.t.iε<(xx~)(i+1)ε}A_{i}:=\big{\{}x\in\widetilde{D}\leavevmode\nobreak\ \hbox{s.t.}\leavevmode\nobreak\ i\varepsilon<(x-\tilde{x})\leqslant(i+1)\varepsilon\big{\}}

and observe that AiD~D~δ~A_{i}\subset\widetilde{D}\setminus\widetilde{D}_{\tilde{\delta}}.

Denote by

(8.5) di(x):=dD~εi(x),d_{i}(x):=d_{\partial\widetilde{D}_{\varepsilon i}}(x),

the distance from xx to the boundary of D~εi\widetilde{D}_{\varepsilon i}. As in the proof of [SV-gamma, Proposition 4.1], we can show that there exists some 0iN10\leqslant i\leqslant N-1 such that

(8.6) Ai|u(x)w(x)|𝑑x+ε2sD~(i+1)εD~δ~|u(x)w(x)|di(x)2s𝑑xσε.\int_{A_{i}}|u(x)-w(x)|\,dx+\varepsilon^{2s}\int_{\widetilde{D}_{(i+1)\varepsilon}\setminus\widetilde{D}_{\tilde{\delta}}}|u(x)-w(x)|d_{i}(x)^{-2s}\,dx\leqslant\sigma\varepsilon.

From now on such an ii is fixed once and for all.

Step 3. (Partition \mathbb{R} and construct vv). We partition \mathbb{R} into the following six disjoint regions:

P:=D~δ~,Q:=D~(i+1)εD~δ~,R:=Ai,S:=ΩD~εi,T:=ΩcBδ+,U:=Bδ+.\begin{array}[]{lll}P:=\widetilde{D}_{\tilde{\delta}},&Q:=\widetilde{D}_{(i+1)\varepsilon}\setminus\widetilde{D}_{\tilde{\delta}},&R:=A_{i},\\ S:=\Omega\setminus\widetilde{D}_{\varepsilon i},&T:=\Omega^{c}\setminus B_{\delta}^{+},&U:=B_{\delta}^{+}.\end{array}

Note that

(8.7) Ω=PQRSandΩc=TU.\Omega=P\cup Q\cup R\cup S\qquad{\mbox{and}}\qquad\Omega^{c}=T\cup U.

Let ϕ\phi be a smooth cutoff function such that ϕ=1\phi=1 on PQP\cup Q, ϕ=0\phi=0 on STS\cup T, and ϕL()3/ε\|\phi^{\prime}\|_{L^{\infty}(\mathbb{R})}\leqslant 3/\varepsilon. Define vv by

v:={ϕu+(1ϕ)win=PQRST,win+=T+U.v:=\begin{cases}\phi u+(1-\phi)w&\hbox{in}\leavevmode\nobreak\ \mathbb{R}^{-}=P\cup Q\cup R\cup S\cup T^{-},\\ w&\hbox{in}\leavevmode\nobreak\ \mathbb{R}^{+}=T^{+}\cup U.\end{cases}

By construction, it is clear that vγv\leqslant\gamma in \mathbb{R}, v=uv=u in PQDδP\cup Q\supset D_{\delta} and v=wv=w in (B1)c=Dc(B_{1}^{-})^{c}=D^{c}. It remains to show that vv satisfies the energy estimate in Proposition 8.1.

Starting with the kinetic energy, we will show that

(8.8) v(QΩ)w(QΩ)w(QDδ)+u(B1,B1)+2u(B1,B1+)+O(σε12s)+O(δ~2s).v(Q_{\Omega})\leqslant w(Q_{\Omega})-w(Q_{D_{\delta}})+u(B_{1}^{-},B_{1}^{-})+2u(B_{1}^{-},B_{1}^{+})+O(\sigma\varepsilon^{1-2s})+O(\tilde{\delta}^{-2s}).

Since SΩDδS\subset\Omega\setminus D_{\delta}, one can check (see [SV-gamma]) that

w(S,S)+2w(S,TU)+w(QDδ)w(QΩ).\displaystyle w(S,S)+2w(S,T\cup U)+w(Q_{D_{\delta}})\leqslant w(Q_{\Omega}).

With this and noting that

u(PQ,PQ)+2u(PQ,U)u(B1,B1)+2u(B1,B1+),u(P\cup Q,P\cup Q)+2u(P\cup Q,U)\leqslant u(B_{1}^{-},B_{1}^{-})+2u(B_{1}^{-},B_{1}^{+}),

in order to prove (8.8), it is enough to show that

(8.9) v(QΩ)w(S,S)+2w(S,TU)+u(PQ,PQ)+2u(PQ,U)+O(σε12s)+O(δ~2s).v(Q_{\Omega})\leqslant w(S,S)+2w(S,T\cup U)+u(P\cup Q,P\cup Q)+2u(P\cup Q,U)+O(\sigma\varepsilon^{1-2s})+O(\tilde{\delta}^{-2s}).

From the construction of vv, note that

v(S,S)=w(S,S),v(S,TU)=w(S,TU),\displaystyle v(S,S)=w(S,S),\quad v(S,T\cup U)=w(S,T\cup U),
and v(PQ,PQ)=u(PQ,PQ).\displaystyle{\mbox{and }}\quad v(P\cup Q,P\cup Q)=u(P\cup Q,P\cup Q).

Recalling (8.7), we find that

v(Ω,Ω)\displaystyle v(\Omega,\Omega) =w(S,S)+2v(S,PQR)+u(PQ,PQ)+2v(PQ,R)+v(R,R).\displaystyle=w(S,S)+2v(S,P\cup Q\cup R)+u(P\cup Q,P\cup Q)+2v(P\cup Q,R)+v(R,R).

and

v(Ω,Ωc)\displaystyle v(\Omega,\Omega^{c}) =w(S,TU)+v(PQ,T)+v(PQ,U)+v(R,TU).\displaystyle=w(S,T\cup U)+v(P\cup Q,T)+v(P\cup Q,U)+v(R,T\cup U).

Therefore,

(8.10) v(QΩ)=[w(S,S)+2w(S,TU)]+[u(PQ,PQ)+2u(PQ,U)]+Ev(Q_{\Omega})=[w(S,S)+2w(S,T\cup U)]+[u(P\cup Q,P\cup Q)+2u(P\cup Q,U)]+E

where

E\displaystyle E :=2v(S,PQR)+2v(PQ,RT)+v(R,R)+2v(R,TU)\displaystyle:=2v(S,P\cup Q\cup R)+2v(P\cup Q,R\cup T)+v(R,R)+2v(R,T\cup U)
+2(v(PQ,U)u(PQ,U)).\displaystyle\quad+2\big{(}v(P\cup Q,U)-u(P\cup Q,U)\big{)}.

Step 4. (Error estimates). We will show that there exists C>0C>0 such that

(8.11) |E|C(σε12s+δ~2s).|E|\leqslant C\big{(}\sigma\varepsilon^{1-2s}+\tilde{\delta}^{-2s}\big{)}.

Notice that (8.11), together with (8.10), gives the desired result in (8.9) (and then in (8.8)). Hence, we now focus on the proof of (8.11).

First note that

QR=D~iεD~δ~D~D~δ~Q\cup R=\widetilde{D}_{i\varepsilon}\setminus\widetilde{D}_{\tilde{\delta}}\subset\widetilde{D}\setminus\widetilde{D}_{\tilde{\delta}}

and, since DδD~δ~D~(i+1)εD_{\delta}\subset\widetilde{D}_{\tilde{\delta}}\subset\widetilde{D}_{(i+1)\varepsilon},

RSTU=(D~(i+1)ε)c(D~δ~)c(Dδ)c.R\cup S\cup T\cup U=(\widetilde{D}_{(i+1)\varepsilon})^{c}\subset(\widetilde{D}_{\tilde{\delta}})^{c}\subset(D_{\delta})^{c}.

Therefore, (8.4) implies that

(8.12) w(QR,RSTU)σε12s.w(Q\cup R,R\cup S\cup T\cup U)\leqslant\sigma\varepsilon^{1-2s}.

Following the proof of [SV-gamma, Proposition 4.1], we use (8.4), (8.6), and (8.12) to find that

(8.13) v(P,STR)+v(Q,STR)+v(STU,R)+v(R,R)C(σε12s+δ~2s).v(P,S\cup T\cup R)+v(Q,S\cup T\cup R)+v(S\cup T\cup U,R)+v(R,R)\leqslant C\big{(}\sigma\varepsilon^{1-2s}+\tilde{\delta}^{-2s}\big{)}.

We are left to show that

|v(PQ,U)u(PQ,U)|Cδ~2s.|v(P\cup Q,U)-u(P\cup Q,U)|\leqslant C\tilde{\delta}^{-2s}.

For this, observe that

PQU|xy|>δ~21|xy|12s𝑑y𝑑xCδ~2+r12s𝑑rCδ~2s.\int_{P\cup Q\cup U}\int_{|x-y|>\frac{\tilde{\delta}}{2}}\frac{1}{|x-y|^{1-2s}}dy\,dx\leqslant C\int_{\frac{\tilde{\delta}}{2}}^{+\infty}r^{-1-2s}\,dr\leqslant C\tilde{\delta}^{-2s}.

Therefore, recalling that w(y)=g(0)w(y)=g(0) and u(y)=g(y)u(y)=g(y) for yUy\in U,

|v(PQ,U)u(PQ,U)|\displaystyle|v(P\cup Q,U)-u(P\cup Q,U)| δ~200δ~2||u(x)g(0)|2|u(x)g(y)|2||xy|1+2s𝑑y𝑑x+Cδ~2s.\displaystyle\leqslant\int_{-\frac{\tilde{\delta}}{2}}^{0}\int_{0}^{\frac{\tilde{\delta}}{2}}\frac{\big{|}|u(x)-g(0)|^{2}-|u(x)-g(y)|^{2}\big{|}}{|x-y|^{1+2s}}\,dy\,dx+C\tilde{\delta}^{-2s}.

Since gg is Lipschitz continuous and |u||u|, |g|1|g|\leqslant 1,

||u(x)g(0)|2|u(x)g(y)|2|\displaystyle\big{|}|u(x)-g(0)|^{2}-|u(x)-g(y)|^{2}\big{|} C|g(0)g(y)|C|y|.\displaystyle\leqslant C|g(0)-g(y)|\leqslant C|y|.

Therefore,

|v(PQ,U)u(PQ,U)|Cδ~200δ~2|y||xy|1+2s𝑑y𝑑x+Cδ~2s\displaystyle|v(P\cup Q,U)-u(P\cup Q,U)|\leqslant C\int_{-\frac{\tilde{\delta}}{2}}^{0}\int_{0}^{\frac{\tilde{\delta}}{2}}\frac{|y|}{|x-y|^{1+2s}}\,dy\,dx+C\tilde{\delta}^{-2s}
C0δ~2y12s𝑑y+Cδ~2sCδ~22s+Cδ~2sCδ~2s.\displaystyle\qquad\qquad\leqslant C\int_{0}^{\frac{\tilde{\delta}}{2}}y^{1-2s}\,dy+C\tilde{\delta}^{-2s}\leqslant C\tilde{\delta}^{2-2s}+C\tilde{\delta}^{-2s}\leqslant C\tilde{\delta}^{-2s}.

Together with (8.13), this proves (8.11).

Conclusion. Regarding the potential energy, we use (8.7) to write

ΩW(v)𝑑x=PQW(u)𝑑x+SW(w)𝑑x+RW(v)𝑑x.\int_{\Omega}W(v)\,dx=\int_{P\cup Q}W(u)\,dx+\int_{S}W(w)\,dx+\int_{R}W(v)\,dx.

In RR, observe that

W(v)W(w)+C|vw|W(w)+C|uw|.W(v)\leqslant W(w)+C|v-w|\leqslant W(w)+C|u-w|.

Therefore, with (8.6),

ΩW(v)𝑑xPQW(u)𝑑x+SRW(w)𝑑x+CR|uw|𝑑xDW(u)𝑑x+ΩDδW(w)𝑑x+Cσε.\begin{split}\int_{\Omega}W(v)\,dx&\leqslant\int_{P\cup Q}W(u)\,dx+\int_{S\cup R}W(w)\,dx+C\int_{R}|u-w|\,dx\\ &\leqslant\int_{D}W(u)\,dx+\int_{\Omega\setminus D_{\delta}}W(w)\,dx+C\sigma\varepsilon.\end{split}

From this and (8.8), we obtain that

ε(1)(v,Ω)\displaystyle\mathcal{F}_{\varepsilon}^{(1)}(v,\Omega) =ε2s1v(QΩ)+1εΩW(v)𝑑x\displaystyle=\varepsilon^{2s-1}v(Q_{\Omega})+\frac{1}{\varepsilon}\int_{\Omega}W(v)\,dx
ε2s1(w(QΩ)w(QDδ)+u(B1,B1)+2u(B1,B1+))\displaystyle\leqslant\varepsilon^{2s-1}\Big{(}w(Q_{\Omega})-w(Q_{D_{\delta}})+u(B_{1}^{-},B_{1}^{-})+2u(B_{1}^{-},B_{1}^{+})\Big{)}
+1ε(DW(u)𝑑x+ΩW(w)𝑑xDδW(w)𝑑x)+C(σ+δ~2sε2s1)\displaystyle\quad+\frac{1}{\varepsilon}\left(\int_{D}W(u)\,dx+\int_{\Omega}W(w)\,dx-\int_{D_{\delta}}W(w)\,dx\right)+C\big{(}\sigma+\tilde{\delta}^{-2s}\varepsilon^{2s-1}\big{)}
=ε(1)(w,Ω)ε(1)(w,Dδ)+Iε(u,B1,Ω)+C(σ+δ~2sε2s1).\displaystyle=\mathcal{F}_{\varepsilon}^{(1)}(w,\Omega)-\mathcal{F}_{\varepsilon}^{(1)}(w,D_{\delta})+I_{\varepsilon}(u,B_{1},\Omega)+C\big{(}\sigma+\tilde{\delta}^{-2s}\varepsilon^{2s-1}\big{)}.

Therefore,

lim supk+ε(1)(v,Ω)lim supk+[ε(1)(w,Ω)ε(1)(w,Dδ)+Iε(u,B1,Ω)]+Cσ.\limsup_{k\to+\infty}\mathcal{F}_{\varepsilon}^{(1)}(v,\Omega)\leqslant\limsup_{k\to+\infty}\Big{[}\mathcal{F}_{\varepsilon}^{(1)}(w,\Omega)-\mathcal{F}_{\varepsilon}^{(1)}(w,D_{\delta})+I_{\varepsilon}(u,B_{1},\Omega)\Big{]}+C\sigma.

Since σ>0\sigma>0 was arbitrary, the proof Proposition 8.1 is complete. ∎

8.2. Contribution near the boundary

We now establish an energy estimate near Ω\partial\Omega.

Proposition 8.2.

Assume that |g|<1|g|<1 on Ω\partial\Omega and let κ(0,|Ω|2)\kappa\in\left(0,\frac{|\Omega|}{2}\right). Let uε:[1,1]u_{\varepsilon}:\mathbb{R}\to[-1,1] be such that

(8.15) uε(x)=g(x) for any xΩ¯,\displaystyle u_{\varepsilon}(x)=g(x){\mbox{ for any }}x\in\mathbb{R}\setminus\overline{\Omega},
limε0uε(x)=±1 for any xBρo(x¯)Ω,\displaystyle\lim_{\varepsilon\searrow 0}u_{\varepsilon}(x)=\pm 1{\mbox{ for any }}x\in B_{\rho_{o}}(\bar{x})\cap\Omega,
(8.16) and either uε(x)g(x¯)0u_{\varepsilon}(x)\geqslant g(\bar{x})\geqslant 0 or uε(x)g(x¯)0u_{\varepsilon}(x)\leqslant g(\bar{x})\leqslant 0, for all xBκ(x¯)Ωx\in B_{\kappa}(\bar{x})\cap\Omega,

for some ρo(0,κ]\rho_{o}\in(0,\kappa] and some x¯Ω\bar{x}\in\partial\Omega.

Then,

lim infρ0lim infε0Iε(uε,Bρ(x¯),Ω)Ψ(±1,g(x¯)).\liminf_{\rho\searrow 0}\,\liminf_{\varepsilon\searrow 0}I_{\varepsilon}(u_{\varepsilon},B_{\rho}(\bar{x}),\Omega)\geqslant\Psi(\pm 1,g(\bar{x})).
Proof.

Without loss of generality, we assume that

0=x¯ΩandBρo(x¯)Ω=Bρo=(ρo,0)0=\bar{x}\in\partial\Omega\qquad\hbox{and}\qquad B_{\rho_{o}}(\bar{x})\cap\Omega=B_{\rho_{o}}^{-}=(-\rho_{o},0)

and that

(8.17) limε0uε(x)=1 for any xBρo.\lim_{\varepsilon\searrow 0}u_{\varepsilon}(x)=-1{\mbox{ for any }}x\in B_{\rho_{o}}^{-}.

