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

Existence and rigidity of the vectorial Peierls-Nabarro model for dislocations in high dimensions

Yuan Gao Department of Mathematics, Duke University, Durham, NC and Department of Mathematics, Purdue University, West Lafayette, IN [email protected] Jian-Guo Liu Department of Mathematics and Department of Physics, Duke University, Durham, NC [email protected]  and  Zibu Liu Department of Mathematics, Duke University, Durham, NC [email protected]
Abstract.

We focus on the existence and rigidity problems of the vectorial Peierls-Nabarro (PN) model for dislocations. Under the assumption that the misfit potential on the slip plane only depends on the shear displacement along the Burgers vector, a reduced non-local scalar Ginzburg-Landau equation with an anisotropic positive (if Poisson ratio belongs to (1/2,1/3)(-1/2,1/3)) singular kernel is derived on the slip plane. We first prove that minimizers of the PN energy for this reduced scalar problem exist. Starting from H1/2H^{1/2} regularity, we prove that these minimizers are smooth 1D profiles only depending on the shear direction, monotonically and uniformly converge to two stable states at far fields in the direction of the Burgers vector. Then a De Giorgi-type conjecture of single-variable symmetry for both minimizers and layer solutions is established. As a direct corollary, minimizers and layer solutions are unique up to translations. The proof of this De Giorgi-type conjecture relies on a delicate spectral analysis which is especially powerful for nonlocal pseudo-differential operators with strong maximal principle. All these results hold in any dimension since we work on the domain periodic in the transverse directions of the slip plane. The physical interpretation of this rigidity result is that the equilibrium dislocation on the slip plane only admits shear displacements and is a strictly monotonic 1D profile provided exclusive dependence of the misfit potential on the shear displacement.

Key words and phrases:
fractional Laplacian, De Giorgi conjecture, stable entire solution, anisotropic nonlocal operator, spectral analysis, energy rearrangement, Peierls-Nabarro model, plastic deformation
2010 Mathematics Subject Classification:
35A02, 35Q74, 35S15, 35J50

1. Introduction and main results

In materials science, the Peierls-Nabarro (PN) model with Poisson ratio ν[1,1/2]\nu\in[-1,1/2] plays a fundamental role in describing dislocations or line defects in materials [6, 27]. Understanding this model provides insights on designing new materials with robust performance [24, 21, 8, 16]. However, the existence and rigidity problem regarding the vector-field PN model has not been explored.

The PN model is a nonlinear model that studies the core structure of the dislocation by incorporating the atomistic effect in the dislocation core into the continuum elastic model. In the PN model in three dimensions, two half-spaces separated by the slip plane of a dislocation are assumed to be linear elastic continua. Here the slip plane is assumed to be a fixed plane Γ={(x,y,z):y=0}\Gamma=\{(x,y,z):\ y=0\}, where the horizontal displacement discontinuity (known as disregistry) happens. Without loss of generality, we assume that the Burgers vector is 𝒃=(b,0,0)\bm{b}=(b,0,0) where b>0b>0. The magnitude of the Burgers vector represents the typical length to observe a heavily distorted region in the dislocation core. Hence it is natural to rescale all the quantities including spatial variable x,y,zx,y,z, the displacement vector 𝒖=(u1,u2,u3)\bm{u}=(u_{1},u_{2},u_{3}) and 𝒃\bm{b} with respect to the magnitude of the Burgers vector. After rescaling, we regard all these quantities (with same notations) as dimensionless quantities and b=1b=1.

In this paper, the shear direction is referred to as the direction of the Burgers vector, i.e. the xx direction; the vertical direction of the slip plane is referred to as the yy direction and the transverse direction in the slip plane is referred to as the zz direction.

The PN model is a minimization problem for the total energy EE which is given by

(1.1) E(𝒖):=Eels(𝒖)+Emis(𝒖).E(\bm{u}):=E_{\mathrm{els}}(\bm{u})+E_{\mathrm{mis}}(\bm{u}).

Here 𝒖=(u1,u2,u3)\bm{u}=(u_{1},u_{2},u_{3}) is the displacement vector. (1.1) incorporates not only the elastic energy in the bulk but also the atomistic effect in the dislocation core. The elastic energy in the two half-spaces is defined as

(1.2) Eels(𝒖)=3\Γ12σ:εdxdydz,\displaystyle E_{\mathrm{els}}(\bm{u})=\int_{\mathbb{R}^{3}\backslash\Gamma}\frac{1}{2}\sigma:\varepsilon\,\mathrm{d}x\,\mathrm{d}y\,\mathrm{d}z,

where σ:ε=i,j=13σijεij\sigma:\varepsilon=\sum_{i,j=1}^{3}\sigma_{ij}\varepsilon_{ij}. Here ε\varepsilon and σ\sigma are the strain tensor and the stress tensor respectively, defined as

(1.3) εij=12(jui+iuj),σij=2Gεij+2νG12νk=13εkkδij,i,j=1,2,3.\varepsilon_{ij}=\frac{1}{2}(\partial_{j}u_{i}+\partial_{i}u_{j}),\ \sigma_{ij}=2G\varepsilon_{ij}+\frac{2\nu G}{1-2\nu}\sum_{k=1}^{3}\varepsilon_{kk}\delta_{ij},\quad i,j=1,2,3.

Here ν[1,1/2]\nu\in[-1,1/2] is the Poisson ratio and GG is the shear modulus.

On the slip plane, we denote the upper limit and lower limit of the displacement as

(1.4) ui+(x,z)=ui(x,0+,z),ui(x,z)=ui(x,0,z),i=1,2,3.\displaystyle u_{i}^{+}(x,z)=u_{i}(x,0^{+},z),\ u_{i}^{-}(x,z)=u_{i}(x,0^{-},z),\ i=1,2,3.

Moreover, we assume that ui,i=1,2,3u_{i},\ i=1,2,3 are subject to the following boundary conditions at the slip plane:

(1.5) u1+(x,z)=u1(x,z),u2+(x,z)=u2(x,z),u3+(x,z)=u3(x,z).\displaystyle u_{1}^{+}(x,z)=-u_{1}^{-}(x,z),\,\,\,u_{2}^{+}(x,z)=u_{2}^{-}(x,z),\,\,\,u_{3}^{+}(x,z)=-u_{3}^{-}(x,z).

We call (1.5) the symmetric assumption. Characterizing the nonlinear atomistic interactions, the misfit energy Emis(𝒖)E_{\mathrm{mis}}(\bm{u}) is defined as the integral of the misfit potential γ:2\gamma:\mathbb{R}^{2}\to\mathbb{R} on the slip plane:

(1.6) Emis(𝒖):=Γγ(u1+u1,u3+u3)dxdz=Γγ(2u1+,2u3+)dxdz.{E_{\mathrm{mis}}}(\bm{u}):=\int_{\Gamma}\gamma(u_{1}^{+}-u_{1}^{-},u_{3}^{+}-u_{3}^{-})\,\mathrm{d}x\,\mathrm{d}z=\int_{\Gamma}\gamma(2u_{1}^{+},2u_{3}^{+})\,\mathrm{d}x\,\mathrm{d}z.