Consequently, we deduce from (8.16) that

(8.18) uε(x)g(0)0 for any xBκBρo.{u_{\varepsilon}(x)\leqslant g(0)\leqslant 0{\mbox{ for any }}x\in B_{\kappa}^{-}\supset B_{\rho_{o}}^{-}.}

We then want to show that

lim infρ0lim infε0Iε(uε,Bρ,Bρ)Ψ(1,g(0)).\liminf_{\rho\searrow 0}\,\liminf_{\varepsilon\searrow 0}I_{\varepsilon}(u_{\varepsilon},B_{\rho}^{-},B_{\rho})\geqslant\Psi(-1,g(0)).

For this, we set

𝒥ρ,ε\displaystyle{\mathcal{J}}_{\rho,\varepsilon} :=Iε(uε,Bρ,Bρ)\displaystyle:=I_{\varepsilon}(u_{\varepsilon},B_{\rho}^{-},B_{\rho})
𝒥ρ\displaystyle{\mathcal{J}}_{\rho} :=lim infε0𝒥ρ,ε=lim infε0Iε(uε,Bρ,Bρ)\displaystyle:=\liminf_{\varepsilon\searrow 0}{\mathcal{J}}_{\rho,\varepsilon}=\liminf_{\varepsilon\searrow 0}I_{\varepsilon}(u_{\varepsilon},B_{\rho}^{-},B_{\rho})
and 𝒥\displaystyle{\mbox{and }}\quad{\mathcal{J}} :=lim infρ0𝒥ρ=lim infρ0lim infε0Iε(uε,Bρ,Bρ).\displaystyle:=\liminf_{\rho\searrow 0}{\mathcal{J}}_{\rho}=\liminf_{\rho\searrow 0}\liminf_{\varepsilon\searrow 0}I_{\varepsilon}(u_{\varepsilon},B_{\rho}^{-},B_{\rho}).

We take a sequence ρi>0\rho_{i}>0 which is infinitesimal as i+i\to+\infty and such that

(8.19) 𝒥=limi+𝒥ρi.{\mathcal{J}}=\lim_{i\to+\infty}{\mathcal{J}}_{\rho_{i}}.

Then, for any fixed ii\in\mathbb{N}, we take a sequence εk,i\varepsilon_{k,i} which is infinitesimal as k+k\to+\infty and such that

(8.20) 𝒥ρi=limk+𝒥ρi,εk,i.{\mathcal{J}}_{\rho_{i}}=\lim_{k\to+\infty}{\mathcal{J}}_{\rho_{i},\varepsilon_{k,i}}.

We point out that, up to extracting a subsequence, we can suppose that

(8.21) εk,iρik.\varepsilon_{k,i}\leqslant\frac{\rho_{i}}{k}.

Since we need to extract subsequences in a delicate and appropriate way, given mm\in\mathbb{N} (to be taken large in the sequel), we use (8.19) to find i0(m)i_{0}(m) such that if ii0(m)i\geqslant i_{0}(m) then

|𝒥𝒥ρi|em2.|{\mathcal{J}}-{\mathcal{J}}_{\rho_{i}}|\leqslant\frac{e^{-m}}{2}.

Then, fixed ii\in\mathbb{N}, we use (8.20) to find k0(i,m)k_{0}(i,m) such that for any kk0(i,m)k\geqslant k_{0}(i,m) we have that

|𝒥ρi𝒥ρi,εk,i|em2.|{\mathcal{J}}_{\rho_{i}}-{\mathcal{J}}_{\rho_{i},\varepsilon_{k,i}}|\leqslant\frac{e^{-m}}{2}.

In particular, if ii0(m)i\geqslant i_{0}(m) and kk0(i,m)k\geqslant k_{0}(i,m),

(8.22) |𝒥𝒥ρi,εk,i|em.|{\mathcal{J}}-{\mathcal{J}}_{\rho_{i},\varepsilon_{k,i}}|\leqslant{e^{-m}}.

Now, we define

(8.23) uk,i(x):=uεk,i(ρix)andwk,i(x):=wεk,i(ρix)=w0(ρixεk,i),u_{k,i}(x):=u_{\varepsilon_{k,i}}({{{\rho}}}_{i}x)\qquad{\mbox{and}}\qquad w_{k,i}(x):=w_{\varepsilon_{k,i}}({{{\rho}}}_{i}x)=w_{0}\left(\frac{{{{\rho}}}_{i}x}{\varepsilon_{k,i}}\right),

where w0(x):=w0(x;1,g(0))w_{0}(x):=w_{0}(x;-1,g(0)) is given by Theorem 1.3 with γ:=g(0)(1,1)\gamma:=g(0)\in(-1,1) (recall the notation in (7.2)). In particular,

(8.24) if x>0x>0,  then wk,i(x)=g(0)w_{k,i}(x)=g(0).

Since w0(x)1w_{0}(x)\to-1 uniformly as xx\to-\infty, by the monotonicity of w0w_{0}, we have that

limk+supxρi|wk,i(x)+1|=limk+supxρiw0(ρixεk,i)+1=limk+w0(ρi2εk,i)+1=0.\lim_{k\to+\infty}\sup_{x\leqslant-\rho_{i}}|w_{k,i}(x)+1|=\lim_{k\to+\infty}\sup_{x\leqslant-\rho_{i}}w_{0}\left(\frac{\rho_{i}x}{\varepsilon_{k,i}}\right)+1=\lim_{k\to+\infty}w_{0}\left(-\frac{\rho_{i}^{2}}{\varepsilon_{k,i}}\right)+1=0.

Hence, for any ii, mm\in\mathbb{N}, there exists k1(i,m)[m,+)k_{1}(i,m)\in\mathbb{N}\cap[m,+\infty) such that

(8.25) supxρi|wk,i(x)+1|emfor anykk1(i,m).\sup_{x\leqslant-\rho_{i}}|w_{k,i}(x)+1|\leqslant e^{-m}\quad\hbox{for any}\leavevmode\nobreak\ k\geqslant k_{1}(i,m).

Now, let i1i_{1}\in\mathbb{N} be such that ρi<1\rho_{i}<1 for all ii1i\geqslant i_{1}. If x(0,ρo)x\in(0,\rho_{o}), we have that

(8.26) ρix(0,ρo)for allii1,\rho_{i}x\in(0,\rho_{o})\quad\hbox{for all}\leavevmode\nobreak\ i\geqslant i_{1},

so, by (8.15),

uk,i(x)=uεk,i(ρix)=g(ρix).u_{k,i}(x)=u_{\varepsilon_{k,i}}({{{\rho}}}_{i}x)=g({{{\rho}}}_{i}x).

Since gg is Lipschitz continuous, there exists i2(m)[m,+)i_{2}(m)\in\mathbb{N}\cap[m,+\infty) such that, for any x(0,ρo)x\in(0,\rho_{o}) and any ii2(m)i\geqslant i_{2}(m),

|g(ρix)g(0)|em.|g({{{\rho}}}_{i}x)-g(0)|\leqslant e^{-m}.

Therefore, if i3(m):=i1+i2(m)i_{3}(m):=i_{1}+i_{2}(m), we have that, for any ii3(m)i\geqslant i_{3}(m), any kk\in\mathbb{N} and any x(0,ρo)x\in(0,\rho_{o}),

(8.27) |uk,i(x)g(0)|em.|u_{k,i}(x)-g(0)|\leqslant e^{-m}.

Now we define

(8.28) im:=i0(m)+i3(m)=i0(m)+i1+i2(m)and km:=k0(im,m)+k1(im,m).\begin{split}&{i_{m}}:=i_{0}(m)+i_{3}(m)=i_{0}(m)+i_{1}+i_{2}(m)\\ {\mbox{and }}\quad&{k_{m}}:=k_{0}(i_{m},m)+k_{1}(i_{m},m).\end{split}

Notice that both km{k_{m}} and im{i_{m}} diverge as m+m\to+\infty. We also set

εm:=εkm,im,wm(x):=wkm,im(x), and um(x):=ukm,im(x).\varepsilon_{m}:=\varepsilon_{{k_{m}},{i_{m}}},\;\qquad w_{m}(x):=w_{{k_{m}},{i_{m}}}(x),\;\qquad{\mbox{ and }}\;\qquad u_{m}(x):=u_{{k_{m}},{i_{m}}}(x).

Note that εm/ρim\varepsilon_{m}/\rho_{i_{m}} is infinitesimal as m+m\to+\infty, thanks to (8.21).

From (8.27), we have that that umg(0)u_{m}\to g(0) in (0,ρo)(0,\rho_{o}), as m+m\to+\infty. An obvious consequence of (8.24) is that also wmg(0)w_{m}\to g(0) in (0,ρo)(0,\rho_{o}), as m+m\to+\infty. From (8.25), we have that wm1w_{m}\to-1 in (ρo,0)(-\rho_{o},0), as m+m\to+\infty. On the other hand, if x(ρo,0)x\in(-\rho_{o},0), then by (8.26) and (8.28),

ρix(ρo,0)for all iim,\rho_{i}x\in(-\rho_{o},0)\quad\hbox{for all }i\geqslant i_{m},

and so, thanks to (8.17),

limk+uk,i(x)=limk+uεk,i(ρix)=1.\lim_{k\to+\infty}u_{k,i}(x)=\lim_{k\to+\infty}u_{\varepsilon_{k,i}}({{{\rho}}}_{i}x)=-1.

That is, for any x(ρo,0)x\in(-\rho_{o},0) and any mm\in\mathbb{N}, there exists k2(m,x)[m,+)k_{2}(m,x)\in\mathbb{N}\cap[m,+\infty) such that, for any ii1i\geqslant i_{1} and any kk2(m,x)k\geqslant k_{2}(m,x),

|uk,i(x)+1|em.|u_{k,i}(x)+1|\leqslant e^{-m}.

In particular, taking mm large enough to guarantee that kmk2(m,x)k_{m}\geqslant k_{2}(m,x), it holds that

|um(x)+1|em|u_{m}(x)+1|\leqslant e^{-m}

and um(x)1u_{m}(x)\to-1 as m+m\to+\infty. Therefore, by collecting these pieces, we find that umwm0u_{m}-w_{m}\to 0 a.e. in (ρo,ρo)(-\rho_{o},\rho_{o}), as m+m\to+\infty.

Without loss of generality, let us take now take ρo=1\rho_{o}=1. Fix δ(0,14)\delta\in\left(0,\frac{1}{4}\right) and set

D\displaystyle D :=B1Ωρim=B1\displaystyle:=B_{1}\cap\frac{\Omega}{\rho_{{i_{m}}}}=B_{1}^{-}
and Dδ\displaystyle{\mbox{and }}\quad D_{\delta} :={xD s.t. dDΩ(x)>δ}=B1δ.\displaystyle:=\left\{x\in D{\mbox{ s.t. }}d_{\partial D\setminus\partial\Omega}(x)>\delta\right\}=B_{1-\delta}^{-}.

By (8.18), we have that umg(0)=γu_{m}\leqslant g(0)=\gamma in B1B_{1}^{-}. We are now in a position to apply Proposition 8.1 to guarantee the existence of vmγv_{m}\leqslant\gamma such that

vm(x)={um(x) if xDδ,wm(x) if xDc,v_{m}(x)=\begin{cases}u_{m}(x)&\hbox{ if }x\in D_{\delta},\\ w_{m}(x)&\hbox{ if }x\in D^{c},\end{cases}

and satisfying

(8.29) lim supm+ε~m(1)(vm,Ω)lim supm+[ε~m(1)(wm,Ω)ε~m(1)(wm,Dδ)+Iε~m(um,D,Ω)]\limsup_{m\to+\infty}{\mathcal{F}}^{(1)}_{\tilde{\varepsilon}_{m}}(v_{m},\Omega)\leqslant\limsup_{m\to+\infty}\left[{\mathcal{F}}^{(1)}_{\tilde{\varepsilon}_{m}}(w_{m},\Omega)-{\mathcal{F}}^{(1)}_{\tilde{\varepsilon}_{m}}(w_{m},D_{\delta})+I_{\tilde{\varepsilon}_{m}}(u_{m},D,\Omega)\right]

where ε~m:=εm/ρim\tilde{\varepsilon}_{m}:=\varepsilon_{m}/\rho_{i_{m}}. We recall that ε~m\tilde{\varepsilon}_{m} is infinitesimal as m+m\to+\infty.

Let s(12,1)s\in\left(\frac{1}{2},1\right). Since vmγv_{m}\leqslant\gamma and vm=wmv_{m}=w_{m} outside Ω\Omega, in particular outside Ω\Omega\cap\mathbb{R}^{-}, we are in the position of applying (5.2) (with A:=ΩA:=\Omega and B:=B:=\mathbb{R}^{-}), obtaining that

ε~m(1)(wm,Ω)ε~m(1)(vm,Ω)=ε~m(1)(wm,)ε(1)(vm,).\mathcal{F}_{\tilde{\varepsilon}_{m}}^{(1)}(w_{m},\Omega)-\mathcal{F}_{\tilde{\varepsilon}_{m}}^{(1)}(v_{m},\Omega)=\mathcal{F}_{\tilde{\varepsilon}_{m}}^{(1)}(w_{m},\mathbb{R}^{-})-\mathcal{F}_{\varepsilon}^{(1)}(v_{m},\mathbb{R}^{-}).

Thus, rescaling as in (5.4) (with A:=A:=\mathbb{R}^{-}) and using the fact that w0w_{0} is a global minimizer of 𝒢s\mathcal{G}_{s}, we find that

ε~m(1)(wm,Ω)ε~m(1)(vm,Ω)=𝒢s(w0,)𝒢s(vm(ε~m()),)0.\mathcal{F}_{\tilde{\varepsilon}_{m}}^{(1)}(w_{m},\Omega)-\mathcal{F}_{\tilde{\varepsilon}_{m}}^{(1)}(v_{m},\Omega)=\mathcal{G}_{s}(w_{0},\mathbb{R}^{-})-\mathcal{G}_{s}(v_{m}(\tilde{\varepsilon}_{m}(\cdot)),\mathbb{R}^{-})\leqslant 0.

Consequently with (8.29) and scaling as in (5.3)-(5.4), we have that

0\displaystyle 0 lim supm+[ε~m(1)(wm,Dδ)+Iε~m(um,D,Ω)]\displaystyle\leqslant\limsup_{m\to+\infty}\left[-{\mathcal{F}}^{(1)}_{\tilde{\varepsilon}_{m}}(w_{m},D_{\delta})+I_{\tilde{\varepsilon}_{m}}(u_{m},D,\Omega)\right]
=lim supm+[ε~m(1)(wm,Dδ)+Iε~m(um,B1,B1)]\displaystyle=\limsup_{m\to+\infty}\left[-{\mathcal{F}}^{(1)}_{\tilde{\varepsilon}_{m}}(w_{m},D_{\delta})+I_{\tilde{\varepsilon}_{m}}(u_{m},B_{1}^{-},B_{1})\right]
lim supm+[𝒢s(w0,Dδ/ε~m)+Iεm(uεm,Bρim,Bρim)].\displaystyle\leqslant\limsup_{m\to+\infty}\left[-\mathcal{G}_{s}(w_{0},D_{\delta}/\tilde{\varepsilon}_{m})+I_{\varepsilon_{m}}(u_{\varepsilon_{m}},B_{\rho_{i_{m}}}^{-},B_{\rho_{i_{m}}})\right].

Since δ>0\delta>0 is arbitrary, we obtain, recalling also (7.1) and (7.3), that

(8.30) lim supm+Iεm(uεm,Bρim,Bρim)lim infm+𝒢s(w0,Bε~m1)=𝒢s(w0)=Ψ(1,g(0)).\limsup_{m\to+\infty}I_{\varepsilon_{m}}(u_{\varepsilon_{m}},B_{\rho_{i_{m}}}^{-},B_{\rho_{i_{m}}})\geqslant\liminf_{m\to+\infty}\mathcal{G}_{s}(w_{0},B_{\tilde{\varepsilon}_{m}^{-1}}^{-})=\mathcal{G}_{s}(w_{0})=\Psi(-1,g(0)).

For the case s=12s=\frac{1}{2}, using here that w0w_{0} is a local minimizer in Bε~m1B_{\tilde{\varepsilon}_{m}^{-1}}^{-}, we similarly estimate (since vm=wmv_{m}=w_{m} in (B1)c(B_{1}^{-})^{c}),

ε~m(1)(wm,Ω)ε(1)(vm,Ω)=ε~m(1)(wm,B1)ε(1)(vm,B1)\displaystyle\mathcal{F}_{\tilde{\varepsilon}_{m}}^{(1)}(w_{m},\Omega)-\mathcal{F}_{\varepsilon}^{(1)}(v_{m},\Omega)=\mathcal{F}_{\tilde{\varepsilon}_{m}}^{(1)}(w_{m},B_{1}^{-})-\mathcal{F}_{\varepsilon}^{(1)}(v_{m},B_{1}^{-})
=1|lnε~m|[𝒢s(w0,Bε~m1)𝒢s((vm(ε~m()),Bε~m1)]0,\displaystyle\qquad\qquad=\frac{1}{|\ln\tilde{\varepsilon}_{m}|}\left[\mathcal{G}_{s}(w_{0},B_{\tilde{\varepsilon}_{m}^{-1}}^{-})-\mathcal{G}_{s}((v_{m}(\tilde{\varepsilon}_{m}(\cdot)),B_{\tilde{\varepsilon}_{m}^{-1}}^{-})\right]\leqslant 0,

so that

0\displaystyle 0 lim supm+[ε~m(1)(wm,Dδ)+Iε~m(um,D,Ω)]\displaystyle\leqslant\limsup_{m\to+\infty}\left[-{\mathcal{F}}^{(1)}_{\tilde{\varepsilon}_{m}}(w_{m},D_{\delta})+I_{\tilde{\varepsilon}_{m}}(u_{m},D,\Omega)\right]
=lim supm+[ε~m(1)(wm,Dδ)+Iε~m(um,B1,B1)]\displaystyle=\limsup_{m\to+\infty}\left[-{\mathcal{F}}^{(1)}_{\tilde{\varepsilon}_{m}}(w_{m},D_{\delta})+I_{\tilde{\varepsilon}_{m}}(u_{m},B_{1}^{-},B_{1})\right]
lim supm+[1|lnε~m|𝒢s(w0,Dδ/ε~m)+Iεm(uεm,Bρim,Bρim)].\displaystyle\leqslant\limsup_{m\to+\infty}\left[-\frac{1}{|\ln\tilde{\varepsilon}_{m}|}\mathcal{G}_{s}(w_{0},D_{\delta}/\tilde{\varepsilon}_{m})+I_{\varepsilon_{m}}(u_{\varepsilon_{m}},B_{\rho_{i_{m}}}^{-},B_{\rho_{i_{m}}})\right].

Consequently,

(8.31) lim supm+Iεm(uεm,Bρim,Bρim)lim infm+1|lnε~m|𝒢s(w0,Bε~m1)=Ψ(1,g(0)).\limsup_{m\to+\infty}I_{\varepsilon_{m}}(u_{\varepsilon_{m}},B_{\rho_{i_{m}}}^{-},B_{\rho_{i_{m}}})\geqslant\liminf_{m\to+\infty}\frac{1}{|\ln\tilde{\varepsilon}_{m}|}\mathcal{G}_{s}(w_{0},B_{\tilde{\varepsilon}_{m}^{-1}}^{-})=\Psi(-1,g(0)).

Now, for all s[12,1)s\in\left[\frac{1}{2},1\right), we have from (8.30) and (8.31) that

(8.32) Ψ(1,g(0))lim supm+Iεm(uεm,Bρim,Bρim)=lim supm+𝒥ρim,εkm,im.\begin{split}\Psi(-1,g(0))\leqslant\limsup_{m\to+\infty}I_{\varepsilon_{m}}(u_{\varepsilon_{m}},B_{\rho_{i_{m}}}^{-},B_{\rho_{i_{m}}})=\limsup_{m\to+\infty}{\mathcal{J}}_{\rho_{i_{m}},\varepsilon_{k_{m},i_{m}}}.\end{split}

Recalling (8.28), we see that imi0(m)i_{m}\geqslant i_{0}(m) and kmk0(im,m)k_{m}\geqslant k_{0}(i_{m},m), thus we can use (8.22) to find that 𝒥ρim,εkm,im𝒥+em{\mathcal{J}}_{\rho_{i_{m}},\varepsilon_{k_{m},i_{m}}}\leqslant{\mathcal{J}}+{e^{-m}}. Hence, we infer from (8.32) that

Ψ(1,g(0))lim supm+(𝒥+em)=𝒥,\Psi(-1,g(0))\leqslant\limsup_{m\to+\infty}(\mathcal{J}+e^{-m})={\mathcal{J}},

as desired. ∎

8.3. Proof of the lim inf\liminf inequality

We recall that XX is the space of all the measurable functions u:nu:\mathbb{R}^{n}\to\mathbb{R} such that the restriction of uu to Ω\Omega belongs to L1(Ω)L^{1}(\Omega). Moreover, XX is endowed with the metric of L1(Ω)L^{1}(\Omega), as made clear in (1.14). Also, the spaces XgX_{g} and YκY_{\kappa} are defined in (1.15) and (1.7), respectively.

Proposition 8.3.

Assume that |g|<1|g|<1 on Ω\partial\Omega and let κ(0,|Ω|2)\kappa\in\left(0,\frac{|\Omega|}{2}\right). Let uju_{j} be such that uju¯u_{j}\to\bar{u} in XX and εj0\varepsilon_{j}\searrow 0 as j+j\to+\infty.

Then,

lim infj+εj(1)(uj)(1)(u¯)\liminf_{j\to+\infty}{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(u_{j})\geqslant{\mathcal{F}}^{(1)}(\bar{u})

where (1)\mathcal{F}^{(1)} is given by

(1)(u):={cPer(E,Ω)+ΩΨ(u(x),g(x))𝑑0(x) if XYκu=χEχEc a.e. in Ω, for some En,+ otherwise.{\mathcal{F}}^{(1)}(u):=\begin{cases}c_{\star}\operatorname{Per}\,(E,\Omega)+\displaystyle\int_{\partial\Omega}\Psi(u(x),g(x))\,d{\mathcal{H}}^{0}(x)&\begin{matrix}{\mbox{ if\leavevmode\nobreak\ $X\cap Y_{\kappa}\ni u=\chi_{E}-\chi_{E^{c}}$ a.e. in\leavevmode\nobreak\ $\Omega$,}}\\ {\mbox{ for some\leavevmode\nobreak\ $E\subset\mathbb{R}^{n}$,}}\end{matrix}\\ +\infty&{\mbox{ otherwise.}}\end{cases}
Proof.

Notice that we can assume that ujXgYκu_{j}\in X_{g}\cap Y_{\kappa}, otherwise εj(1)(uj)=+\mathcal{F}_{\varepsilon_{j}}^{(1)}(u_{j})=+\infty (recall the setting in Section 3) and we are done.

Also, in light of [SV-gamma, Proposition 3.3], we can assume that u¯=χEχEc\bar{u}=\chi_{E}-\chi_{E^{c}} a.e. in Ω\Omega, with EE of finite perimeter. In particular, u¯\bar{u} can be defined along Ω\partial\Omega in the trace sense (see e.g. [Giusti]). Moreover, since ujYκu_{j}\in Y_{\kappa} and uju¯u_{j}\to\bar{u} in XX, it must be that u¯Yκ\bar{u}\in Y_{\kappa}.

For clarity in the proof, we assume that Ω=(x¯1,x¯2)\Omega=(\bar{x}_{1},\bar{x}_{2}). Consider a covering of small intervals {Bρi(x¯1),Bρi(x¯2)}\{B_{\rho_{i}}(\bar{x}_{1}),B_{\rho_{i}}(\bar{x}_{2})\} of Ω={x¯1,x¯2}\partial\Omega=\{\bar{x}_{1},\bar{x}_{2}\}, with ρi\rho_{i} infinitesimal as i+i\to+\infty, and with

(8.33) either u¯=1\bar{u}=-1 or u¯=1\bar{u}=1 in Bρi(x¯k)ΩB_{{\rho}_{i}}(\bar{x}_{k})\cap\Omega, with k=1,2k=1,2.

We can satisfy (8.33) since EE\subset\mathbb{R} has finite perimeter.

Now we fix an interval ΩΩ\Omega^{\prime}\Subset\Omega. Then, for large ii, the balls Bρi(x±)B_{{{{\rho}}}_{i}}(x_{\pm}) lie outside Ω\Omega^{\prime}, and, by inspection, one sees that

(8.34) εj(1)(uj,Ω)\displaystyle{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(u_{j},\Omega) εj(1)(uj,Ω)+Iεj(uj,Bρi(x¯1),Ω)+Iεj(uj,Bρi(x¯2),Ω).\displaystyle\geqslant{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(u_{j},\Omega^{\prime})+I_{\varepsilon_{j}}(u_{j},B_{{\rho}_{i}}(\bar{x}_{1}),\Omega)+I_{\varepsilon_{j}}(u_{j},B_{{\rho}_{i}}(\bar{x}_{2}),\Omega).

By [SV-gamma, Proposition 4.5], it holds that

lim infj+εj(1)(uj,Ω)cPer(E,Ω).\liminf_{j\to+\infty}{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(u_{j},\Omega^{\prime})\geqslant c_{\star}\operatorname{Per}\,(E,\Omega^{\prime}).

With this and Proposition 8.2, we find in (8.34) that

lim infi+lim infj+εj(1)(uj,Ω)cPer(E,Ω)+Ψ(u¯(x¯1),g(x¯1))+Ψ(u¯(x¯2),g(x¯2)).\liminf_{i\to+\infty}\liminf_{j\to+\infty}{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(u_{j},\Omega)\geqslant c_{\star}\operatorname{Per}\,(E,\Omega^{\prime})+\Psi(\bar{u}(\bar{x}_{1}),g(\bar{x}_{1}))+\Psi(\bar{u}(\bar{x}_{2}),g(\bar{x}_{2})).

We remark that Proposition 8.2 can be used in this framework, since (8.15) and (8.16) are consequences of the fact that ujXgYκu_{j}\in X_{g}\cap Y_{\kappa}, while (8.15) follows from (8.33). The desired result now follows by taking Ω\Omega^{\prime} arbitrarily close to Ω\Omega and noticing that

ΩΨ(u¯(x),g(x))𝑑0(x)=Ψ(u¯(x¯1),g(x¯1))+Ψ(u¯(x¯2),g(x¯2)).\int_{\partial\Omega}\Psi(\bar{u}(x),g(x))\,d{\mathcal{H}}^{0}(x)=\Psi(\bar{u}(\bar{x}_{1}),g(\bar{x}_{1}))+\Psi(\bar{u}(\bar{x}_{2}),g(\bar{x}_{2})).\qed

9. Bounding the energy from above for s[12,1)s\in\left[\frac{1}{2},1\right)

In this section, we establish the lim sup\limsup-inequality in Theorem 1.6 by constructing a recovery sequence corresponding to some infinitesimal sequence εj0\varepsilon_{j}\searrow 0. To this aim, we let ε:=εj\varepsilon:=\varepsilon_{j} for short and we take ρ:=ρj>0\rho:=\rho_{j}>0 which is infinitesimal as j+j\to+\infty and such that, recalling (3.2),

(9.1) 0=limj+a~ερ2s={limj+1|lnε|ρ if s=12,limj+ε2s1ρ2s if s(12,1).0=\lim_{j\to+\infty}\frac{\tilde{a}_{\varepsilon}}{\rho^{2s}}=\begin{cases}\displaystyle\lim_{j\to+\infty}\frac{1}{|\ln\varepsilon|\rho}&\hbox{ if }s=\frac{1}{2},\\[7.5pt] \displaystyle\lim_{j\to+\infty}\frac{\varepsilon^{2s-1}}{\rho^{2s}}&\hbox{ if }s\in\left(\frac{1}{2},1\right).\end{cases}

The quantity ρ\rho will play a crucial role in the detection of the recovery sequence near the boundary of Ω\Omega. We will use the notation r:=ρ/εr:=\rho/\varepsilon. Notice that r+r\to+\infty as j+j\to+\infty, thanks to (9.1). We also denote by o(1)o(1) quantities that are infinitesimal as j+j\to+\infty.

Now, let EE\subset\mathbb{R} be such that

(9.2) Per(E,Ω)<+\operatorname{Per}(E,\Omega)<+\infty

and let d~\tilde{d} denote the signed distance to to E\partial E with the convention that d~0\tilde{d}\geqslant 0 in EE. Then, by [SV-gamma, Proposition 4.6], for any ΩΩ\Omega^{\prime}\Subset\Omega, it holds that

(9.3) lim supε0ε(1)(uε)cPer(E,Ω)whereuε(x):=u0(d~(x)ε)\limsup_{\varepsilon\searrow 0}{\mathcal{F}}^{(1)}_{\varepsilon}(u_{\varepsilon})\leqslant c_{\star}\operatorname{Per}(E,\Omega^{\prime})\qquad\hbox{where}\quad u_{\varepsilon}(x):=u_{0}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)

and c>0c_{\star}>0 depends only on ss and WW. Here above and in the rest of this section, u0u_{0} is the unique solution of (6.2).

We denote by dd the signed distance from Ω\partial\Omega, with the convention that d0d\geqslant 0 in Ω\Omega. Set π:Ω\pi:\mathbb{R}\to\partial\Omega to be the projection along the boundary of Ω\Omega. In particular, we have that

d(x)={|xπ(x)| if xΩ,|xπ(x)| if xΩc.d(x)=\begin{cases}|x-\pi(x)|&\hbox{ if }x\in\Omega,\\ -|x-\pi(x)|&\hbox{ if }x\in\Omega^{c}.\end{cases}

Also, we set

Ωρ:={xΩ s.t. d(x)(0,ρ)}.\Omega_{\rho}:=\{x\in\Omega{\mbox{ s.t. }}d(x)\in(0,\rho)\}.

We fix ρo>0\rho_{o}>0 sufficiently small so that |Ω|10ρo|\Omega|\geqslant 10\rho_{o}. Moreover, since (9.2) holds, up to taking ρo\rho_{o} smaller, we can suppose that

(9.4) sgn(d~(x))=±1\operatorname{sgn}(\tilde{d}(x))=\pm 1 does not change sign in Ω3ρo\Omega_{3\rho_{o}}.

For clarity and ease in notation, we also define the trace of sgn(d~)\operatorname{sgn}(\tilde{d}) in Ω¯\overline{\Omega} by setting

sgnE(x):={sgn(d~(x)) if xΩ,limΩyxsgn(d~(y)) if xΩ.\operatorname{sgn}_{E}(x):=\begin{cases}\operatorname{sgn}(\tilde{d}(x))&\hbox{ if }x\in\Omega,\\ \displaystyle\lim_{\Omega\ni y\to x}\operatorname{sgn}(\tilde{d}(y))&\hbox{ if }x\in\partial\Omega.\end{cases}

Finally, for sufficiently large jj (so that ρρo\rho\leqslant\rho_{o}), we define the function vε:[1,1]v_{\varepsilon}:\mathbb{R}\to[-1,1] by

(9.5) vε(x):={g(x) if xΩc,wε(x) if xΩρ,2ρd(x)ρ[wε(x)uε(x)]+uε(x) if xΩ2ρΩρ,uε(x) if xΩΩ2ρ.v_{\varepsilon}(x):=\begin{cases}g(x)&{\mbox{ if }}x\in\Omega^{c},\\ w_{\varepsilon}(x)&{\mbox{ if }}x\in\Omega_{\rho},\\ \frac{2\rho-d(x)}{\rho}\left[w_{\varepsilon}(x)-u_{\varepsilon}(x)\right]+u_{\varepsilon}(x)&{\mbox{ if }}x\in\Omega_{2\rho}\setminus\Omega_{\rho},\\ u_{\varepsilon}(x)&{\mbox{ if }}x\in\Omega\setminus\Omega_{2\rho}.\end{cases}

where uεu_{\varepsilon} is as given by (9.3) and

wε(x):=w0(d(x)ε;sgnE(x),g(π(x))).w_{\varepsilon}(x):=w_{0}\left(-\frac{d(x)}{\varepsilon};\operatorname{sgn}_{E}(x),g(\pi(x))\right).

Recall the notation in (7.2) for w0w_{0}.

We will show that vεv_{\varepsilon} in (9.5) is the so-called recovery sequence.