The last equality is due to the symmetric assumption (1.5). Notice, 𝒖\bm{u} is already dimensionless quantity so EmisE_{\mathrm{mis}} is well-defined. For brevity, we will omit factor 2 in (1.6) before u1+u_{1}^{+} and u3+u_{3}^{+} which makes no difference in the conclusions.

In this paper, to characterize the key property imposed by the Burgers vector, i.e., the direction of the dislocation and existence of two stable states, we assume that γ\gamma depends only on the shear displacement in the Burgers direction, i.e.,

(1.7) γ(u1,u3)=γ(u1),\displaystyle\gamma(u_{1},u_{3})=\gamma(u_{1}),

where the misfit potential is a double-well type potential, i.e. γ:\gamma:\mathbb{R}\to\mathbb{R} is a C2C^{2} function satisfying

(1.8) γ(x)>0if1<x<1,γ(±1)=0,γ′′(±1)>0.\gamma(x)>0\ \mathrm{if}\ {-1<x<1},\quad\gamma(\pm 1)=0,\ \ \gamma^{\prime\prime}(\pm 1)>0.

We remark that this assumption on the misfit potential also includes some other typical periodic potentials which satisfy γ(x+2)=γ(x)\gamma(x+2)=\gamma(x) and represent the periodic lattice structure of crystalline materials; see γ0\gamma_{0} in an explicit example (1.16).

Because the magnitude of the rescaled Burgers vector is order 11, for convenience, we take u1=±1u_{1}=\pm 1 as bi-states at far fields, i.e.,

(1.9) limx1±u1(𝒙)=±1,\lim\limits_{x_{1}\to\pm\infty}u_{1}(\bm{x})=\pm 1,

which equivalent to the magnitude of the dislocation is assumed to be b=4b=4.

In this bi-states case (1.9), the total energy EE in (1.1) in the whole space is always infinite. Therefore, we consider the global minimizer in the following perturbed sense. However we remark that for a dislocation loop, i.e., a disregistry with compact support instead of the bi-states far field condition, energy EE in (1.1) is finite.

Definition 1.

A function 𝐮:33\bm{u}:\mathbb{R}^{3}\to\mathbb{R}^{3} satisfying (1.9) is called a global minimizer of EE defined in (1.1) if it satisfies

(1.10) E(𝒖+𝝋)E(𝒖)0E(\bm{u}+\bm{\varphi})-E(\bm{u})\geq 0

for any perturbation 𝛗=(φ1,φ2,φ3)C(3\Γ;3)\bm{\varphi}=(\varphi_{1},\varphi_{2},\varphi_{3})\in C^{\infty}(\mathbb{R}^{3}\backslash\Gamma;\mathbb{R}^{3}) supported in some ball B(R)3B(R)\subset\mathbb{R}^{3} satisfying (1.5), i.e.

(1.11) φ1+(x,z)=φ1(x,z),φ2+(x,z)=φ2(x,z),φ3+(x,z)=φ3(x,z).\varphi_{1}^{+}(x,z)=-\varphi_{1}^{-}(x,z),\ \varphi_{2}^{+}(x,z)=\varphi_{2}^{-}(x,z),\ \varphi_{3}^{+}(x,z)=-\varphi_{3}^{-}(x,z).

The problem of existence and rigidity for the PN model interests us the most:

  1. (i)

    Does the minimizer of total energy (1.1) in the sense of Definition 1 exist?

  2. (ii)

    Do minimizers in (i) and layer solutions (see Definition 3) have 1D symmetry on the slip plane, i.e. are only depending on the shear direction, but independent with the transverse direction?

The answers to these two questions are both positive. To provide explicit and complete answers to these two questions, we consider the resulting Euler-Lagrange equation satisfied by the minimizer, which is a Lamé system with nonlinear boundary conditions on the slip plane (see (1.13)). Because we assume (1.7), i.e., the misfit potential γ\gamma depends only on the shear displacement u1u_{1}, this Euler-Lagrange equation is reduced to a nonlocal semi-linear scalar equation on the slip plane with an elliptic pseudo-differential operator of order 11 (see (1.15)). In particular, when ν(1/2,1/3)\nu\in(-1/2,1/3), the pseudo-differential operator can be described in the singular kernel formulation; see Assumptions (A)-(D).

After these simplifications and reformulation, we only need to focus on the existence and rigidity of this reduced scalar nonlocal equation (see (1.23)). This equation is the Euler-Lagrange equation of a reduced energy function FF on the slip plane (see (1.37)). We will first prove that minimizers of energy functional FF (see (1.37)) in set (1.35) exist by constructing a minimizing sequence in which each function is an H1/2H^{1/2} perturbation of a given 1D profile; see Theorem 1. Although starting from this weak regularity, we finally prove that these minimizers are smooth 1D profiles that monotonically and uniformly converge to stable states of the misfit functional γ\gamma in the shear direction, i.e., they converge to ±1\pm 1 as x±x\to\pm\infty.

After proving Theorem 1, we also establish a rigidity result of De Giorgi-type conjecture on 1D symmetry for all minimizers in set (1.35), and more generally for all layer solutions (see Definition 3). As a corollary, the uniqueness of these minimizers, as well as layer solutions, is also demonstrated; see Theorem 2 and Theorem 3. The existence and rigidity results are also stated for the original vectorial PN model (1.13) in Theorem 4.

Our results on both existence and rigidity hold in any dimension d1d\geq 1 due to the periodic assumption: we are interested in solutions that are periodic in d1d-1 transverse directions. This dimension-independent rigidity is also observed in other equations if the domain is armed with periodicity [22].

In terms of materials science, our results provide a compatible physical interpretation. For Poisson ratio ν(1/2,1/3)\nu\in(-1/2,1/3), if the misfit potential γ\gamma depends only on the shear displacement, then the equilibrium dislocation profile only admits shear displacements on the slip plane. Furthermore, this uniquely (up to translations) determined shear displacement is a strictly monotonic 1D profile connecting two stable states. In view of this rigidity result, the vectorial PN model (1.13) in three dimensions is reduced to a two-dimensional problem which was thoroughly investigated in our previous work [19].

In the remaining parts of the introduction, we will introduce the vectorial PN model and its reduced scalar equation (see (1.15)) in Section 1.1. From the reduced scalar equation, in Section 1.3, we introduce the nonlocal high-dimensional equation (see (1.23)) which contains the PN model as a special case. Finally, in this general context, we will present our main results and strategies in Section 1.4 to provide a rigorous and complete answer to the main questions (i) and (ii).

1.1. The vectorial Peierls-Nabarro model and its reduced scalar equation

Denote the unit torus /\mathbb{R}/\mathbb{Z} as 𝕋\mathbb{T}. Instead of minimizing the total energy (1.1) on 3\mathbb{R}^{3}, we consider the model on 2×𝕋\mathbb{R}^{2}\times\mathbb{T}. Correspondingly, the slip plane Γ\Gamma is replaced by Γ\Gamma^{\prime} which is defined as

(1.12) Γ={(x,y,z)2×𝕋:y=0}.\displaystyle\Gamma^{\prime}=\{(x,y,z)\in\mathbb{R}^{2}\times\mathbb{T}:\ y=0\}.

A standard calculation of the first variation of the total energy (1.1) derives the following Euler–Lagrange equation satisfied by minimizers of (1.1) in the sense of Definition 1. The proof of this lemma can be found in Appendix B or our previous work [19].

Lemma 1.1.

Assume that 𝐮C2(2×𝕋\Γ)\bm{u}\in C^{2}(\mathbb{R}^{2}\times\mathbb{T}\backslash\Gamma^{\prime}) is a minimizer of the total energy EE in the sense of Definition 1 satisfying the boundary conditions (1.5). Then 𝐮\bm{u} satisfies the Euler–Lagrange equation

(1.13) {Δ𝒖+112ν(𝒖)=0,in2×𝕋Γ,σ12++σ12=γu1(u1+,u3+),onΓ,σ22+=σ22,onΓ,σ32++σ32=γu3(u1+,u3+),onΓ.\displaystyle\begin{cases}\Delta\bm{u}+\dfrac{1}{1-2\nu}\nabla(\nabla\cdot\bm{u})=0,\ &\mathrm{in\ }\mathbb{R}^{2}\times\mathbb{T}\setminus\Gamma^{\prime},\\ \sigma^{+}_{12}+\sigma^{-}_{12}=\dfrac{\partial\gamma}{\partial u_{1}}(u_{1}^{+},u_{3}^{+}),\ &\mathrm{on\ }\Gamma^{\prime},\\ \sigma^{+}_{22}=\sigma^{-}_{22},\ &\mathrm{on\ }\Gamma^{\prime},\\ \sigma^{+}_{32}+\sigma^{-}_{32}=\dfrac{\partial\gamma}{\partial u_{3}}(u_{1}^{+},u_{3}^{+}),\ &\mathrm{on\ }\Gamma^{\prime}.\\ \end{cases}
Remark 1.

We can view (1.13), especially the second and the fourth equations as an incorporation of the linear response theory. Moreover, they are coupled equations of u1u_{1} and u3u_{3}. Notice that taking trace in (1.3), σ12=G(1u2+2u1)\sigma_{12}=G(\partial_{1}u_{2}+\partial_{2}u_{1}) on Γ\Gamma^{\prime}. (i) Regard the elastic bulks 3\Γ\mathbb{R}^{3}\backslash\Gamma^{\prime} as an environment and the slip plane Γ\Gamma as an open system. (ii) Given a Dirichlet disregistry boundary condition u1+,u3+u^{+}_{1},u^{+}_{3}, by solving σ=0\nabla\cdot\sigma=0 in the environment, one can obtain the trace σ12,σ32\sigma_{12},\sigma_{32} on Γ\Gamma^{\prime}. (iii) We call this operator (u1+,u3+)(σ12±,σ32±)(u^{+}_{1},u^{+}_{3})\mapsto(\sigma_{12}^{\pm},\sigma_{32}^{\pm}) the Dirichlet to Neumann map; also known as a nonlocal linear response operator. As a consequence, this enables us to consider a nonlocal semi-linear elliptic system on Γ\Gamma^{\prime}; see Section 2 on the kernel representation of this Dirichlet to Neumann map.

For the special case in (1.7), where γ\gamma only depends on the shear displacement u1+u_{1}^{+}, we can simplify and decouple the system (1.13) into two independent equations and finally drive a reduced scalar equation of u1+u_{1}^{+}, i.e. (1.15). In details, one can employ the Dirichlet to Neumann map and the elastic extension introduced in [19] to reduce the problem from 2×𝕋\mathbb{R}^{2}\times\mathbb{T} onto Γ\Gamma^{\prime}, i.e., to equations of (u1+,u3+)(u_{1}^{+},u_{3}^{+}) on the split plane Γ\Gamma^{\prime}. Second, if further employing (1.7), one can derive a linear representation formula between u1+u_{1}^{+} and u3+u_{3}^{+} on the Fourier side, i.e.,

(1.14) u^3+(𝒌)=νk1k2(1ν)k12+k22u^1+(𝒌).\displaystyle\hat{u}_{3}^{+}(\bm{k})=-\dfrac{\nu k_{1}k_{2}}{(1-\nu)k_{1}^{2}+k_{2}^{2}}\hat{u}_{1}^{+}(\bm{k}).

Here u^i+(k1,k2),i=1,3\hat{u}_{i}^{+}(k_{1},k_{2}),i=1,3 are the Fourier transform of ui+,i=1,3u_{i}^{+},i=1,3 with frequency vector 𝒌=(k1,k2),k1,k22π\bm{k}=(k_{1},k_{2}),k_{1}\in\mathbb{R},k_{2}\in 2\pi\mathbb{Z}. Substituting (1.14) into system (1.13), an independent equation of u1+u_{1}^{+} is derived, which contains a pseudo-differential operator \mathcal{L} defined on H1(×𝕋)H^{1}(\mathbb{R}\times\mathbb{T}):

(1.15) u1+(x,z)+γ(u1+(x,z))2G=0,u1^(𝒌)=|k|3u^1(𝒌)(1ν)k12+k22.\displaystyle\mathcal{L}u_{1}^{+}(x,z)+\dfrac{\gamma^{\prime}(u_{1}^{+}(x,z))}{2G}=0,\quad\widehat{\mathcal{L}u_{1}}(\bm{k})=\dfrac{|k|^{3}\hat{u}_{1}(\bm{k})}{(1-\nu)k_{1}^{2}+k_{2}^{2}}.

The derivation of (1.15) is standard and can be found in Section 2. Therefore, as long as we can solve the non-local semi-linear equation of u1+u_{1}^{+}, i.e. the first equation in (1.15), we can also find u3+u_{3}^{+} by (1.14), and then derive the solution of the original system (1.13). For brevity, we will omit the superscript ’++’ in the following sections. We call (1.15) the reduced scalar equation.

To solve (1.15), a meaningful observation is that we can write down an explicit solution to it for certain double-well potential γ\gamma’s. Highly compatible with dislocations in Halite, the cosine potential γ0=1π2(cos(πu)+1)\gamma_{0}=\dfrac{1}{\pi^{2}}(\cos(\pi u)+1) in the PN model implements an explicit solution [6, 17] to (1.15) and (1.14):

(1.16) u1(x,z)=2πarctan((1ν)x2G),u3(x,z)=0.\displaystyle u_{1}(x,z)=\dfrac{2}{\pi}\arctan\left(\dfrac{(1-\nu)x}{2G}\right),\quad u_{3}(x,z)=0.

In particular, u1u_{1} in (1.16) is a layer solution (see Definition 3) since it is strictly monotonic in xx direction and satisfies assumption (1.9). In fact, (1.16) is a good candidate for minimizers of total energy (1.1) in the sense of Definition 1. We will prove that this solution is the unique minimizer up to translations which concludes the question on existence and rigidity. We remark here that this is just a concrete example of our general result: for general double-well type potentials, we prove that minimizers of the total energy FF (see (1.37)) in function set (1.35) exist and they are layer solutions (see Definition 3). Moreover, they are unique up to translations.

1.2. Unsolved problems on the vectorial PN model

The existence and rigidity of the vectorial PN model are important in understanding dislocations. Previous literature on the vectorial PN model mainly focused on numerical simulations [40, 42, 28, 36] and physical experiments [42, 39], while only few rigorous mathematical results [19] were derived. In this section, we aim at mathematically formulating those important but unsolved problems into a framework and embedding our result into this macroscopic framework.