Let Ω:=ΩΩ2ρoΩ\Omega^{\prime}:=\Omega\setminus\Omega_{2\rho_{o}}\Subset\Omega be fixed. Moreover, for clarity, assume that

(9.6) Ω=(x¯1,x¯2)\Omega=(\bar{x}_{1},\bar{x}_{2})

and note that Bρ(x¯1)Bρ(x¯2)B_{\rho}(\bar{x}_{1})\cup B_{\rho}(\bar{x}_{2}) is a finite disjoint covering of Ω\partial\Omega for ρρo\rho\leqslant\rho_{o}. We also have for jj sufficiently large that

Ωρ\displaystyle\Omega_{\rho} =(Bρ(x¯1)Ω)(Bρ(x¯2)Ω)=Bρ+(x¯1)Bρ(x¯2)\displaystyle=(B_{\rho}(\bar{x}_{1})\cap\Omega)\cup(B_{\rho}(\bar{x}_{2})\cap\Omega)=B_{\rho}^{+}(\bar{x}_{1})\cup B_{\rho}^{-}(\bar{x}_{2})
and Ω2ρ\displaystyle{\mbox{and }}\quad\Omega_{2\rho} =(B2ρ(x¯1)Ω)(B2ρ(x¯2)Ω)=B2ρ+(x¯1)B2ρ(x¯2),\displaystyle=(B_{2\rho}(\bar{x}_{1})\cap\Omega)\cup(B_{2\rho}(\bar{x}_{2})\cap\Omega)=B_{2\rho}^{+}(\bar{x}_{1})\cup B_{2\rho}^{-}(\bar{x}_{2}),

where

Bδ+(x¯):={xBδ(x¯) s.t. x>x¯}andBδ(x¯):={xBδ(x¯) s.t. x<x¯}.B_{\delta}^{+}(\bar{x}):=\{x\in B_{\delta}(\bar{x})\hbox{ s.t. }x>\bar{x}\}\qquad\hbox{and}\qquad B_{\delta}^{-}(\bar{x}):=\{x\in B_{\delta}(\bar{x})\hbox{ s.t. }x<\bar{x}\}.

Recalling the definition of the set YκY_{\kappa} in (1.7), we assume that d~\tilde{d} (or equivalently the set EE) satisfies, for i=1,2i=1,2,

ifg(x¯i)>0,thend~(x)>0for allxBκ(x¯i)Ω;ifg(x¯i)<0,thend~(x)<0for allxBκ(x¯i)Ω;ifg(x¯i)=0,then eitherd~(x)>0ord~(x)<0for allxBκ(x¯i)Ω;\begin{array}[]{ll}\hbox{if}\leavevmode\nobreak\ g(\bar{x}_{i})>0,&\hbox{then}\leavevmode\nobreak\ \tilde{d}(x)>0\leavevmode\nobreak\ \hbox{for all}\leavevmode\nobreak\ x\in B_{\kappa}(\bar{x}_{i})\cap\Omega;\\ \hbox{if}\leavevmode\nobreak\ g(\bar{x}_{i})<0,&\hbox{then}\leavevmode\nobreak\ \tilde{d}(x)<0\leavevmode\nobreak\ \hbox{for all}\leavevmode\nobreak\ x\in B_{\kappa}(\bar{x}_{i})\cap\Omega;\\ \hbox{if}\leavevmode\nobreak\ g(\bar{x}_{i})=0,&\hbox{then either}\leavevmode\nobreak\ \tilde{d}(x)>0\leavevmode\nobreak\ \hbox{or}\leavevmode\nobreak\ \tilde{d}(x)<0\leavevmode\nobreak\ \hbox{for all}\leavevmode\nobreak\ x\in B_{\kappa}(\bar{x}_{i})\cap\Omega;\end{array}

Consequently, vεXgYκv_{\varepsilon}\in X_{g}\cap Y_{\kappa}.

We now estimate the kinetic and potential energies for ε(1)(vε)\mathcal{F}_{\varepsilon}^{(1)}(v_{\varepsilon}) separately.

9.1. Kinetic energy estimates

We first estimate the kinetic energy for vεv_{\varepsilon} as

(9.7) vε(Ω,Ω)+2vε(Ω,Ωc)\displaystyle v_{\varepsilon}(\Omega,\Omega)+2v_{\varepsilon}(\Omega,\Omega^{c})
(9.8) vε(Ω,Ω)+2vε(Ω,(Ω)c)\displaystyle\qquad\leqslant v_{\varepsilon}(\Omega^{\prime},\Omega^{\prime})+2v_{\varepsilon}(\Omega^{\prime},(\Omega^{\prime})^{c})
(9.9) +vε(Bρ+(x¯1),Bρ+(x¯1))+2vε(Bρ+(x¯1),Bρ(x¯1))\displaystyle\qquad\quad+v_{\varepsilon}(B_{\rho}^{+}(\bar{x}_{1}),B_{\rho}^{+}(\bar{x}_{1}))+2v_{\varepsilon}(B_{\rho}^{+}(\bar{x}_{1}),B_{\rho}^{-}(\bar{x}_{1}))
(9.10) +vε(Bρ(x¯2),Bρ(x¯2))+2vε(Bρ(x¯2),Ωρ+(x¯2))\displaystyle\qquad\quad+v_{\varepsilon}(B_{\rho}^{-}(\bar{x}_{2}),B_{\rho}^{-}(\bar{x}_{2}))+2v_{\varepsilon}(B_{\rho}^{-}(\bar{x}_{2}),\Omega_{\rho}^{+}(\bar{x}_{2}))
(9.11) +2vε(Bρ+(x¯1),(ΩcΩρ)Bρ(x¯1))+2vε(Bρ(x¯2),(ΩcΩρ)Bρ(x¯2))\displaystyle\qquad\quad+2v_{\varepsilon}(B_{\rho}^{+}(\bar{x}_{1}),(\Omega^{c}\cup\Omega_{\rho})\setminus B_{\rho}(\bar{x}_{1}))+2v_{\varepsilon}(B_{\rho}^{-}(\bar{x}_{2}),(\Omega^{c}\cup\Omega_{\rho})\setminus B_{\rho}(\bar{x}_{2}))
(9.12) +2vε(Ω2ρoΩρ,Ω2ρoΩc).\displaystyle\qquad\quad+2v_{\varepsilon}(\Omega_{2\rho_{o}}\setminus\Omega_{\rho},\Omega_{2\rho_{o}}\cup\Omega^{c}).

Under the scaling (3.2), we will use (9.3) (with the corresponding potential energy in Ω\Omega^{\prime}) to estimate the quantity in line (9.8). We then show that the boundary energy arises in-part from lines (9.9) and (9.10), and that the remaining terms (9.11) and (9.12) vanish as j+j\to+\infty.

First, we estimate from above the kinetic energy contributions of vεv_{\varepsilon} near the boundary of Ω\Omega, coming from inside Ω\Omega. To this end, for γ(1,1)\gamma\in(-1,1), we define

(9.13) Ψ1r(±1,γ):=brBr×Br|w0(x;±1,γ)w0(y;±1,γ)|2|xy|1+2s𝑑x𝑑y\Psi_{1}^{r}(\pm 1,\gamma):=b_{r}\iint_{B_{r}^{-}\times B_{r}^{-}}\frac{|w_{0}(x;\pm 1,\gamma)-w_{0}(y;\pm 1,\gamma)|^{2}}{|x-y|^{1+2s}}\,dx\,dy

where, recalling (1.3), we set

(9.14) br={1|lnr| if s=12,1 if s(12,1).b_{r}=\begin{cases}\frac{1}{|\ln r|}&\hbox{ if }s=\frac{1}{2},\\ 1&\hbox{ if }s\in\left(\frac{1}{2},1\right).\end{cases}

This quantity is well-defined for a fixed r>0r>0 and, since w0w_{0} is a global minimizer of 𝒢s\mathcal{G}_{s}, recalling the notation in (7.3),

lim supr+Ψ1r(±1,γ)Ψ(±1,γ)<+.\limsup_{r\to+\infty}\Psi_{1}^{r}(\pm 1,\gamma)\leqslant\Psi(\pm 1,\gamma)<+\infty.
Lemma 9.1.

If x¯Ω\bar{x}\in\partial\Omega, then

a~εvε(Bρ(x¯)Ω,Bρ(x¯)Ω)Ψ1r(sgnE(x¯),g(x¯)).\tilde{a}_{\varepsilon}v_{\varepsilon}(B_{\rho}(\bar{x})\cap\Omega,\,B_{\rho}(\bar{x})\cap\Omega)\leqslant\Psi_{1}^{r}\big{(}\operatorname{sgn}_{E}(\bar{x}),\,g(\bar{x})\big{)}.
Proof.

Without loss of generality, assume that

(9.15) x¯=0andBρo(x¯)Ω=Bρo=(ρo,0).\bar{x}=0\qquad\hbox{and}\qquad B_{\rho_{o}}(\bar{x})\cap\Omega=B_{\rho_{o}}^{-}=(-\rho_{o},0).

Consequently,

(9.16) π(x)=0andd(x)=xfor xBρo.\pi(x)=0\qquad\hbox{and}\qquad d(x)=-x\quad\hbox{for }x\in B_{\rho_{o}}^{-}.

For ease, set

(9.17) w0(x):=w0(x;sgnE(0),g(0)).w_{0}(x):=w_{0}(x;\operatorname{sgn}_{E}(0),g(0)).

With this and recalling (9.4) and (9.5), we use the change of variables x~=x/ε\tilde{x}=x/\varepsilon and y~=y/ε\tilde{y}=y/\varepsilon to find that

a~εvε(BρΩ,BρΩ)\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}(B_{\rho}\cap\Omega,\,B_{\rho}\cap\Omega) =a~ερ0ρ0|w0(xε)w0(yε)|2|xy|1+2s𝑑y𝑑x\displaystyle=\tilde{a}_{\varepsilon}\int_{-\rho}^{0}\int_{-\rho}^{0}\frac{\left|w_{0}\left(\frac{x}{\varepsilon}\right)-w_{0}\left(\frac{y}{\varepsilon}\right)\right|^{2}}{|x-y|^{1+2s}}\,dy\,dx
=ε12sa~ερε0ρε0|w0(x~)w0(y~)|2|x~y~|1+2s𝑑y~𝑑x~.\displaystyle=\varepsilon^{1-2s}\tilde{a}_{\varepsilon}\int_{-\frac{\rho}{\varepsilon}}^{0}\int_{-\frac{\rho}{\varepsilon}}^{0}\frac{\left|w_{0}(\tilde{x})-w_{0}(\tilde{y})\right|^{2}}{|\tilde{x}-\tilde{y}|^{1+2s}}\,d\tilde{y}\,d\tilde{x}.

Recall that r=ρ/εr=\rho/\varepsilon. Thus, when s(12,1)s\in\left(\frac{1}{2},1\right), we have that ε12sa~ε=1\varepsilon^{1-2s}\tilde{a}_{\varepsilon}=1 and the lemma holds with equality.

If instead s=12s=\frac{1}{2}, recall that a~ε=1/|lnε|\tilde{a}_{\varepsilon}=1/|\ln\varepsilon|. Notice that, for jj large,

(9.18) 1|lnε|=|lnr||lnε|1|lnr|=|lnε||lnρ||lnε|1|lnr|1|lnr|.\frac{1}{|\ln\varepsilon|}=\frac{|\ln r|}{|\ln\varepsilon|}\frac{1}{|\ln r|}=\frac{|\ln\varepsilon|-|\ln\rho|}{|\ln\varepsilon|}\frac{1}{|\ln r|}\leqslant\frac{1}{|\ln r|}.

Therefore,

1|lnε|vε(BρΩ,BρΩ)\displaystyle\frac{1}{|\ln\varepsilon|}v_{\varepsilon}(B_{\rho}\cap\Omega,\,B_{\rho}\cap\Omega) 1|lnr|r0r0|w0(x~)w0(y~)|2|x~y~|2𝑑y~𝑑x~,\displaystyle\leqslant\frac{1}{|\ln r|}\int_{-r}^{0}\int_{-r}^{0}\frac{\left|w_{0}(\tilde{x})-w_{0}(\tilde{y})\right|^{2}}{|\tilde{x}-\tilde{y}|^{2}}\,d\tilde{y}\,d\tilde{x},

as desired. ∎

Now we address the counterpart of Lemma 9.1, in which we estimate from above the mixed energy contributions of vεv_{\varepsilon} near the boundary of Ω\Omega, coming from the interactions between Ω\Omega and its complement. To this aim, we set

(9.19) Ψ2r(±1,γ):=brBr×Br+|w0(x;±1,γ)γ|2|xy|1+2s𝑑x𝑑y.\Psi_{2}^{r}(\pm 1,\gamma):=b_{r}\iint_{B_{r}^{-}\times B_{r}^{+}}\frac{\left|w_{0}(x;\pm 1,\gamma)-\gamma\right|^{2}}{|x-y|^{1+2s}}\,dx\,dy.

The reader may compare (9.13) and (9.19) to appreciate the different interactions coming from inside the domain and the ones coming from the inside/outside relations. Note that

lim supr+Ψ2r(±1,γ)Ψ(±1,γ)<+.\limsup_{r\to+\infty}\Psi_{2}^{r}(\pm 1,\gamma)\leqslant\Psi(\pm 1,\gamma)<+\infty.
Lemma 9.2.

If x¯Ω\bar{x}\in\partial\Omega, then

a~εvε(Bρ(x¯)Ω,Bρ(x¯)Ω)Ψ2r(sgnE(x¯),g(x¯))+o(1).\tilde{a}_{\varepsilon}v_{\varepsilon}(B_{\rho}(\bar{x})\cap\Omega,\,B_{\rho}(\bar{x})\setminus\Omega)\leqslant\Psi_{2}^{r}\big{(}\operatorname{sgn}_{E}(\bar{x}),g(\bar{x})\big{)}+o(1).
Proof.

As in the proof of Lemma 9.1, assume, without loss of generality, that (9.15), (9.16), and (9.17) hold. We recall (9.5) to find that

vε(x)vε(y)=w0(xε)g(y)for xBρ and yBρ+.v_{\varepsilon}(x)-v_{\varepsilon}(y)=w_{0}\left(\frac{x}{\varepsilon}\right)-g(y)\quad\hbox{for\leavevmode\nobreak\ $x\in B_{\rho}^{-}$ and $y\in B_{\rho}^{+}$.}

Consequently,

|vε(x)vε(y)|\displaystyle|v_{\varepsilon}(x)-v_{\varepsilon}(y)| |w0(xε)g(0)|+|g(0)g(y)|\displaystyle\leqslant\left|w_{0}\left(\frac{x}{\varepsilon}\right)-g(0)\right|+|g(0)-g(y)|
|w0(xε)g(0)|+Cmin{1,|xy|}.\displaystyle\leqslant\left|w_{0}\left(\frac{x}{\varepsilon}\right)-g(0)\right|+C\min\{1,|x-y|\}.

Now, we use the following elementary inequality: if α\alpha, β0\beta\geqslant 0, then

2αβ=2α|lnρ|(|lnρ|β)α2|lnρ|+|lnρ|β2,2\alpha\beta=2\,\frac{\alpha}{\sqrt{|\ln\rho|}}\cdot({\sqrt{|\ln\rho|}}\,\beta)\leqslant\frac{\alpha^{2}}{{|\ln\rho|}}+|\ln\rho|\,\beta^{2},

so that

(9.20) (α+β)2=α2+2αβ+β2(1+1|lnρ|)α2+(1+|lnρ|)β2.(\alpha+\beta)^{2}=\alpha^{2}+2\alpha\beta+\beta^{2}\leqslant\left(1+\frac{1}{|\ln\rho|}\right)\alpha^{2}+\big{(}1+|\ln\rho|\big{)}\beta^{2}.

Hence, we find that

|vε(x)vε(y)|2(1+1|lnρ|)|w0(xε)g(0)|2+C|lnρ|min{1,|xy|2},\displaystyle|v_{\varepsilon}(x)-v_{\varepsilon}(y)|^{2}\leqslant\left(1+\frac{1}{|\ln\rho|}\right)\,\left|w_{0}\left(\frac{x}{\varepsilon}\right)-g(0)\right|^{2}+C\,|\ln\rho|\,\min\{1,\,|x-y|^{2}\},

for jj sufficiently large and for some C>0C>0.

This gives that

a~εvε(Bρ,Bρ+)\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}(B_{\rho}^{-},\,B_{\rho}^{+}) (1+1|lnρ|)a~ερ00ρ|w0(xε)g(0)|2|xy|1+2s𝑑y𝑑x\displaystyle\leqslant\left(1+\frac{1}{|\ln\rho|}\right)\,\tilde{a}_{\varepsilon}\int_{-\rho}^{0}\int_{0}^{\rho}\frac{\left|w_{0}\left(\frac{x}{\varepsilon}\right)-g(0)\right|^{2}}{|x-y|^{1+2s}}\,dy\,dx
+Ca~ε|lnρ|ρ00ρmin{1,|xy|2}|xy|1+2s𝑑y𝑑x\displaystyle\qquad+C\,\tilde{a}_{\varepsilon}\,|\ln\rho|\int_{-\rho}^{0}\int_{0}^{\rho}\frac{\min\{1,\,|x-y|^{2}\}}{|x-y|^{1+2s}}\,dy\,dx
(1+1|lnρ|)a~ερ00ρ|w0(xε)g(0)|2|xy|1+2s𝑑y𝑑x+Ca~ε|lnρ|ρ\displaystyle\leqslant\left(1+\frac{1}{|\ln\rho|}\right)\,\tilde{a}_{\varepsilon}\int_{-\rho}^{0}\int_{0}^{\rho}\frac{\left|w_{0}\left(\frac{x}{\varepsilon}\right)-g(0)\right|^{2}}{|x-y|^{1+2s}}\,dy\,dx+C\,\tilde{a}_{\varepsilon}\,|\ln\rho|\cdot\rho
=(1+1|lnρ|)ε12sa~ερε00ρε|w0(x~)g(0)|2|x~y~|1+2s𝑑y~𝑑x~+Ca~ε|lnρ|ρ\displaystyle=\left(1+\frac{1}{|\ln\rho|}\right)\,\varepsilon^{1-2s}\tilde{a}_{\varepsilon}\int_{-\frac{\rho}{\varepsilon}}^{0}\int_{0}^{\frac{\rho}{\varepsilon}}\frac{\left|w_{0}\left(\tilde{x}\right)-g(0)\right|^{2}}{|\tilde{x}-\tilde{y}|^{1+2s}}\,d\tilde{y}\,d\tilde{x}+C\,\tilde{a}_{\varepsilon}\,|\ln\rho|\cdot\rho

where we use the substitutions x~=x/ε\tilde{x}=x/\varepsilon and y~=y/ε\tilde{y}=y/\varepsilon.