Consider the Euler-Lagrange equation of the vectorial model. We first observed that the second and fourth equations of (1.13) can be rewritten as

(1.17) 𝒜(u1+u3+)=12G(γu1+γu3+),𝒜^=(k22k+11νk12kν1νk1k2kν1νk1k2kk12k+11νk22k),\mathcal{A}\begin{pmatrix}u_{1}^{+}\\ u_{3}^{+}\end{pmatrix}=\dfrac{1}{2G}\begin{pmatrix}\dfrac{\partial\gamma}{\partial u_{1}^{+}}\\ \dfrac{\partial\gamma}{\partial u_{3}^{+}}\end{pmatrix},\quad\hat{\mathcal{A}}=\begin{pmatrix}\dfrac{k_{2}^{2}}{\|k\|}+\dfrac{1}{1-\nu}\cdot\dfrac{k_{1}^{2}}{\|k\|}&\dfrac{\nu}{1-\nu}\cdot\dfrac{k_{1}k_{2}}{\|k\|}\\ \dfrac{\nu}{1-\nu}\cdot\dfrac{k_{1}k_{2}}{\|k\|}&\dfrac{k_{1}^{2}}{\|k\|}+\dfrac{1}{1-\nu}\cdot\dfrac{k_{2}^{2}}{\|k\|}\end{pmatrix},

where 𝒜\mathcal{A} is a pseudo-differential operator with Fourier symbol 𝒜^\hat{\mathcal{A}}. Equation (1.17) is a nonlocal reduced elliptic equations on slip plane Γ\Gamma, which is an open system. Meanwhile, the misfit potential here may depend on both u1u_{1} and u3u_{3}: γ=γ(u1,u3)\gamma=\gamma(u_{1},u_{3}). As far as we know, neither existence nor rigidity of (1.17) was studied by previous literatures.

In fact, if the misfit potential is carefully selected [42, 28, 36, 6], solutions of (1.17) determine the underlying structure of dislocations in crystals such as Cu (ν=0.36\nu=0.36) and Al (ν=0.33\nu=0.33), no matter straight ones or curved ones. Considering symmetry of the crystal lattice, authors of [40] adopted a truncated Fourier expansion for the generalized stacking fault energy (see eq.(14) in [40]) as the misfit energy γ(u1,u3)\gamma(u_{1},u_{3}), with different coefficients for Cu and Al.

Numerical simulations in [40] indicated that straight edge dislocations in both Cu (figure 4 in [40]) and Al (figure 5 in [40]) possess the following structure: the displacement in the shear direction (i.e. u1u_{1}) is a layer solution (see Definition 2) and the displacement in the transverse direction (i.e. u3u_{3}) is a solitary wave. As claimed by authors in [40], this numerical result agrees with data from experiments on real materials [39].

To understand this consistency between the numerical result and the experimental data, we consider (1.17) where the Poisson ratio ν=0\nu=0, which is exactly the case for cork. This special case is less obscure since equations for u1u_{1} and u3u_{3} are decoupled now. Furthermore, we assume that the misfit potential consists of two parts which depend merely on u1u_{1} and u3u_{3} respectively:

(1.18) γ(u1,u3)=γ1(u1)+γ3(u3).\displaystyle\gamma(u_{1},u_{3})=\gamma_{1}(u_{1})+\gamma_{3}(u_{3}).

Suppose that γ1\gamma_{1} is a double-well potential (see (1.8)) and γ3\gamma_{3} is the nonlinear potential in the Benjamin-Ono equation [7], i.e.

(1.19) γ3(u3)=u322u333.\displaystyle\gamma_{3}(u_{3})=\dfrac{u_{3}^{2}}{2}-\dfrac{u_{3}^{3}}{3}.

If 𝒖\bm{u} is a minimizer of (1.1) with boundary conditions

(1.20) limx±u1(x,y,z)=±1,limx±u3(x,y,z)=0,\displaystyle\lim\limits_{x\to\pm\infty}u_{1}(x,y,z)=\pm 1,\ \lim\limits_{x\to\pm\infty}u_{3}(x,y,z)=0,

then separately, u1u_{1} and u3u_{3} satisfy

(1.21) 2G(Δ)1/2ui++γi(ui+)=0,i=1,3.\displaystyle 2G\cdot(-\Delta)^{1/2}u_{i}^{+}+\gamma_{i}^{\prime}(u_{i}^{+})=0,\ i=1,3.

For i=1i=1, because γ1\gamma_{1} is a double-well potential, [9] proved that (1.21) admits layer solutions u1u_{1} (unique up to translations) in the sense of Definition 2. For i=3i=3, (1.21) is the traveling wave form of the Benjamin-Ono equation which admits

(1.22) u3+(x,z)=4G4G2+x2\displaystyle u_{3}^{+}(x,z)=\dfrac{4G}{4G^{2}+x^{2}}

as a solitary solution [7]. These solutions also satisfy the boundary condition (1.20). These special solutions partially explain the structure of minimizers observed in [40, 39], i.e., it is a layer solution in the shear direction while it is a solitary wave in the transverse direction.

As far as we know, a complete answer on the rigidity of minimizers is still unknown even for the case ν=0\nu=0. More explicitly, does the De Giorgi conjecture hold in this case? Does (1.17) (or (1.21)) admit any other solution? Existing evidence indicates negative results. If (Δ)1/2(-\Delta)^{1/2} is replaced by Δ-\Delta in (1.21) and we take i=3i=3, the author of [11] proved that there exist solutions being a soliton in the xx direction while being periodic (but non-constant) in the zz direction. Thus, the one-dimensional symmetry (or the De Giorgi conjecture) fails in this case. This evidence strongly indicates the existence of high-dimensional solutions of u3u_{3} in (1.21). High dimensionality physically indicates the existence of curved dislocations but the construction of a counterexample for the De Giorgi conjecture in the nonlocal case is still open.

For general cases where the Poisson ratio ν\nu is non-zero, neither existence nor rigidity result is proved to our best knowledge. In particular, no matter u3u_{3} has a one-dimensional profile or not, no conclusion can be drawn on the rigidity of u1u_{1}.

Another question regarding the De Giorgi conjecture is also of great interest: for what misfit potential γ\gamma, there exist one-dimensional solutions for (1.17)? For what misfit potential γ\gamma, the De Giorgi conjecture holds, i.e. all solutions of (1.17) are one dimensional? No previous study has ever considered these problems as far as we know.

In summary, the rigidity and existence of the vectorial PN model is an important problem that is central to studies of dislocations, both straight and curved dislocations. Our contribution to this macroscopic framework is that, under the assumption γ=γ(u1)\gamma=\gamma(u_{1}), i.e. γ\gamma only depends on u1u_{1}, a complete answer to existence and rigidity is justified even for high dimensions. See the following sections in the introduction.

1.3. The nonlocal scalar equation in high dimensions

We remind our audience here that we will focus on the case where γ\gamma only depends on u1u_{1} and γ\gamma is a double-well type potential (see (1.8)) in the following sections.

We extend the discussion to any dimension d1d\geq 1 and clarify the set up. Denote Ωd:=×𝕋d1\Omega_{d}:=\mathbb{R}\times\mathbb{T}^{d-1}, consider the high-dimensional reduced scalar equation in Ωd\Omega_{d}:

(1.23) u(𝒘)+γ(u(𝒘))=0,𝒘Ωd.\displaystyle\mathcal{L}u(\bm{w})+\gamma^{\prime}(u(\bm{w}))=0,\ \bm{w}\in\Omega_{d}.

The potential function γC2()\gamma\in C^{2}(\mathbb{R}) is a double-well potential that satisfies (1.8). The linear operator \mathcal{L} is a convolution-type singular integral operator [38] which is defined as

(1.24) (u)(𝒘):=P.V.Ωd(u(𝒘)u(𝒘))K(𝒘𝒘)d𝒘\displaystyle(\mathcal{L}u)(\bm{w}):=\mathrm{P.V.}\int_{\Omega_{d}}(u(\bm{w})-u(\bm{w}^{\prime}))K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}^{\prime}

whose convolution kernel K(𝒘)K(\bm{w}) can be written as

(1.25) K(x,𝒚)=𝒋d1H(x,𝒚+𝒋)\displaystyle K(x,\bm{y})=\sum_{\bm{j}\in\mathbb{Z}^{d-1}}H(x,\bm{y}+\bm{j})

where 𝒘=(x,𝒚),x,𝒚𝕋d1\bm{w}=(x,\bm{y}),\ x\in\mathbb{R},\ \bm{y}\in\mathbb{T}^{d-1}.

We impose several assumptions on the operator \mathcal{L} and its kernel HH. To clarify these assumptions, we first introduce the Fourier transform on Ωd\Omega_{d} and the Sobolev spaces Hs(Ωd)H^{s}(\Omega_{d}).

Denote Ωd=×(2π)d1\Omega_{d}^{\prime}=\mathbb{R}\times(2\pi\mathbb{Z})^{d-1}. The Fourier transform on Ωd\Omega_{d} is understood as a composition in two directions: the Fourier transform on \mathbb{R} in the xx direction and the Fourier series expansion on 𝕋d1\mathbb{T}^{d-1} in the 𝒚\bm{y} direction. Denote 𝝂=(ξ,𝒌)\bm{\nu}=(\xi,\bm{k}) where ξ\xi\in\mathbb{R} and 𝒌(2π)d1\bm{k}\in(2\pi\mathbb{Z})^{d-1}, then the Fourier transform of u(𝒘)u(\bm{w}), denoted as u^(𝝂)\hat{u}(\bm{\nu}), is defined as

u^(𝝂)=Ωde2πi(xξ+𝒚𝒌)u(x,𝒚)dxd𝒚.\displaystyle\hat{u}(\bm{\nu})=\int_{\Omega_{d}}e^{-2\pi i(x\xi+\bm{y}\cdot\bm{k})}u(x,\bm{y})\mathrm{d}x\mathrm{d}\bm{y}.

Thus the Fourier transform on Ωd\Omega_{d} maps functions defined on Ωd\Omega_{d} into functions defined on Ωd\Omega_{d}^{\prime}. For any s>0s>0, we define Sobolev spaces Hs(Ωd)H^{s}(\Omega_{d}) in the classical way on the Fourier side:

(1.26) Hs(Ωd):={uL2(Ωd):|𝝂|su^(𝝂)L2(Ωd)}.\displaystyle H^{s}(\Omega_{d}):=\{u\in L^{2}(\Omega_{d}):\ |\bm{\nu}|^{s}\hat{u}(\bm{\nu})\in L^{2}(\Omega_{d}^{\prime})\}.

Hs(Ωd),s>0H^{s}(\Omega_{d}),\ s>0 are Hilbert spaces with inner product

(1.27) u,vHs(Ωd):=u^,v^L2(Ωd)+|𝝂|su^,|𝝂|sv^L2(Ωd).\displaystyle\langle u,v\rangle_{H^{s}(\Omega_{d})}:=\langle\hat{u},\hat{v}\rangle_{L^{2}(\Omega_{d}^{\prime})}+\langle|\bm{\nu}|^{s}\hat{u},|\bm{\nu}|^{s}\hat{v}\rangle_{L^{2}(\Omega_{d}^{\prime})}.

Denote the norm induced by this inner product as Hs(Ωd)\|\cdot\|_{H^{s}(\Omega_{d})}. We also define the homogeneous norm H˙s(Ωd)\|\cdot\|_{\dot{H}^{s}(\Omega_{d})} as

(1.28) uH˙s(Ωd):=|𝝂|su^L2(Ωd).\displaystyle\|u\|_{\dot{H}^{s}(\Omega_{d})}:=\||\bm{\nu}|^{s}\hat{u}\|_{L^{2}(\Omega_{d}^{\prime})}.

Now we are ready to impose assumptions of \mathcal{L} in (1.24) and introduce several important properties of it. We assume that:

  1. (A)

    (symbol of order 1) The Fourier symbol of \mathcal{L} is positive with same order as |𝝂||\bm{\nu}|, i.e. for any 𝝂Ωd=×(2π)d1\bm{\nu}\in\Omega_{d}^{\prime}=\mathbb{R}\times(2\pi\mathbb{Z})^{d-1}, there exist positive constants cc and CC such that

    (1.29) u^(𝝂)=σ(𝝂)u^(𝝂),c|𝝂|σ(𝝂)C|𝝂|\displaystyle\widehat{\mathcal{L}u}(\bm{\nu})=\sigma_{\mathcal{L}}(\bm{\nu})\hat{u}(\bm{\nu}),\ c|\bm{\nu}|\leq\sigma_{\mathcal{L}}(\bm{\nu})\leq C|\bm{\nu}|
  2. (B)

    (positivity and continuity) H(𝒛):dH(\bm{z}):\mathbb{R}^{d}\to\mathbb{R} is positive and continuous on d{𝟎}\mathbb{R}^{d}\setminus\{\bm{0}\}.

  3. (C)

    (homogeneity) For any 𝒛𝟎\bm{z}\neq\bm{0} and a>0a>0,

    (1.30) H(a𝒛)\displaystyle\ H(a\bm{z}) =ad1H(𝒛).\displaystyle=a^{-d-1}H(\bm{z}).
  4. (D)

    (symmetry) For any 𝒛d\bm{z}\in\mathbb{R}^{d}, H(𝒛)=H(𝒛)H(\bm{z})=H(-\bm{z}).

The assumptions we impose on \mathcal{L} and its kernel KK include two important cases. First, in dimension d=2d=2, the non-local operator in equation (1.15) derived from the PN model is included if the Poisson ratio ν(1/2,1/3)\nu\in(-1/2,1/3). In this case, the operator \mathcal{L} has Fourier symbol |𝒌|3/((1ν)k12+k22)|\bm{k}|^{3}/((1-\nu)k_{1}^{2}+k_{2}^{2}) (see (1.15)) which is of the same order as |k||k|, so Assumption (A) is satisfied. Moreover, the authors of [14] proved that KK satisfies assumption (B), (C) and (D) if and only if ν(1/2,1/3)\nu\in(-1/2,1/3). So equation (1.15) is included in this context if ν(1/2,1/3)\nu\in(-1/2,1/3). Second, in arbitrary dimensions, if we take ν=0\nu=0, then =(Δ)1/2\mathcal{L}=(-\Delta)^{1/2} defined on Ωd\Omega_{d} is also included. In this case, the Fourier symbol is exactly |k||k| and according to [26], there exists a constant Cd>0C_{d}>0 such that

(1.31) H(z)=Cd|𝒛|d+1.\displaystyle H(z)=\dfrac{C_{d}}{|\bm{z}|^{d+1}}.