For s(12,1)s\in\left(\frac{1}{2},1\right), we have that

ε2s1vε(BρΩ,BρΩ)\displaystyle\varepsilon^{2s-1}v_{\varepsilon}(B_{\rho}\cap\Omega,\,B_{\rho}\setminus\Omega) (1+1|lnρ|)r00r|w0(x~)g(0)|2|x~y~|1+2s𝑑y~𝑑x~+C|lnρ|ρ2\displaystyle\leqslant\left(1+\frac{1}{|\ln\rho|}\right)\int_{-r}^{0}\int_{0}^{r}\frac{\left|w_{0}\left(\tilde{x}\right)-g(0)\right|^{2}}{|\tilde{x}-\tilde{y}|^{1+2s}}\,d\tilde{y}\,d\tilde{x}+C\,|\ln\rho|\cdot\rho^{2}
r00r|w0(x~)g(0)|2|x~y~|1+2s𝑑y~𝑑x~+o(1),\displaystyle\leqslant\int_{-r}^{0}\int_{0}^{r}\frac{\left|w_{0}\left(\tilde{x}\right)-g(0)\right|^{2}}{|\tilde{x}-\tilde{y}|^{1+2s}}\,d\tilde{y}\,d\tilde{x}+o(1),

as desired.

For s=12s=\frac{1}{2}, recalling (9.18), we see that

1|lnε|\displaystyle\frac{1}{|\ln\varepsilon|} vε(BρΩ,BρΩ)\displaystyle v_{\varepsilon}(B_{\rho}\cap\Omega,\,B_{\rho}\setminus\Omega)
(1+1|lnρ|)1|lnε|r00r|w0(x~)g(0)|2|x~y~|2𝑑y~𝑑x~+C|lnε||lnρ|ρ\displaystyle\leqslant\left(1+\frac{1}{|\ln\rho|}\right)\frac{1}{|\ln\varepsilon|}\int_{-r}^{0}\int_{0}^{r}\frac{\left|w_{0}\left(\tilde{x}\right)-g(0)\right|^{2}}{|\tilde{x}-\tilde{y}|^{2}}\,d\tilde{y}\,d\tilde{x}+\frac{C}{|\ln\varepsilon|}\,|\ln\rho|\cdot\rho
=(1+1|lnρ|)1|lnr|r00r|w0(x~)g(0)|2|x~y~|2𝑑y~𝑑x~+C|lnε||lnρ|ρ\displaystyle=\left(1+\frac{1}{|\ln\rho|}\right)\frac{1}{|\ln r|}\int_{-r}^{0}\int_{0}^{r}\frac{\left|w_{0}\left(\tilde{x}\right)-g(0)\right|^{2}}{|\tilde{x}-\tilde{y}|^{2}}\,d\tilde{y}\,d\tilde{x}+\frac{C}{|\ln\varepsilon|}\,|\ln\rho|\cdot\rho
1|lnr|r00r|w0(x~)g(0)|2|x~y~|2𝑑y~𝑑x~+o(1).\displaystyle\leqslant\frac{1}{|\ln r|}\int_{-r}^{0}\int_{0}^{r}\frac{\left|w_{0}\left(\tilde{x}\right)-g(0)\right|^{2}}{|\tilde{x}-\tilde{y}|^{2}}\,d\tilde{y}\,d\tilde{x}+o(1).

This completes the proof. ∎

It remains to estimate the error terms in lines (9.11) and (9.12). For this, we first prove the following regularity estimate away from Ω\partial\Omega.

Lemma 9.3.

Let x¯Ω\bar{x}\in\partial\Omega. If xx, y(B2ρo(x¯)Ω)(Bρ10(x¯)Ω)y\in(B_{2\rho_{o}}(\bar{x})\cap\Omega)\setminus(B_{\frac{\rho}{10}}(\bar{x})\cap\Omega), then, for any α(0,1)\alpha\in(0,1), there exists a constant C>0C>0 such that

|vε(x)vε(y)|C(|xy|αρα+|xy|ρ).|v_{\varepsilon}(x)-v_{\varepsilon}(y)|\leqslant C\left(\frac{|x-y|^{\alpha}}{\rho^{\alpha}}+\frac{|x-y|}{\rho}\right).

Moreover, if xx, y(B2ρo(x¯)Ω)(B2ρ(x¯)Ω)y\in(B_{2\rho_{o}}(\bar{x})\cap\Omega)\setminus(B_{2\rho}(\bar{x})\cap\Omega), there exists C>0C>0 such that

|vε(x)vε(y)|C|xy|ρ.|v_{\varepsilon}(x)-v_{\varepsilon}(y)|\leqslant C\frac{|x-y|}{{\rho}}.
Remark 9.4.

If s(12,1)s\in\left(\frac{1}{2},1\right) then, thanks to (1.13), one can check that Lemma 9.3 holds for α=1\alpha=1.

Proof of Lemma 9.3.

Recall the notation in (9.6) and notice that, without loss of generality, we can assume that x¯=x¯2=0\bar{x}=\bar{x}_{2}=0.

In this setting, we write that d(x)=xd(x)=-x for all x(2ρo,0)x\in(-2\rho_{o},0). In light of (9.4), we also have that sgnE(x)=sgnE(0)\operatorname{sgn}_{E}(x)=\operatorname{sgn}_{E}(0) for all x(2ρo,0)x\in(-2\rho_{o},0).

Assume also that sgnE(x)=1\operatorname{sgn}_{E}(x)=-1 and set w0(x):=w0(x;1,g(0))w_{0}(x):=w_{0}(x;-1,g(0)).

We now break the proof into cases based on the definition of vε(x)v_{\varepsilon}(x) and vε(y)v_{\varepsilon}(y), according to (9.5).

Case 1. Suppose that ρ<x<y<ρ10-\rho<x<y<-\frac{\rho}{10}. Note that xx, yΩρy\in\Omega_{\rho}, therefore,

|vε(x)vε(y)|=|wε(x)wε(y)|=|w0(xε)w0(yε)|.|v_{\varepsilon}(x)-v_{\varepsilon}(y)|=|w_{\varepsilon}(x)-w_{\varepsilon}(y)|=\left|w_{0}\left(\frac{x}{\varepsilon}\right)-w_{0}\left(\frac{y}{\varepsilon}\right)\right|.

Thus, using also (1.12), we have that

|vε(x)vε(y)|[w0]Cα((ρε,ρ10ε))|xyε|αC|xyρ|α.|v_{\varepsilon}(x)-v_{\varepsilon}(y)|\leqslant[w_{0}]_{C^{\alpha}((-\frac{\rho}{\varepsilon},-\frac{\rho}{10\varepsilon}))}\left|\frac{x-y}{\varepsilon}\right|^{\alpha}\leqslant C\left|\frac{x-y}{\rho}\right|^{\alpha}.

Case 2. Suppose that 2ρ<x<y<ρ-2\rho<x<y<-\rho. Since xx, yΩ2ρΩρy\in\Omega_{2\rho}\setminus\Omega_{\rho}, we write

|vε(x)vε(y)|\displaystyle|v_{\varepsilon}(x)-v_{\varepsilon}(y)|
=|(2ρ+xρ(wε(x)uε(x))+uε(x))(2ρ+yρ(wε(y)uε(y))+uε(y))|\displaystyle=\left|\left(\frac{2\rho+x}{\rho}\left(w_{\varepsilon}(x)-u_{\varepsilon}(x)\right)+u_{\varepsilon}(x)\right)-\left(\frac{2\rho+y}{\rho}\left(w_{\varepsilon}(y)-u_{\varepsilon}(y)\right)+u_{\varepsilon}(y)\right)\right|
|2ρ+xρ||wε(x)wε(y)|+|ρ+xρ||uε(x)uε(y)|+|yxρ||wε(y)uε(y)|\displaystyle\leqslant\left|\frac{2\rho+x}{\rho}\right|\left|w_{\varepsilon}(x)-w_{\varepsilon}(y)\right|+\left|\frac{\rho+x}{\rho}\right|\left|u_{\varepsilon}(x)-u_{\varepsilon}(y)\right|+\left|\frac{y-x}{\rho}\right|\left|w_{\varepsilon}(y)-u_{\varepsilon}(y)\right|
|wε(x)wε(y)|+|uε(x)uε(y)|+2|yxρ|.\displaystyle\leqslant\left|w_{\varepsilon}(x)-w_{\varepsilon}(y)\right|+\left|u_{\varepsilon}(x)-u_{\varepsilon}(y)\right|+2\left|\frac{y-x}{\rho}\right|.

We use (1.12) to estimate

|wε(x)wε(y)||w0(xε)w0(yε)|[w0]Cα((2ρε,ρε))|xyε|αC|xyρ|α.\left|w_{\varepsilon}(x)-w_{\varepsilon}(y)\right|\leqslant\left|w_{0}\left(\frac{x}{\varepsilon}\right)-w_{0}\left(\frac{y}{\varepsilon}\right)\right|\leqslant[w_{0}]_{C^{\alpha}((-\frac{2\rho}{\varepsilon},-\frac{\rho}{\varepsilon}))}\left|\frac{x-y}{\varepsilon}\right|^{\alpha}\leqslant C\left|\frac{x-y}{\rho}\right|^{\alpha}.

Also, as a consequence of (9.4), we have that |d~(z)|>ρ|\tilde{d}(z)|>\rho for z(2ρ,ρ)z\in(-2\rho,-\rho) for jj sufficiently large. Hence, by (6.3), we have that

|uε(z)|=|u0(d~(z)ε)||d~(z)ε|Cε1+2s|d~(z)|1+2s1εCε2sρ1+2s.|u_{\varepsilon}^{\prime}(z)|=\left|u_{0}^{\prime}\left(\frac{\tilde{d}(z)}{\varepsilon}\right)\right|\left|\frac{\tilde{d}^{\prime}(z)}{\varepsilon}\right|\leqslant\frac{C\varepsilon^{1+2s}}{|\tilde{d}(z)|^{1+2s}}\frac{1}{\varepsilon}\leqslant\frac{C\varepsilon^{2s}}{\rho^{1+2s}}.

With this, we obtain that

(9.21) |uε(x)uε(y)|uεL((2ρ,ρ))|xy|Cε2sρ1+2s|xy|C|xy|ρ.\left|u_{\varepsilon}(x)-u_{\varepsilon}(y)\right|\leqslant\|u_{\varepsilon}^{\prime}\|_{L^{\infty}((-2\rho,-\rho))}|x-y|\leqslant\frac{C\varepsilon^{2s}}{\rho^{1+2s}}|x-y|\leqslant C\frac{|x-y|}{\rho}.

Therefore,

|vε(x)vε(y)|C|xyρ|α+C|xy|ρ+2|xy|ρ.|v_{\varepsilon}(x)-v_{\varepsilon}(y)|\leqslant C\left|\frac{x-y}{\rho}\right|^{\alpha}+C\frac{|x-y|}{\rho}+2\frac{|x-y|}{\rho}.

Case 3. Suppose that 2ρo<x<y<2ρ-2\rho_{o}<x<y<-2\rho. Since xx, yΩ2ρoΩ2ρy\in\Omega_{2\rho_{o}}\setminus\Omega_{2\rho}, we recall (9.4) and estimate as in (9.21) to find that

|vε(x)vε(y)|=|uε(x)uε(y)|uεL((2ρo,2ρ))|xy|Cε2sρ1+2s|xy|C|xy|ρ.|v_{\varepsilon}(x)-v_{\varepsilon}(y)|=\left|u_{\varepsilon}(x)-u_{\varepsilon}(y)\right|\leqslant\|u_{\varepsilon}^{\prime}\|_{L^{\infty}((-2\rho_{o},-2\rho))}|x-y|\leqslant\frac{C\varepsilon^{2s}}{\rho^{1+2s}}|x-y|\leqslant C\frac{|x-y|}{\rho}.

Case 4. Suppose that 2ρ<x<ρ<y<ρ10-2\rho<x<-\rho<y<-\frac{\rho}{10}. In this case, xΩ2ρΩρx\in\Omega_{2\rho}\setminus\Omega_{\rho} and yΩρy\in\Omega_{\rho}, so we write

|vε(x)vε(y)|\displaystyle|v_{\varepsilon}(x)-v_{\varepsilon}(y)| =|(2ρ+xρ(wε(x)uε(x))+uε(x))wε(y)|\displaystyle=\left|\left(\frac{2\rho+x}{\rho}\left(w_{\varepsilon}(x)-u_{\varepsilon}(x)\right)+u_{\varepsilon}(x)\right)-w_{\varepsilon}(y)\right|
|2ρ+xρ||w0(xε)w0(yε)|+|ρ+xρ||w0(yε)uε(x)|.\displaystyle\leqslant\left|\frac{2\rho+x}{\rho}\right|\left|w_{0}\left(\frac{x}{\varepsilon}\right)-w_{0}\left(\frac{y}{\varepsilon}\right)\right|+\left|\frac{\rho+x}{\rho}\right|\left|w_{0}\left(\frac{y}{\varepsilon}\right)-u_{\varepsilon}(x)\right|.

Recalling (1.12) and using that |ρ+x|=ρx<yx=|xy||\rho+x|=-\rho-x<y-x=|x-y|, we find that

|vε(x)vε(y)|\displaystyle|v_{\varepsilon}(x)-v_{\varepsilon}(y)| [w0]Cα((2ρε,ρ10ε))|xyε|α+2|ρ+x|ρC|xyρ|α+2|xy|ρ.\displaystyle\leqslant[w_{0}]_{C^{\alpha}((-\frac{2\rho}{\varepsilon},-\frac{\rho}{10\varepsilon}))}\left|\frac{x-y}{\varepsilon}\right|^{\alpha}+2\frac{|\rho+x|}{\rho}\leqslant C\left|\frac{x-y}{\rho}\right|^{\alpha}+2\frac{|x-y|}{\rho}.

Case 5. Suppose that 2ρo<x<2ρ<y<ρ-2\rho_{o}<x<-2\rho<y<-\rho. This is the case in which xΩ2ρoΩ2ρx\in\Omega_{2\rho_{o}}\setminus\Omega_{2\rho} and yΩ2ρΩρy\in\Omega_{2\rho}\setminus\Omega_{\rho}. Estimating as in (9.21) and using that |2ρ+y|<|xy||2\rho+y|<|x-y|, we obtain that

|vε(x)vε(y)|\displaystyle|v_{\varepsilon}(x)-v_{\varepsilon}(y)| =|uε(x)(2ρ+yρ(wε(y)uε(y))+uε(y))|\displaystyle=\left|u_{\varepsilon}(x)-\left(\frac{2\rho+y}{\rho}\left(w_{\varepsilon}(y)-u_{\varepsilon}(y)\right)+u_{\varepsilon}(y)\right)\right|
|uε(x)uε(y)|+|2ρ+yρ||wε(y)uε(y)|\displaystyle\leqslant|u_{\varepsilon}(x)-u_{\varepsilon}(y)|+\left|\frac{2\rho+y}{\rho}\right|\left|w_{\varepsilon}(y)-u_{\varepsilon}(y)\right|
uεL((2ρo,ρ))|xy|+2|xy|ρ\displaystyle\leqslant\|u_{\varepsilon}^{\prime}\|_{L^{\infty}((-2\rho_{o},-\rho))}|x-y|+2\frac{|x-y|}{\rho}
C|xy|ρ+2|xy|ρ\displaystyle\leqslant C\frac{|x-y|}{\rho}+2\frac{|x-y|}{\rho}
C|xy|ρ.\displaystyle\leqslant C\frac{|x-y|}{\rho}.