So all assumptions are satisfied.

We remark here that we adopt two different but equivalent definitions for \mathcal{L}: one is as a Fourier multiplier and the other is as a singular convolution. The result that these two definitions for fractional Laplacian (Δ)α,α(0,1)(-\Delta)^{\alpha},\alpha\in(0,1) are equivalent is thoroughly investigated in [26]. For equation (1.15), the equivalence of these two definitions is also well-studied in [14]. So in later sections, we will switch between these two definitions for the sake of convenience.

1.4. Main results and strategies

Before presenting the main results, we introduce the fractional Allen-Cahn equation [5]:

(1.32) (Δ)1/2u(𝒙)+γ(u(𝒙))=0,𝒙d.\displaystyle(-\Delta)^{1/2}u(\bm{x})+\gamma^{\prime}(u(\bm{x}))=0,\ \bm{x}\in\mathbb{R}^{d}.

Here the double-well type potential γC2()\gamma\in C^{2}(\mathbb{R}) satisfying (1.8) is exactly the misfit potential in the PN model. Taking ν=0\nu=0 in (1.15), we see that (1.32) is a special case of (1.23) and (1.15). (1.32) has already been thoroughly investigated in the literatures [9, 32, 37, 34, 35, 15]. In particular, the well-posedness result of (1.32) is completely developed. A long standing conjecture named after De Giorgi [12, 21] (which originally discussed the local case, i.e., one replaces (Δ)1/2(-\Delta)^{1/2} by Δ-\Delta in (1.32), but then generalized to the non-local case (1.32)) is proved for dimension d8d\leq 8. The De Giorgi conjecture claims that any layer solution (see below) to (1.32) is a simple 1D profile for dimensions d8d\leq 8. In the classical Allen-Cahn equation, this conjecture is optimal in the sense that a counterexample in dimension d=9d=9 is constructed [13]. The layer solution in the De Giorgi conjecture is defined as:

Definition 2.

u:du:\mathbb{R}^{d}\to\mathbb{R} is a layer solution to equation (1.32) if

(1.33) u(𝒙)x1>0,limx1±u(𝒙)=±1.\displaystyle\dfrac{\partial u(\bm{x})}{\partial x_{1}}>0,\ \lim\limits_{x_{1}\to\pm\infty}u(\bm{x})=\pm 1.

The layer solution is also of main interest in the PN model since it models a dislocation profile that monotonically converges to two stable states at far field in the shear direction.

Now we are ready to articulate our existence and rigidity result on the PN model and (1.23). Although the PN model and (1.23) share the common double-well nonlinearity and non-localness with the fractional Landau-Ginzburg equation (see (1.32)), the main difference between our setting and previous work on (1.32) is that we work on a partially periodic domain Ωd\Omega_{d} and Hilbert spaces Hs(Ωd)H^{s}(\Omega_{d}) while previous work focused on the whole domain d\mathbb{R}^{d} and Banach spaces Ck,α(d)C^{k,\alpha}(\mathbb{R}^{d}). This discrepancy in the setting urges us to develop more appropriate methods while referring to some valuable techniques introduced in previous work.

For the existence problem, although we know that (1.16) is a solution to (1.15) and (1.14), we are still not aware of whether a minimizer of the total energy (1.1) in function set (1.35) exists. In [9], authors worked on Hölder spaces Ck,α(d),k=0,1,2,C^{k,\alpha}(\mathbb{R}^{d}),\ k=0,1,2, and derived some Schauder’s estimates of the weak solution to (1.32). Based on these estimates, they proved the existence of the classical solutions to (1.32) for d=1d=1 by considering the harmonic extension of (1.32) on the upper half-plane. In [32], the authors adopted the direct method in calculus of variations minimizing the total energy on a subset of Lloc1()L^{1}_{\mathrm{loc}}(\mathbb{R}), i.e.

(1.34) 𝒳:={fLloc1():limx±f(x)=±1}.\displaystyle\mathcal{X}:=\{f\in L^{1}_{\mathrm{loc}}(\mathbb{R}):\ \lim\limits_{x\to\pm\infty}f(x)=\pm 1\}.

They proved the existence and uniqueness of the minimizer in one dimension and the existence result is generalized to any dimension dd.

For the high-dimensional equation (1.23), we will follow the idea of [32] by using the direct method in the calculus of variations to prove that the minimizer of a functional F(u)F(u) exists. However, instead of requiring the far field assumption (1.9), we only consider H1/2H^{1/2} perturbation of a given 1D profile η\eta who satisfies (1.9):

(1.35) 𝒜:={uHloc1/2(Ωd):uηH1/2(Ωd)}.\displaystyle\mathcal{A}:=\{u\in H^{1/2}_{\mathrm{loc}}(\Omega_{d}):u-\eta\in H^{1/2}(\Omega_{d})\}.

Here η(x,𝒚)\eta(x,\bm{y}) is a smooth 1D profile, i.e. η(x,𝒚)=η(x)\eta(x,\bm{y})=\eta(x) for any (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}, satisfying

(1.36) η(x)\displaystyle\eta(x) C(),η(x)={1ifx[1,+),1ifx(,1].\displaystyle\in C^{\infty}(\mathbb{R}),\quad\eta(x)=\left\{\begin{array}[]{cc}1\quad\mathrm{if}\quad x\in[1,+\infty),\\ -1\quad\mathrm{if}\quad x\in(-\infty,-1].\end{array}\right.

We will abuse the notation η\eta to represent the profile defined on either Ωd\Omega_{d} or \mathbb{R}. We remark here that the weak H1/2H^{1/2} regularity does not ensure any far field limit behavior in any dimension, even in dimension d=1d=1.

The functional FF that we aim to minimize is the perturbed version of the total energy (1.1):

(1.37) F(u)\displaystyle F(u) :=12ΩdΩd|(u(x,𝒚)u(x,𝒚)|2K(xx,𝒚𝒚)\displaystyle:=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(u(x,\bm{y})-u(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})
|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚+Ωdγ(u(x,𝒚))dxd𝒚.\displaystyle-{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y^{\prime}}+\int_{\Omega_{d}}\gamma(u(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}.

Here γC2()\gamma\in C^{2}(\mathbb{R}) is the double-well type potential considered in (1.23) that satisfies (1.8).

Starting from functions only with H1/2H^{1/2} regularity and even not necessarily satisfying the far field limit condition, we construct a minimizer with H2H^{2} regularity (in fact smooth) that also satisfies the desired rigidity result that we aim to prove: it is a layer solution with 1D symmetry.

Theorem 1.

(Existence of the minimizer) Suppose that γC2()\gamma\in C^{2}(\mathbb{R}) is a double-well type potential satisfying condition (1.8). Consider set 𝒜\mathcal{A} defined in (1.35) and energy functional FF defined in (1.37). Then:

  1. (i)

    (existence) There exists u𝒜u^{*}\in\mathcal{A} such that F(u)=minu𝒜F(u)F(u^{*})=\min\limits_{u\in\mathcal{A}}F(u). In particular, uu^{*} is a weak solution to (1.23), i.e.

    u+γ(u)=0.\displaystyle\mathcal{L}u^{*}+\gamma^{\prime}(u^{*})=0.

    Here \mathcal{L} is defined as in (1.24).

  2. (ii)

    (regularity) uu^{*} in (i)(i) satisfies uηH2(Ωd)u^{*}-\eta\in H^{2}(\Omega_{d}). In particular, uu^{*} solves equation (1.23) in L2L^{2} sense and satisfies the far end limit condition uniformly in 𝒚\bm{y}:

    limx±u(x,𝒚)=±1\displaystyle\lim\limits_{x\to\pm\infty}u^{*}(x,\bm{y})=\pm 1
  3. (iii)

    (monotonicity) uu^{*} in (i)(i) satisfies u(x,𝒚)x>0\dfrac{\partial u^{*}(x,\bm{y})}{\partial x}>0 for any (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}, i.e. u(x,𝒚)u^{*}(x,\bm{y}) is strictly increasing in xx direction.

  4. (iv)

    (symmetry) uu^{*} in (i)(i) satisfies 𝒚u(x,𝒚)=𝟎\nabla_{\bm{y}}u^{*}(x,\bm{y})=\bm{0} for any (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}, i.e. u(x,𝒚)=u(x)u^{*}(x,\bm{y})=u^{*}(x) is a 1D profile.

The critical technique in the proof is the energy decreasing rearrangement method in [32]. This method relies on the rearrangement inequality (see Lemma 3.1) whose proof is quite elementary. However, driven by this basic inequality, the energy decreasing method is powerful in proving monotonicity and 1D symmetry. We will introduce this method in Section 3.1.

For the rigidity problem, it worth mentioning the De Giorgi conjecture on the fractional Ginzburg-Landau equation (1.32). It claims that at least for d8d\leq 8, layer solutions to (1.32) are in fact just 1D profiles. Here a layer solution is defined in Definition 3. This conjecture was proved for dimension d=2d=2 in [9] and finally completely proved by Savin in his series of work [34, 35, 32]. We also mention the asymptotic analysis for the sharp interface limit of the fractional diffusion-reaction equation with isotropic/anisotropic nonlocal kernel of order 1rd+2s,s(0,1)\frac{1}{r^{d+2s}},s\in(0,1) in [4, 20, 23, 30] .

The common approach to prove the De Giorgi conjecture is to develop a Liouville-type theorem and then apply the theorem on ratios of partial derivatives in different directions uxi/ux1,i=1,2,,d1u_{x_{i}}/u_{x_{1}},i=1,2,...,d-1. Then one can conclude that there exist constants ci,i=1,2,,d1c_{i},i=1,2,...,d-1 such that

uxi=ciux1,i=1,2,..,d1.\displaystyle u_{x_{i}}=c_{i}u_{x_{1}},\quad i=1,2,..,d-1.

So uu in fact only depends on x1x_{1} variable and hence is a 1D profile. For (1.23), we can prove the following theorem which is true for any dimension d1d\geq 1, not only for dimension d8d\leq 8:

Theorem 2.

(De Giorgi Conjecture) For any dimension d1d\geq 1, suppose that u:Ωdu:\Omega_{d}\to\mathbb{R} satisfies uηH1(Ωd)u-\eta\in H^{1}(\Omega_{d}) and is a layer solution (defined in Definition 3) to equation (1.23), i.e.

u+γ(u)=0.\displaystyle\mathcal{L}u+\gamma^{\prime}(u)=0.

Here γC()\gamma\in C^{\infty}(\mathbb{R}) is a double-well type potential satisfies (1.8) and \mathcal{L} is defined in (1.24) satisfying Assumptions (A)-(D). Then u(x,𝐲)u(x,\bm{y}) only depends on xx variable, i.e. there exists ϕ:\phi:\mathbb{R}\to\mathbb{R} such that u(x,𝐲)=ϕ(x)u(x,\bm{y})=\phi(x).

We emphasize here that the main reason of Theorem 2 being true in any dimension d1d\geq 1 instead of only dimensions less than eight is that we fully employed the compactness of the torus 𝕋d1\mathbb{T}^{d-1}. The compactness ensures convergence of a sequence which is a key step in our proof (see the proof of Theorem 2). As we explained in Section 1.2, without periodicity, the De Giorgi conjecture may fail in the classical case [11]. Therefore, domain Ωd=×𝕋d1\Omega^{d}=\mathbb{R}\times\mathbb{T}^{d-1} is critical to our result which is also physically meaningful since it incorporates the periodicity in materials.

Instead of using Liouville type theorems, we prove Theorem 2 by analyzing the spectrum of a linear operator. This method sufficiently respects the maximal property of operator \mathcal{L} (see Lemma 2.1) which is realized by the positivity assumption (Assumption (B)). Utilized in our previous work [18], this spectral analysis method is straightforward and appropriate in our setting since the working spaces are selected as Hilbert spaces Hs(Ωd)H^{s}(\Omega_{d}) instead of Banach spaces Ck,α(d)C^{k,\alpha}(\mathbb{R}^{d}) in [9]. Under this setting, we can use the perturbation theory of self-adjoint operators on Hilbert spaces [25].

Specifically speaking, suppose that uu is a given solution to (1.23) that satisfies conditions in Theorem 2. Differentiating on both sides of equation (1.23), we see that uxu_{x} and uyi,i=1,2,,d1u_{y_{i}},i=1,2,...,d-1 are solutions to the following non-local linear elliptic equation of ϕ\phi on Ωd\Omega_{d} which is given by:

[+γ′′(u)]ϕ=0.\displaystyle\left[\mathcal{L}+\gamma^{\prime\prime}(u)\right]\phi=0.

Equivalently, uxu_{x} and uyi,i=1,2,,d1u_{y_{i}},i=1,2,...,d-1 are eigenfunctions of eigenvalue 0 for the linear operator linearized along profile uu:

(1.38) L:H1(Ωd)L2(Ωd)L2(Ωd),Lϕ=ϕ+γ′′(u)ϕ.\displaystyle L:H^{1}(\Omega_{d})\subset L^{2}(\Omega_{d})\to L^{2}(\Omega_{d}),\ L\phi=\mathcal{L}\phi+\gamma^{\prime\prime}(u)\phi.

Therefore, as long as we can prove that 0 is a simple eigenvalue of LL, i.e. the eigenspace of 0 is only 1 dimension, then we prove that uxu_{x} and uyi,i=1,2,,d1u_{y_{i}},i=1,2,...,d-1 are linearly dependent, which indicates 1D symmetry. This is the main idea and approach we will utilize to prove Theorem 2.

As a direct corollary of Theorem 2, we can prove that both layer solutions and minimizers of FF on 𝒜\mathcal{A} are unique up to translations. Define 𝒜\mathcal{A}_{\ell} and 𝒜m\mathcal{A}_{m} as

(1.39) 𝒜\displaystyle\mathcal{A}_{\ell} :={uH˙1(Ωd):uηH1(Ωd),uisalayersolutionto(1.23)},\displaystyle:=\{u\in\dot{H}^{1}(\Omega_{d}):u-\eta\in H^{1}(\Omega_{d}),\ u\ \mathrm{is\ a\ layer\ solution\ to}\ \eqref{eq:generalized1d}\},
𝒜m\displaystyle\mathcal{A}_{m} :={u𝒜:F(u)=minv𝒜F(v)},\displaystyle:=\{u\in\mathcal{A}:F(u)=\min\limits_{v\in\mathcal{A}}F(v)\},

i.e. 𝒜\mathcal{A}_{\ell} is the set of layer solutions to (1.23) with H1H^{1} regularity and 𝒜m\mathcal{A}_{m} is the set of minimizers of FF on set 𝒜\mathcal{A}. Then we can prove the following theorem:

Theorem 3.