Case 6. Suppose that 2ρo<x<2ρ-2\rho_{o}<x<-2\rho and ρ<y<ρ10-\rho<y<-\frac{\rho}{10}. This case is trivial since |vε|1|v_{\varepsilon}|\leqslant 1:

|vε(x)vε(y)|22|xy|ρ.|v_{\varepsilon}(x)-v_{\varepsilon}(y)|\leqslant 2\leqslant 2\frac{|x-y|}{\rho}.

We have exhausted all cases and the proof of Lemma 9.3 is complete. ∎

Lemma 9.5.

It holds that

(9.22) a~ε[vε(Bρ(x¯1),(ΩcΩρ)Bρ(x¯1))+vε(Bρ(x¯2),(ΩcΩρ)Bρ(x¯2))]=o(1)\tilde{a}_{\varepsilon}\Big{[}v_{\varepsilon}(B_{\rho}^{-}(\bar{x}_{1}),(\Omega^{c}\cup\Omega_{\rho})\setminus B_{\rho}(\bar{x}_{1}))+v_{\varepsilon}(B_{\rho}^{-}(\bar{x}_{2}),(\Omega^{c}\cup\Omega_{\rho})\setminus B_{\rho}(\bar{x}_{2}))\Big{]}=o(1)

and

(9.23) a~εvε(Ω2ρoΩρ,Ω2ρoΩc)=o(1).\tilde{a}_{\varepsilon}v_{\varepsilon}(\Omega_{2\rho_{o}}\setminus\Omega_{\rho},\Omega_{2\rho_{o}}\cup\Omega^{c})=o(1).
Proof.

For any δ>0\delta>0, we use that vε[1,1]v_{\varepsilon}\in[-1,1] to note that

(9.24) |xy|>δ|vε(x)vε(y)|2|xy|1+2s𝑑yCδ+r12s𝑑r=Cδ2s.\int_{|x-y|>\delta}\frac{|v_{\varepsilon}(x)-v_{\varepsilon}(y)|^{2}}{|x-y|^{1+2s}}\,dy\leqslant C\int_{\delta}^{+\infty}r^{-1-2s}\,dr=C\delta^{-2s}.

With this, we first estimate

a~εvε(Bρ(x¯2),(ΩcΩρ)Bρ(x¯2))a~εBρ(x¯2)|xy|>ρ|vε(x)vε(y)|2|xy|1+2s𝑑y𝑑x\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}(B_{\rho}^{-}(\bar{x}_{2}),(\Omega^{c}\cup\Omega_{\rho})\setminus B_{\rho}(\bar{x}_{2}))\leqslant\tilde{a}_{\varepsilon}\int_{B_{\rho}^{-}(\bar{x}_{2})}\int_{|x-y|>\rho}\frac{|v_{\varepsilon}(x)-v_{\varepsilon}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx
Ca~ερ2s|Bρ(x¯2)|=Ca~ερ2s1=o(1)\displaystyle\qquad\qquad\leqslant C\frac{\tilde{a}_{\varepsilon}}{\rho^{2s}}|B_{\rho}^{-}(\bar{x}_{2})|=C\frac{\tilde{a}_{\varepsilon}}{\rho^{2s-1}}=o(1)

and similarly find that vε(Bρ+(x¯1),(ΩcΩρ)Bρ(x¯1))=o(1)v_{\varepsilon}(B_{\rho}^{+}(\bar{x}_{1}),(\Omega^{c}\cup\Omega_{\rho})\setminus B_{\rho}(\bar{x}_{1}))=o(1). This proves (9.22).

Regarding (9.23), we first estimate the long-range interactions in the same way to find that

a~εvε(B2ρo(x¯2)Bρ(x¯2),B2ρo+(x¯1)Ωc)a~εx22ρox2ρ|xy|>ρ|vε(x)vε(y)|2|xy|1+2s𝑑y𝑑x\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}\big{(}B_{2\rho_{o}}^{-}(\bar{x}_{2})\setminus B_{\rho}^{-}(\bar{x}_{2}),B_{2\rho_{o}}^{+}(\bar{x}_{1})\cup\Omega^{c}\big{)}\leqslant\tilde{a}_{\varepsilon}\int^{x_{2}-\rho}_{x_{2}-2\rho_{o}}\int_{|x-y|>\rho}\frac{|v_{\varepsilon}(x)-v_{\varepsilon}(y)|^{2}}{|x-y|^{1+2s}}\,dydx
Ca~ερ2s(2ρ0ρ)=o(1).\displaystyle\qquad\qquad\leqslant C\frac{\tilde{a}_{\varepsilon}}{\rho^{2s}}(2\rho_{0}-\rho)=o(1).

and similarly find that a~εvε(B2ρo+(x¯1)Bρ+(x¯1),B2ρo(x¯2)Ωc)=o(1)\tilde{a}_{\varepsilon}v_{\varepsilon}(B_{2\rho_{o}}^{+}(\bar{x}_{1})\setminus B_{\rho}^{+}(\bar{x}_{1}),B_{2\rho_{o}}^{-}(\bar{x}_{2})\cup\Omega^{c})=o(1).

It is left to show that

a~ε[vε(B2ρo(x¯2),B2ρo(x¯2)Bρ(x¯2))+vε(B2ρo+(x¯1),B2ρo+(x¯1)Bρ+(x¯1))]=o(1).\tilde{a}_{\varepsilon}\Big{[}v_{\varepsilon}(B_{2\rho_{o}}^{-}(\bar{x}_{2}),B_{2\rho_{o}}^{-}(\bar{x}_{2})\setminus B_{\rho}^{-}(\bar{x}_{2}))+v_{\varepsilon}(B_{2\rho_{o}}^{+}(\bar{x}_{1}),B_{2\rho_{o}}^{+}(\bar{x}_{1})\setminus B_{\rho}^{+}(\bar{x}_{1}))\Big{]}=o(1).

We only prove the estimate around x¯2\bar{x}_{2} since the estimate near x¯1\bar{x}_{1} is similar. For this, we assume, without loss of generality, that x¯2=0\bar{x}_{2}=0, so that

B2ρo(x¯2)=(2ρo,0)andB2ρo(x¯2)Bρ(x¯2)=(2ρo,ρ).B_{2\rho_{o}}^{-}(\bar{x}_{2})=(-2\rho_{o},0)\qquad\hbox{and}\qquad B_{2\rho_{o}}^{-}(\bar{x}_{2})\setminus B_{\rho}^{-}(\bar{x}_{2})=(-2\rho_{o},-\rho).

We thus want to prove that

(9.25) a~εvε((2ρo,0),(2ρo,ρ))=o(1).\tilde{a}_{\varepsilon}v_{\varepsilon}((-2\rho_{o},0),(-2\rho_{o},-\rho))=o(1).

For this, we split the remaining error around x¯2=0\bar{x}_{2}=0 as

(9.26) vε((2ρo,0),(2ρo,ρ))=vε((ρ/10,0),(2ρo,ρ))+vε((2ρo,ρ/10),(2ρ,ρ))+vε((2ρ,ρ/10),(2ρo,2ρ))+vε((2ρo,2ρ),(2ρo,2ρ)).\begin{split}v_{\varepsilon}((-2\rho_{o},0),(-2\rho_{o},-\rho))&=v_{\varepsilon}((-\rho/10,0),(-2\rho_{o},-\rho))\\ &\quad+v_{\varepsilon}((-2\rho_{o},-\rho/10),(-2\rho,-\rho))\\ &\quad+v_{\varepsilon}((-2\rho,-\rho/10),(-2\rho_{o},-2\rho))\\ &\quad+v_{\varepsilon}((-2\rho_{o},-2\rho),(-2\rho_{o},-2\rho)).\end{split}

Using again (9.24), we see that

a~εvε((ρ/10,0),(2ρo,ρ))\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}((-\rho/10,0),(-2\rho_{o},-\rho)) a~ε2ρoρ|xy|>9ρ10|vε(x)vε(y)|2|xy|1+2s𝑑y𝑑x=o(1).\displaystyle\leqslant\tilde{a}_{\varepsilon}\int^{-\rho}_{-2\rho_{o}}\int_{|x-y|>\frac{9\rho}{10}}\frac{|v_{\varepsilon}(x)-v_{\varepsilon}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx=o(1).

In the following, we let

α=α(s):={910 if s=12,1 if s(12,1).\alpha=\alpha(s):=\begin{cases}\frac{9}{10}&\hbox{ if }s=\frac{1}{2},\\ 1&\hbox{ if }s\in\left(\frac{1}{2},1\right).\end{cases}

We use Lemma 9.3 (with Remark 9.4) to estimate

a~εvε((2ρo,ρ/10),(2ρ,ρ))\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}((-2\rho_{o},-\rho/10),(-2\rho,-\rho))
\displaystyle\leqslant a~ε[2ρρ{|xy|<ρ}{y<ρ/10}|vε(x)vε(y)|2|xy|1+2sdydx\displaystyle\tilde{a}_{\varepsilon}\left[\int^{-\rho}_{-2\rho}\int_{\{|x-y|<\rho\}\cap\{y<-\rho/10\}}\frac{|v_{\varepsilon}(x)-v_{\varepsilon}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\right.
+2ρρ{|xy|>ρ}|vε(x)vε(y)|2|xy|1+2sdydx]\displaystyle\qquad\qquad\left.+\int^{-\rho}_{-2\rho}\int_{\{|x-y|>\rho\}}\frac{|v_{\varepsilon}(x)-v_{\varepsilon}(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\right]
\displaystyle\leqslant Ca~ε[2ρρ{|xy|<ρ}ρ2α|xy|2α|xy|1+2sdydx+2ρρ{|xy|<ρ}ρ2|xy|2|xy|1+2sdydx\displaystyle C\tilde{a}_{\varepsilon}\left[\int^{-\rho}_{-2\rho}\int_{\{|x-y|<\rho\}}\frac{\rho^{-2\alpha}|x-y|^{2\alpha}}{|x-y|^{1+2s}}\,dy\,dx+\int^{-\rho}_{-2\rho}\int_{\{|x-y|<\rho\}}\frac{\rho^{-2}|x-y|^{2}}{|x-y|^{1+2s}}\,dy\,dx\right.
+2ρρ{|xy|>ρ}1|xy|1+2sdydx]\displaystyle\qquad\qquad\left.+\int^{-\rho}_{-2\rho}\int_{\{|x-y|>\rho\}}\frac{1}{|x-y|^{1+2s}}\,dy\,dx\right]
\displaystyle\leqslant Ca~ερ[ρ2(αs)ρ2α+ρ2(1s)ρ2+1ρ2s]=Ca~ερ2s1=o(1).\displaystyle C\tilde{a}_{\varepsilon}\rho\left[\frac{\rho^{2(\alpha-s)}}{\rho^{2\alpha}}+\frac{\rho^{2(1-s)}}{\rho^{2}}+\frac{1}{\rho^{2s}}\right]=\frac{C\tilde{a}_{\varepsilon}}{\rho^{2s-1}}=o(1).

Similarly,

a~εvε((2ρ,ρ/10),(2ρo,2ρ))Ca~ερ2s1=o(1).\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}((-2\rho,-\rho/10),(-2\rho_{o},-2\rho))\leqslant\frac{C\tilde{a}_{\varepsilon}}{\rho^{2s-1}}=o(1).

Lastly, by Lemma 9.3, we similarly estimate to find that

a~εvε((2ρo,2ρ),(2ρo,2ρ))\displaystyle\tilde{a}_{\varepsilon}v_{\varepsilon}((-2\rho_{o},-2\rho),(-2\rho_{o},-2\rho))
\displaystyle\leqslant Ca~ε[2ρo2ρ{|xy|<ρ}(2ρo,2ρ)ρ2|xy|2|xy|1+2sdydx\displaystyle C\tilde{a}_{\varepsilon}\left[\int^{-2\rho}_{-2\rho_{o}}\int_{\{|x-y|<\rho\}\cap(-2\rho_{o},-2\rho)}\frac{\rho^{-2}|x-y|^{2}}{|x-y|^{1+2s}}\,dy\,dx\right.
+2ρo2ρ{|xy|>ρ}(2ρo,2ρ)1|xy|1+2sdydx]\displaystyle\qquad\qquad\left.+\int^{-2\rho}_{-2\rho_{o}}\int_{\{|x-y|>\rho\}\cap(-2\rho_{o},-2\rho)}\frac{1}{|x-y|^{1+2s}}\,dy\,dx\right]
\displaystyle\leqslant C(ρoρ)a~ερ2s=o(1).\displaystyle C(\rho_{o}-\rho)\frac{\tilde{a}_{\varepsilon}}{\rho^{2s}}=o(1).

Putting together these pieces, we obtain from (9.26) that a~εvε((2ρo,0),(2ρo,ρ))=o(1)\tilde{a}_{\varepsilon}v_{\varepsilon}((-2\rho_{o},0),(-2\rho_{o},-\rho))=o(1), which is the desired claim in (9.25). The proof of Lemma 9.5 is thereby complete. ∎

Recalling (9.7), we combine Lemmata 9.1, 9.2, and 9.5 to obtain following estimate for the kinetic part of the energy:

Lemma 9.6.

It holds that

a~ε[vε(Ω,Ω)+vε(Ω,Ωc)]\displaystyle\tilde{a}_{\varepsilon}[v_{\varepsilon}(\Omega,\Omega)+v_{\varepsilon}(\Omega,\Omega^{c})] a~ε[vε(Ω,Ω)+2vε(Ω,(Ω)c)]\displaystyle\leqslant\tilde{a}_{\varepsilon}[v_{\varepsilon}(\Omega^{\prime},\Omega^{\prime})+2v_{\varepsilon}(\Omega^{\prime},(\Omega^{\prime})^{c})]
+Ψ1r(sgnE(x¯1),g(x¯1))+2Ψ2r(sgnE(x¯1),g(x¯1))\displaystyle\quad+\Psi_{1}^{r}(\operatorname{sgn}_{E}(\bar{x}_{1}),g(\bar{x}_{1}))+2\Psi_{2}^{r}(\operatorname{sgn}_{E}(\bar{x}_{1}),g(\bar{x}_{1}))
+Ψ1r(sgnE(x¯2),g(x¯2))+2Ψ2r(sgnE(x¯2),g(x¯2))+o(1).\displaystyle\quad+\Psi_{1}^{r}(\operatorname{sgn}_{E}(\bar{x}_{2}),g(\bar{x}_{2}))+2\Psi_{2}^{r}(\operatorname{sgn}_{E}(\bar{x}_{2}),g(\bar{x}_{2}))+o(1).

9.2. Potential energy estimates

We now estimate the potential energy. For this, recall that Ω=ΩΩ2ρo\Omega^{\prime}=\Omega\setminus\Omega_{2\rho_{o}} and write

ΩW(vε(x))𝑑x\displaystyle\int_{\Omega}W(v_{\varepsilon}(x))\,dx =ΩW(vε(x))𝑑x+Ω2ρoΩρW(vε(x))𝑑x\displaystyle=\int_{\Omega^{\prime}}W(v_{\varepsilon}(x))\,dx+\int_{\Omega_{2\rho_{o}}\setminus\Omega_{\rho}}W(v_{\varepsilon}(x))\,dx
+Bρ+(x¯1)W(vε(x))𝑑x+Bρ(x¯2)W(vε(x))𝑑x.\displaystyle\qquad\quad+\int_{B_{\rho}^{+}(\bar{x}_{1})}W(v_{\varepsilon}(x))\,dx+\int_{B_{\rho}^{-}(\bar{x}_{2})}W(v_{\varepsilon}(x))\,dx.

Near the boundary Ω\partial\Omega, we have the following.

Lemma 9.7.

If x¯Ω\bar{x}\in\partial\Omega, then

b~εBρ(x¯)ΩW(vε(x))dxbrBrW(w0(x;sgnE(x¯),g(x¯))dx,\tilde{b}_{\varepsilon}\int_{B_{\rho}(\bar{x})\cap\Omega}W(v_{\varepsilon}(x))\,dx\leqslant b_{r}\int_{B_{r}^{-}}W(w_{0}(x;\operatorname{sgn}_{E}(\bar{x}),g(\bar{x}))\,dx,

with brb_{r} given by (9.14).

Proof.

As in the proof of Lemma 9.1, assume without loss of generality that (9.15), (9.16), and (9.17) hold. Then, the result follows with the change of variables x~=x/ε\tilde{x}=x/\varepsilon:

b~εBρΩW(vε(x))𝑑x=b~ερ0W(w0(xε))𝑑x\displaystyle\tilde{b}_{\varepsilon}\int_{B_{\rho}\cap\Omega}W(v_{\varepsilon}(x))\,dx=\tilde{b}_{\varepsilon}\int_{-\rho}^{0}W\left(w_{0}\left(\frac{x}{\varepsilon}\right)\right)\,dx
=εb~ερε0W(w0(x~))𝑑x~=εb~εr0W(w0(x~))𝑑x~.\displaystyle\qquad\qquad=\varepsilon\tilde{b}_{\varepsilon}\int_{-\frac{\rho}{\varepsilon}}^{0}W\left(w_{0}(\tilde{x})\right)\,d\tilde{x}=\varepsilon\tilde{b}_{\varepsilon}\int_{-r}^{0}W\left(w_{0}(\tilde{x})\right)\,d\tilde{x}.

When s(12,1)s\in\left(\frac{1}{2},1\right), note that εb~ε=1=br\varepsilon\tilde{b}_{\varepsilon}=1=b_{r}, so the lemma holds with equality. If s=12s=\frac{1}{2}, then εb~ε=1/|lnε|\varepsilon\tilde{b}_{\varepsilon}=1/|\ln\varepsilon| and, using (9.18), the statement holds with an inequality. ∎

We now check the error terms.

Lemma 9.8.

It holds that

b~εΩ2ρoΩρW(vε(x))𝑑x=o(1).\tilde{b}_{\varepsilon}\int_{\Omega_{2\rho_{o}}\setminus\Omega_{\rho}}W(v_{\varepsilon}(x))\,dx=o(1).
Proof.

We split Ω2ρoΩρ=(Ω2ρoΩ2ρ)(Ω2ρΩρ)\Omega_{2\rho_{o}}\setminus\Omega_{\rho}=(\Omega_{2\rho_{o}}\setminus\Omega_{2\rho})\cup(\Omega_{2\rho}\setminus\Omega_{\rho}). First let xΩ2ρoΩ2ρx\in\Omega_{2\rho_{o}}\setminus\Omega_{2\rho}. Recalling (9.4), we use that W(±1)=0W(\pm 1)=0 and (6.3) to find that

W(vε(x))=W(u0(d~(x)ε))W(sgn(d~(x)ε))\displaystyle W(v_{\varepsilon}(x))=W\left(u_{0}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)\right)-W\left(\operatorname{sgn}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)\right)
C|u0(d~(x)ε)sgn(d~(x)ε)|Cε2sρo2s=Cε2s.\displaystyle\qquad\quad\leqslant C\left|u_{0}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)-\operatorname{sgn}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)\right|\leqslant C\frac{\varepsilon^{2s}}{\rho_{o}^{2s}}=C\varepsilon^{2s}.

Therefore,

(9.27) b~εΩ2ρoΩ2ρW(vε(x))𝑑xb~εΩ2ρoΩ2ρCε2s𝑑xCb~εε2s=o(1).\tilde{b}_{\varepsilon}\int_{\Omega_{2\rho_{o}}\setminus\Omega_{2\rho}}W(v_{\varepsilon}(x))\,dx\leqslant\tilde{b}_{\varepsilon}\int_{\Omega_{2\rho_{o}}\setminus\Omega_{2\rho}}C\varepsilon^{2s}\,dx\leqslant C\tilde{b}_{\varepsilon}\varepsilon^{2s}=o(1).

Now let xΩ2ρΩρx\in\Omega_{2\rho}\setminus\Omega_{\rho}. As a consequence of (9.4), we have that |d~(x)|>ρ|\tilde{d}(x)|>\rho. Using the regularity of WW, (1.11), and (6.3), we estimate

W\displaystyle W (vε(x))=W(2ρd(x)ρ[w0(d(x)ε)uε(x)]+uε(x))\displaystyle(v_{\varepsilon}(x))=W\left(\frac{2\rho-d(x)}{\rho}\,\left[w_{0}\left(\frac{-d(x)}{\varepsilon}\right)-u_{\varepsilon}(x)\right]+u_{\varepsilon}(x)\right)
W(uε(x))+C|2ρd(x)ρ[w0(d(x)ε)uε(x)]|\displaystyle\leqslant W(u_{\varepsilon}(x))+C\left|\frac{2\rho-d(x)}{\rho}\,\left[w_{0}\left(\frac{-d(x)}{\varepsilon}\right)-u_{\varepsilon}(x)\right]\right|
C|u0(d~(x)ε)sgn(d~(x)ε)|+C|w0(d(x)ε)sgn(d~(x)ε)|\displaystyle\leqslant C\left|u_{0}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)-\operatorname{sgn}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)\right|+C\left|w_{0}\left(\frac{-d(x)}{\varepsilon}\right)-\operatorname{sgn}\left(\frac{\tilde{d}(x)}{\varepsilon}\right)\right|
Cε2sρ2s,\displaystyle\leqslant\frac{C\varepsilon^{2s}}{\rho^{2s}},

where w0(x)=w0(x;sgn(d~(x)),g(π(x)))w_{0}(x)=w_{0}(x;\operatorname{sgn}(\tilde{d}(x)),g(\pi(x))). Therefore,

(9.28) b~εΩ2ρΩρW(vε(x))𝑑xb~εΩ2ρΩρCε2sρ2s𝑑x=Cb~εε2sρ2s1=o(1).\tilde{b}_{\varepsilon}\int_{\Omega_{2\rho}\setminus\Omega_{\rho}}W(v_{\varepsilon}(x))\,dx\leqslant\tilde{b}_{\varepsilon}\int_{\Omega_{2\rho}\setminus\Omega_{\rho}}\frac{C\varepsilon^{2s}}{\rho^{2s}}\,dx=\frac{C\tilde{b}_{\varepsilon}\varepsilon^{2s}}{\rho^{2s-1}}=o(1).

The lemma follows from (9.27) and (9.28). ∎

Combing the previous two lemmata, we have the following:

Lemma 9.9.

It holds that

b~εΩW(vε(x))𝑑x\displaystyle\tilde{b}_{\varepsilon}\int_{\Omega}W(v_{\varepsilon}(x))\,dx =b~εΩW(uε(x))dx+brBrW(w0(x;sgnE(x¯1),g(x¯1))dx\displaystyle=\tilde{b}_{\varepsilon}\int_{\Omega^{\prime}}W(u_{\varepsilon}(x))\,dx+b_{r}\int_{B_{r}^{-}}W(w_{0}(x;\operatorname{sgn}_{E}(\bar{x}_{1}),g(\bar{x}_{1}))\,dx
+brBrW(w0(x;sgnE(x¯2),g(x¯2))dx+o(1),\displaystyle\qquad\quad+b_{r}\int_{B_{r}^{-}}W(w_{0}(x;\operatorname{sgn}_{E}(\bar{x}_{2}),g(\bar{x}_{2}))\,dx+o(1),

with brb_{r} given by (9.14).

9.3. Proof of lim sup\limsup inequality

We are finally prepared to prove the lim sup\limsup inequality in Theorem 1.6. Let Ψ(±1,γ)\Psi(\pm 1,\gamma) be as in (7.3). Recalling (9.13) and (9.19), notice that

(9.29) Ψ(±1,γ)=limj+[Ψ1r(±1,γ)+2Ψ2r(±1,γ)+brBrW(w0(x;±1,γ))𝑑x].\Psi(\pm 1,\gamma)=\lim_{j\to+\infty}\left[\Psi_{1}^{r}(\pm 1,\gamma)+2\Psi_{2}^{r}(\pm 1,\gamma)+b_{r}\int_{B_{r}^{-}}W(w_{0}(x;\pm 1,\gamma))\,dx\right].

We recall that XX is the space of all the measurable functions u:nu:\mathbb{R}^{n}\to\mathbb{R} such that the restriction of uu to Ω\Omega belongs to L1(Ω)L^{1}(\Omega). Moreover, XX is endowed with the metric of L1(Ω)L^{1}(\Omega), as made clear in (1.14). Also, the space YκY_{\kappa} is defined in (1.7).

Proposition 9.10.

Assume that |g|<1|g|<1 on Ω\partial\Omega. Let κ[0,|Ω|2)\kappa\in\left[0,\frac{|\Omega|}{2}\right) and EE\subset\mathbb{R} be a measurable set.

Then, for any sequence εj0\varepsilon_{j}\searrow 0, there exists a sequence vjXv_{j}\in X such that vjχEχEc=:uv_{j}\to\chi_{E}-\chi_{E^{c}}=:u in XX and

lim supj+εj(1)(vj)cPer(E,Ω)+ΩΨ(u(x),g(x))𝑑0(x).\limsup_{j\to+\infty}{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(v_{j})\leqslant c_{\star}\,\operatorname{Per}(E,\Omega)+\int_{\partial\Omega}\Psi(u(x),g(x))\,d\mathcal{H}^{0}(x).

We stress that Proposition 9.10 holds also for κ=0\kappa=0.

Proof of Proposition 9.10.

We assume that (9.2) holds and uXYκu\in X\cap Y_{\kappa}, otherwise we are done.

Let Ω:=ΩΩ2ρo\Omega^{\prime}:=\Omega\setminus\Omega_{2\rho_{o}} and vj:=vεjv_{j}:=v_{\varepsilon_{j}}, where vεjv_{\varepsilon_{j}} is as in (9.5). Observe that vjχEχEc=uv_{j}\to\chi_{E}-\chi_{E^{c}}=u in XX. In the trace sense, u(x¯i)=sgnE(x¯i)u(\bar{x}_{i})=\operatorname{sgn}_{E}(\bar{x}_{i}) for i=1,2i=1,2.

By Lemmata 9.6 and 9.9, it holds that

εj(1)(vj)\displaystyle{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(v_{j}) a~ε[vε(Ω,Ω)+2vε(Ω,(Ω)c)]+b~εΩW(vε(x))𝑑x\displaystyle\leqslant\tilde{a}_{\varepsilon}[v_{\varepsilon}(\Omega^{\prime},\Omega^{\prime})+2v_{\varepsilon}(\Omega^{\prime},(\Omega^{\prime})^{c})]+\tilde{b}_{\varepsilon}\int_{\Omega^{\prime}}W(v_{\varepsilon}(x))\,dx
+Ψ1r(u(x¯1),g(x¯1))+2Ψ2r(u(x¯1),g(x¯1))+brBrW(w0(x;u(x¯1),g(x¯1)))𝑑x\displaystyle\quad+\Psi_{1}^{r}(u(\bar{x}_{1}),g(\bar{x}_{1}))+2\Psi_{2}^{r}(u(\bar{x}_{1}),g(\bar{x}_{1}))+b_{r}\int_{B_{r}^{-}}W(w_{0}(x;u(\bar{x}_{1}),g(\bar{x}_{1})))\,dx
+Ψ1r(u(x¯2),g(x¯2))+2Ψ2r(u(x¯2),g(x¯2))+brBrW(w0(x;u(x¯2),g(x¯2)))𝑑x\displaystyle\quad+\Psi_{1}^{r}(u(\bar{x}_{2}),g(\bar{x}_{2}))+2\Psi_{2}^{r}(u(\bar{x}_{2}),g(\bar{x}_{2}))+b_{r}\int_{B_{r}^{-}}W(w_{0}(x;u(\bar{x}_{2}),g(\bar{x}_{2})))\,dx
+o(1).\displaystyle\quad+o(1).

Now, by (9.3),

lim supj+[a~ε[vε(Ω,Ω)+2vε(Ω,(Ω)c)]+b~εΩW(vε(x))𝑑x]cPer(E,Ω).\limsup_{j\to+\infty}\left[\tilde{a}_{\varepsilon}\big{[}v_{\varepsilon}(\Omega^{\prime},\Omega^{\prime})+2v_{\varepsilon}(\Omega^{\prime},(\Omega^{\prime})^{c})\big{]}+\tilde{b}_{\varepsilon}\int_{\Omega^{\prime}}W(v_{\varepsilon}(x))\,dx\right]\leqslant c_{\star}\operatorname{Per}(E,\Omega^{\prime}).

Therefore, with (9.29), we have that

lim supj+εj(1)(vj)\displaystyle\limsup_{j\to+\infty}{\mathcal{F}}^{(1)}_{\varepsilon_{j}}(v_{j}) cPer(E,Ω)+Ψ(u(x¯1),g(x¯1))+Ψ(u(x¯2),g(x¯2)).\displaystyle\leqslant c_{\star}\operatorname{Per}(E,\Omega^{\prime})+\Psi(u(\bar{x}_{1}),g(\bar{x}_{1}))+\Psi(u(\bar{x}_{2}),g(\bar{x}_{2})).

Since Ω=ΩΩ2ρo\Omega^{\prime}=\Omega\setminus\Omega_{2\rho_{o}}, we send ρo0\rho_{o}\to 0 to obtain the desired result. ∎

9.4. Proof of Theorem 1.6

We can now complete the proof of Theorem 1.6.

Proof of Theorem 1.6.

By Propositions 8.3 and 9.10, we have that (1)=Γlimε0ε(1)\displaystyle\mathcal{F}^{(1)}=\Gamma-\lim_{\varepsilon\searrow 0}\mathcal{F}^{(1)}_{\varepsilon}. Together with Lemma 2.1 and Remark 2.2, this proves Theorem 1.6. ∎

Appendix A Some auxiliary results

Here, we collect some auxiliary results that are used in Section 7.

First, we show that minimizers uXγu\in X_{\gamma} of 𝒢s\mathcal{G}_{s} in (7.1) must be strictly less than γ\gamma in a left-sided neighborhood of the origin.

Lemma A.1.

Let s(12,1)s\in\left(\frac{1}{2},1\right) and γ(1,1)\gamma\in(-1,1). Suppose that uXγu\in X_{\gamma} is such that 1uγ-1\leqslant u\leqslant\gamma. Assume that u=γu=\gamma in (a,0)(a,0) for some a<0a<0 and define

w(x):={γifx>0u(x+a)ifx0.w(x):=\begin{cases}\gamma&\hbox{if}\leavevmode\nobreak\ x>0\\ u(x+a)&\hbox{if}\leavevmode\nobreak\ x\leqslant 0.\end{cases}

Then, 𝒢s(w)<𝒢s(u)\mathcal{G}_{s}(w)<\mathcal{G}_{s}(u).

Proof.

By a change of variables and using that uγu\equiv\gamma in (a,+)(a,+\infty), we have that

w(,)\displaystyle w(\mathbb{R}^{-},\mathbb{R}^{-}) =00|u(x+a)u(y+a)|2|xy|1+2s𝑑y𝑑x\displaystyle=\int_{-\infty}^{0}\int_{-\infty}^{0}\frac{|u(x+a)-u(y+a)|^{2}}{|x-y|^{1+2s}}\,dy\,dx
=aa|u(x¯)u(y¯)|2|x¯y¯|1+2s𝑑y¯𝑑x¯\displaystyle=\int_{-\infty}^{a}\int_{-\infty}^{a}\frac{|u(\bar{x})-u(\bar{y})|^{2}}{|\bar{x}-\bar{y}|^{1+2s}}\,d\bar{y}\,d\bar{x}
=u((,a),(,a))\displaystyle=u((-\infty,a),(-\infty,a))
=u(,)2u((,a),(a,0))u((a,0),(a,0))\displaystyle=u(\mathbb{R}^{-},\mathbb{R}^{-})-2u((-\infty,a),(a,0))-u((a,0),(a,0))
=u(,)2u((,a),(a,0))\displaystyle=u(\mathbb{R}^{-},\mathbb{R}^{-})-2u((-\infty,a),(a,0))

and

w(,+)\displaystyle w(\mathbb{R}^{-},\mathbb{R}^{+}) =00+|u(x+a)γ|2|xy|1+2s𝑑y𝑑x\displaystyle=\int_{-\infty}^{0}\int_{0}^{+\infty}\frac{|u(x+a)-\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx
=aa+|u(x¯)γ|2|x¯y¯|1+2s𝑑y¯𝑑x¯\displaystyle=\int_{-\infty}^{a}\int_{a}^{+\infty}\frac{|u(\bar{x})-\gamma|^{2}}{|\bar{x}-\bar{y}|^{1+2s}}\,d\bar{y}\,d\bar{x}
=00+|u(x¯)γ|2|x¯y¯|1+2s𝑑y¯𝑑x¯+aa0|u(x¯)u(y¯)|2|xy|1+2s𝑑y¯𝑑x¯\displaystyle=\int_{-\infty}^{0}\int_{0}^{+\infty}\frac{|u(\bar{x})-\gamma|^{2}}{|\bar{x}-\bar{y}|^{1+2s}}\,d\bar{y}\,d\bar{x}+\int_{-\infty}^{a}\int_{a}^{0}\frac{|u(\bar{x})-u(\bar{y})|^{2}}{|x-y|^{1+2s}}\,d\bar{y}\,d\bar{x}
=u(,+)+u((,a),(a,0)).\displaystyle=u(\mathbb{R}^{-},\mathbb{R}^{+})+u((-\infty,a),(a,0)).

Therefore, the kinetic energy for ww satisfies

w(,)+2w(,+)=u(,)+2u(,+).w(\mathbb{R}^{-},\mathbb{R}^{-})+2w(\mathbb{R}^{-},\mathbb{R}^{+})=u(\mathbb{R}^{-},\mathbb{R}^{-})+2u(\mathbb{R}^{-},\mathbb{R}^{+}).

On the other hand, since γ(1,1)\gamma\in(-1,1), we have that W(γ)>0W(\gamma)>0. In particular, |a|W(γ)>0|a|W(\gamma)>0, so that

0W(w(x))𝑑x=0W(u(x+a))𝑑x=aW(u(x¯))𝑑x¯\displaystyle\int_{-\infty}^{0}W(w(x))\,dx=\int_{-\infty}^{0}W(u(x+a))\,dx=\int_{-\infty}^{a}W(u(\bar{x}))\,d\bar{x}
<aW(u(x¯))𝑑x¯+|a|W(γ)=0W(u(x))𝑑x.\displaystyle\qquad\qquad<\int_{-\infty}^{a}W(u(\bar{x}))\,d\bar{x}+|a|W(\gamma)=\int_{-\infty}^{0}W(u(x))\,dx.

The result follows from the previous two displays. ∎

Next, we present a useful competitor for energy estimates in Section 7.

Lemma A.2.

Let s(12,1)s\in\left(\frac{1}{2},1\right) and 𝒢s\mathcal{G}_{s} as in (7.1). Define the function

h(x):={1 if x2,(γ+1)x+2γ+1 if x(2,1),γ if x1.h(x):=\begin{cases}-1&\hbox{ if }x\leqslant-2,\\ (\gamma+1)x+2\gamma+1&\hbox{ if }x\in(-2,-1),\\ \gamma&\hbox{ if }x\geqslant-1.\end{cases}

Then, there exists a constant C>0C>0 such that, for all s(12,1)s\in\left(\frac{1}{2},1\right),

(A.1) 𝒢s(h)C(1+12s1).\mathcal{G}_{s}(h)\leqslant C\left(1+\frac{1}{2s-1}\right).

Moreover, there exists a constant C>0C>0 such that, for all R>3R>3,

(A.2) lims12𝒢s(h,[R,0])C(1+lnR).\lim_{s\searrow\frac{1}{2}}\mathcal{G}_{s}(h,[-R,0])\leqslant C(1+\ln R).
Proof.

To start, we calculate some integrals that will be used to prove both (A.1) and (A.2). In the following, CC is an arbitrary constant, independent of both RR and ss.

Observe that

(A.3) 2110|h(x)h(y)|2|xy|1+2s𝑑y𝑑x=2110(γ+1)2|x+1|2(yx)1+2s𝑑y𝑑x\displaystyle\int_{-2}^{-1}\int_{-1}^{0}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx=\int_{-2}^{-1}\int_{-1}^{0}\frac{(\gamma+1)^{2}|x+1|^{2}}{(y-x)^{1+2s}}\,dy\,dx
=(γ+1)221|x+1|2101(yx)1+2s𝑑y𝑑x\displaystyle\qquad\quad=(\gamma+1)^{2}\int_{-2}^{-1}|x+1|^{2}\int_{-1}^{0}\frac{1}{(y-x)^{1+2s}}\,dy\,dx
=(γ+1)22s21|x+1|2[1|x+1|2s1|x|2s]𝑑x\displaystyle\qquad\quad=\frac{(\gamma+1)^{2}}{2s}\int_{-2}^{-1}|x+1|^{2}\left[\frac{1}{|x+1|^{2s}}-\frac{1}{|x|^{2s}}\right]\,dx
(γ+1)22s21|x+1|22s𝑑xC\displaystyle\qquad\quad\leqslant\frac{(\gamma+1)^{2}}{2s}\int_{-2}^{-1}|x+1|^{2-2s}\,dx\leqslant C

and

(A.4) 2121|h(x)h(y)|2|xy|1+2s𝑑y𝑑x=2121(γ+1)2|xy|2|xy|1+2s𝑑y𝑑x\displaystyle\int_{-2}^{-1}\int_{-2}^{-1}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx=\int_{-2}^{-1}\int_{-2}^{-1}\frac{(\gamma+1)^{2}|x-y|^{2}}{|x-y|^{1+2s}}\,dy\,dx
=(γ+1)22121|xy|12s𝑑y𝑑xC.\displaystyle\qquad\qquad=(\gamma+1)^{2}\int_{-2}^{-1}\int_{-2}^{-1}|x-y|^{1-2s}\,dy\,dx\leqslant C.

Next, we find

(A.5) 221|1h(y)|2(yx)1+2s𝑑y𝑑x=221(γ+1)2|y+2|2(yx)1+2s𝑑y𝑑x\displaystyle\int_{-\infty}^{-2}\int_{-2}^{-1}\frac{|-1-h(y)|^{2}}{(y-x)^{1+2s}}\,dy\,dx=\int_{-\infty}^{-2}\int_{-2}^{-1}\frac{(\gamma+1)^{2}|y+2|^{2}}{(y-x)^{1+2s}}\,dy\,dx
=(γ+1)221|y+2|2[2dx(yx)1+2s]𝑑y\displaystyle\qquad\qquad=(\gamma+1)^{2}\int_{-2}^{-1}|y+2|^{2}\left[\int_{-\infty}^{-2}\frac{dx}{(y-x)^{1+2s}}\right]\,dy
=(γ+1)22s21|y+2|22s𝑑yC.\displaystyle\qquad\qquad=\frac{(\gamma+1)^{2}}{2s}\int_{-2}^{-1}|y+2|^{2-2s}\,dy\leqslant C.

Finally, we estimate

(A.6) 200+|h(x)γ|2|xy|1+2s𝑑y𝑑x=210+(γ+1)2|x+1|2(yx)1+2s𝑑y𝑑x\displaystyle\int_{-2}^{0}\int_{0}^{+\infty}\frac{|h(x)-\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx=\int_{-2}^{-1}\int_{0}^{+\infty}\frac{(\gamma+1)^{2}|x+1|^{2}}{(y-x)^{1+2s}}\,dy\,dx
=(γ+1)221|x+1|20+1(yx)1+2s𝑑y𝑑x\displaystyle\qquad\qquad=(\gamma+1)^{2}\int_{-2}^{-1}|x+1|^{2}\int_{0}^{+\infty}\frac{1}{(y-x)^{1+2s}}\,dy\,dx
=(γ+1)22s21|x+1|2|x|2s𝑑xC.\displaystyle\qquad\qquad=\frac{(\gamma+1)^{2}}{2s}\int_{-2}^{-1}\frac{|x+1|^{2}}{|x|^{2s}}\,dx\leqslant C.

We now prove (A.1). We begin with the interactions in ×\mathbb{R}^{-}\times\mathbb{R}^{-}. Using (A.3), (A.4), and (A.5), we find that

h(,)\displaystyle h(\mathbb{R}^{-},\mathbb{R}^{-}) =2220|1h(y)|2(yx)1+2s𝑑y𝑑x+2020|h(x)h(y)|2|xy|1+2s𝑑y𝑑x\displaystyle=2\int_{-\infty}^{-2}\int_{-2}^{0}\frac{|-1-h(y)|^{2}}{(y-x)^{1+2s}}\,dy\,dx+\int_{-2}^{0}\int_{-2}^{0}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx
=2221|1h(y)|2(yx)1+2s𝑑y𝑑x+2210(1+γ)2(yx)1+2s𝑑y𝑑x\displaystyle=2\int_{-\infty}^{-2}\int_{-2}^{-1}\frac{|-1-h(y)|^{2}}{(y-x)^{1+2s}}\,dy\,dx+2\int_{-\infty}^{-2}\int_{-1}^{0}\frac{(1+\gamma)^{2}}{(y-x)^{1+2s}}\,dy\,dx
+2121|h(x)h(y)|2|xy|1+2s𝑑y𝑑x+22110|h(x)γ|2(yx)1+2s𝑑y𝑑x\displaystyle\quad+\int_{-2}^{-1}\int_{-2}^{-1}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx+2\int_{-2}^{-1}\int_{-1}^{0}\frac{|h(x)-\gamma|^{2}}{(y-x)^{1+2s}}\,dy\,dx
=2210(1+γ)2(yx)1+2s𝑑y𝑑x+C.\displaystyle=2\int_{-\infty}^{-2}\int_{-1}^{0}\frac{(1+\gamma)^{2}}{(y-x)^{1+2s}}\,dy\,dx+C.

Since

210(γ+1)2(yx)1+2s𝑑y𝑑x\displaystyle\int_{-\infty}^{-2}\int_{-1}^{0}\frac{(\gamma+1)^{2}}{(y-x)^{1+2s}}\,dy\,dx (γ+1)22s2dx|1+x|2s(γ+1)22s(2s1)C2s1,\displaystyle\leqslant\frac{(\gamma+1)^{2}}{2s}\int_{-\infty}^{-2}\frac{dx}{|1+x|^{2s}}\leqslant\frac{(\gamma+1)^{2}}{2s(2s-1)}\leqslant\frac{C}{2s-1},

we arrive at

h(,)C(12s1+1).h(\mathbb{R}^{-},\mathbb{R}^{-})\leqslant C\left(\frac{1}{2s-1}+1\right).

Next, we consider the interactions in ×+\mathbb{R}^{-}\times\mathbb{R}^{+}. For this, we use (A.6) to estimate

00+|h(x)γ|2|xy|1+2s𝑑y𝑑x\displaystyle\int_{-\infty}^{0}\int_{0}^{+\infty}\frac{|h(x)-\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx
=20+(1+γ)2(yx)1+2s𝑑y𝑑x+210+|h(x)γ|2(xy)1+2s𝑑y𝑑x\displaystyle\quad=\int_{-\infty}^{-2}\int_{0}^{+\infty}\frac{(1+\gamma)^{2}}{(y-x)^{1+2s}}\,dy\,dx+\int_{-2}^{-1}\int_{0}^{+\infty}\frac{|h(x)-\gamma|^{2}}{(x-y)^{1+2s}}\,dy\,dx
=(1+γ)22s2dx|x|2s+C=(1+γ)22s(2s1)212s+C\displaystyle\quad=\frac{(1+\gamma)^{2}}{2s}\int_{-\infty}^{-2}\frac{dx}{|x|^{2s}}+C\ =\frac{(1+\gamma)^{2}}{2s(2s-1)}2^{1-2s}+C
C(12s1+1).\displaystyle\quad\leqslant C\left(\frac{1}{2s-1}+1\right).

Lastly, for the potential energy, we simply notice that

0W(h(x))𝑑x=20W(h(x))𝑑xC.\int_{-\infty}^{0}W(h(x))\,dx=\int_{-2}^{0}W(h(x))\,dx\leqslant C.

Combining the last three displays gives (A.1).

Next, we prove (A.2). Beginning with the interactions in the container [R,0]×[R,0][-R,0]\times[-R,0], we use (A.3), (A.4), and (A.5) to find

(A.7) R0R0|h(x)h(y)|2|xy|1+2s𝑑y𝑑x=2R221|1+h(y)|2|xy|1+2s𝑑y𝑑x+2R210|1+γ|2|xy|1+2s𝑑y𝑑x+2121|h(x)h(y)|2|xy|1+2s𝑑y𝑑x+22110|h(x)γ|2|xy|1+2s𝑑y𝑑x2R210|1+γ|2|xy|1+2s𝑑y𝑑x+C.\begin{split}&\int_{-R}^{0}\int_{-R}^{0}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &=2\int_{-R}^{-2}\int_{-2}^{-1}\frac{|1+h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx+2\int_{-R}^{-2}\int_{-1}^{0}\frac{|1+\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &\quad+\int_{-2}^{-1}\int_{-2}^{-1}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx+2\int_{-2}^{-1}\int_{-1}^{0}\frac{|h(x)-\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx\\ &\leqslant 2\int_{-R}^{-2}\int_{-1}^{0}\frac{|1+\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx+C.\end{split}

Observe that, exploiting (6.7) (with a:=1a:=1 and b:=2b:=2 and with a:=R1a:=R-1 and b:=Rb:=R),

lims12R210(γ+1)2(yx)1+2s𝑑y𝑑x\displaystyle\lim_{s\searrow\frac{1}{2}}\int_{-R}^{-2}\int_{-1}^{0}\frac{(\gamma+1)^{2}}{(y-x)^{1+2s}}\,dy\,dx
=lims12(γ+1)22s(2s1)[(1212s)((R1)12sR12s)]\displaystyle=\lim_{s\searrow\frac{1}{2}}\frac{(\gamma+1)^{2}}{2s(2s-1)}\Big{[}(1-2^{1-2s})-\big{(}(R-1)^{1-2s}-R^{1-2s}\big{)}\Big{]}
=(γ+1)2[ln2ln(RR1)]C.\displaystyle=(\gamma+1)^{2}\left[\ln 2-\ln\left(\frac{R}{R-1}\right)\right]\leqslant C.

This and (A.7) give that

(A.8) lims12R0R0|h(x)h(y)|2|xy|1+2s𝑑x𝑑yC.\lim_{s\searrow\frac{1}{2}}\int_{-R}^{0}\int_{-R}^{0}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dx\,dy\leqslant C.

Now, we estimate the interactions in [R,0]×[R,0]c[-R,0]\times[-R,0]^{c}. With (A.6), we have that

R00+|h(x)γ|2|xy|1+2s𝑑y𝑑xR20+|1+γ|2|xy|1+2s𝑑y𝑑x+C.\int_{-R}^{0}\int_{0}^{+\infty}\frac{|h(x)-\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx\leqslant\int_{-R}^{-2}\int_{0}^{+\infty}\frac{|1+\gamma|^{2}}{|x-y|^{1+2s}}\,dy\,dx+C.

Using again (6.7) (with a:=2a:=2 and b:=Rb:=R), we next observe that

lims12R20+|h(x)h(y)|2|xy|1+2s𝑑y𝑑x=lims12R20+(γ+1)2(yx)1+2s𝑑y𝑑x\displaystyle\lim_{s\searrow\frac{1}{2}}\int_{-R}^{-2}\int_{0}^{+\infty}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx=\lim_{s\searrow\frac{1}{2}}\int_{-R}^{-2}\int_{0}^{+\infty}\frac{(\gamma+1)^{2}}{(y-x)^{1+2s}}\,dy\,dx
=lims12(γ+1)22s(2s1)(212sR12s)=(γ+1)2(lnRln2)ClnR.\displaystyle\qquad\quad=\lim_{s\searrow\frac{1}{2}}\frac{(\gamma+1)^{2}}{2s(2s-1)}\big{(}2^{1-2s}-R^{1-2s}\big{)}=(\gamma+1)^{2}\big{(}\ln R-\ln 2\big{)}\leqslant C\ln R.

For the remaining term, we use that h[1,γ]h\in[-1,\gamma] and that R3R\geqslant 3 to find that

R0R|h(x)h(y)|2|xy|1+2s𝑑y𝑑x=20R|h(x)h(y)|2(xy)1+2s𝑑y𝑑x\displaystyle\int_{-R}^{0}\int_{-\infty}^{-R}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dy\,dx=\int_{-2}^{0}\int_{-\infty}^{-R}\frac{|h(x)-h(y)|^{2}}{(x-y)^{1+2s}}\,dy\,dx
203(1+γ)2(xy)1+2s𝑑y𝑑x=(1+γ)22s201(x+3)2s𝑑xC.\displaystyle\qquad\quad\leqslant\int_{-2}^{0}\int_{-\infty}^{-3}\frac{(1+\gamma)^{2}}{(x-y)^{1+2s}}\,dy\,dx=\frac{(1+\gamma)^{2}}{2s}\int_{-2}^{0}\frac{1}{(x+3)^{2s}}\,dx\leqslant C.

Gathering these pieces of information, we see that

(A.9) lims12[R,0][R,0]c|h(x)h(y)|2|xy|1+2s𝑑x𝑑yC(1+lnR).\lim_{s\searrow\frac{1}{2}}\int_{[-R,0]}\int_{[-R,0]^{c}}\frac{|h(x)-h(y)|^{2}}{|x-y|^{1+2s}}\,dx\,dy\leqslant C(1+\ln R).

The inequality in (A.2) follows from (A.8), (A.9), and the simply observing that

lims12R0W(h(x))𝑑x=20W(h(x))𝑑xC.\lim_{s\searrow\frac{1}{2}}\int_{-R}^{0}W(h(x))\,dx=\int_{-2}^{0}W(h(x))\,dx\leqslant C.\qed

Acknowledgments

It is a pleasure to thank Hidde Schönberger for fruitful discussions. SD has been supported by the Australian Future Fellowship FT230100333 “New perspectives on nonlocal equations”. SP has been supported by the NSF Grant DMS-2155156 “Nonlinear PDE methods in the study of interphases.” EV and MV have been supported by the Australian Laureate Fellowship FL190100081 “Minimal surfaces, free boundaries and partial differential equations.”

References