(uniqueness of minimizers and layer solutions) For any dimension d1d\geq 1, suppose that γC()\gamma\in C^{\infty}(\mathbb{R}) is a double-well type potential satisfying (1.8). Consider functional energy FF in (1.37), set 𝒜\mathcal{A} in (1.35), set 𝒜\mathcal{A}_{\ell} and set 𝒜m\mathcal{A}_{m} in (1.39). Then

(1.40) 𝒜m=𝒜={u:u(x,𝒚)=u(x+x0)forsomex0}.\displaystyle\mathcal{A}_{m}=\mathcal{A}_{\ell}=\{u:\ u(x,\bm{y})=u^{*}(x+x_{0})\ \mathrm{for\ some}\ x_{0}\in\mathbb{R}\}.

Here u(x)u^{*}(x) is the unique solution to equation

(1.41) c(xx)1/2u+γ(u)=0,u(0)=0.\displaystyle c_{\mathcal{L}}(-\partial_{xx})^{1/2}u^{*}+\gamma^{\prime}(u^{*})=0,\ u^{*}(0)=0.

Here cc_{\mathcal{L}} is the constant in Lemma 2.2.

Theorem 3 provides a compatible physical interpretation of the PN model in three dimensions (with periodicity in the transverse direction): if we assume exclusive dependence of the misfit potential on the shear displacement, then the equilibrium dislocation on the slip plane only admits shear displacements. Furthermore, this uniquely (up to translations) determined shear displacement is a strictly monotonic 1D profile connecting two stable states. This reduces the vectorial PN model to the two-dimensional PN model which was investigated in our previous work [19]. In summary, we have the following theorem:

Theorem 4.

Suppose that γC()\gamma\in C^{\infty}(\mathbb{R}) is a double-well type potential satisfying (1.8). Consider the functional energy EE in (1.1) integrating on 2×𝕋\mathbb{R}^{2}\times\mathbb{T}, i.e.

(1.42) E~(𝒖)=2×𝕋Γ12σ:εdxdydz+Γγ(u1+)dxdz.\displaystyle\tilde{E}(\bm{u})=\int_{\mathbb{R}^{2}\times\mathbb{T}\setminus\Gamma^{\prime}}\dfrac{1}{2}\sigma:\varepsilon\mathrm{d}x\mathrm{d}y\mathrm{d}z+\int_{\Gamma^{\prime}}\gamma(u_{1}^{+})\mathrm{d}x\mathrm{d}z.

Here 𝐮=(u1,u2,u3)\bm{u}=(u_{1},u_{2},u_{3}) is the displacement vector, σ\sigma and ε\varepsilon are the strain tensor and the stress tensor respectively given by (1.3) and Γ\Gamma^{\prime} is the slip plane defined in (1.12). Assume ν(1/2,1/3)\nu\in(-1/2,1/3). Suppose that 𝐮\bm{u} is a global minimizer of E~\tilde{E} as in Definition 1, then:

  1. (i)

    (regularity) The displacement vector 𝒖\bm{u} is smooth in 2×𝕋Γ\mathbb{R}^{2}\times\mathbb{T}\setminus{\Gamma^{\prime}}.

  2. (ii)

    (rigidity) The displacement in transverse direction is 0, i.e. u3=0u_{3}=0 in 2×𝕋\mathbb{R}^{2}\times\mathbb{T}; u1+u_{1}^{+} is the unique (up to translation in xx direction) 1D profile independent with zz variable, strictly monotonic in xx direction satisfying

    (1.43) limx±u1+(x)=±1.\displaystyle\lim\limits_{x\to\pm\infty}u^{+}_{1}(x)=\pm 1.
  3. (iii)

    (Fourier representation) u1u_{1} and u2u_{2} only depend on xx and yy in 2×𝕋\mathbb{R}^{2}\times\mathbb{T}. On the Fourier side, u1u_{1} and u2u_{2} can be uniquely represented by u1±(x)u_{1}^{\pm}(x):

    (1.44) u^1±(ξ,y)\displaystyle\hat{u}_{1}^{\pm}(\xi,y) =u^1±(ξ)(1|ξy|22ν)e|ξy|\displaystyle=\hat{u}_{1}^{\pm}(\xi)\left(1-\dfrac{|\xi y|}{2-2\nu}\right)e^{-|\xi y|}
    (1.45) u^2±(ξ,y)\displaystyle\hat{u}_{2}^{\pm}(\xi,y) =u^1±(ξ)22ν((12ν)iξ|ξ|+iξ|y|)e|ξy|.\displaystyle=-\dfrac{\hat{u}_{1}^{\pm}(\xi)}{2-2\nu}\left((1-2\nu)\dfrac{i\xi}{|\xi|}+i\xi|y|\right)e^{-|\xi y|}.
  4. (iv)

    (Dirichlet to Neumann map) On Γ\Gamma^{\prime}, the stress tensor can be expressed as

    (1.46) σ12+(x)\displaystyle\sigma^{+}_{12}(x) =σ12(x)=G(1ν)πP.V.(u1+)(s)xsds\displaystyle=\sigma^{-}_{12}(x)=-\dfrac{G}{(1-\nu)\pi}\mathrm{P.V.}\int_{\mathbb{R}}\dfrac{(u^{+}_{1})^{\prime}(s)}{x-s}\mathrm{d}s
    (1.47) σ22+(x)\displaystyle\sigma^{+}_{22}(x) =σ22=0.\displaystyle=\sigma^{-}_{22}=0.
  5. (v)

    The stress tensor is divergence free, i.e.

    (1.48) σ=𝟎,holdsinD(2×𝕋).\displaystyle\nabla\cdot\sigma=\bm{0},\ \mathrm{holds}\ \mathrm{in}\ D^{\prime}(\mathbb{R}^{2}\times\mathbb{T}).

    This also holds point-wisely in 2×𝕋Γ\mathbb{R}^{2}\times\mathbb{T}\setminus\Gamma^{\prime}.

These four theorems (Theorem 1, 2, 3 and 4) are the main results for this work which completely close the problem of existence and rigidity in a general setting including the original PN model with Poisson ratio ν(1/2,1/3)\nu\in(-1/2,1/3). Following this logic, we will first conduct a preliminary analysis in Section 2 to assist readers to bridge some gaps in understanding the derivation of (1.23) and be aware of some important properties of the linear operator \mathcal{L}. Then we prove Theorem 1 in Section 3 and Theorem 2, 3 and 4 in Section 4. Finally, the spectral analysis of operator LL is established in Section 5 which proves that 0 is simple and the principle eigenvalue of LL. For facts in functional analysis and details in the spectral analysis, readers may refer to Appendix A; for proofs of some lemmas in the proof of the theorems, readers may refer to Appendix B.

2. Preliminary analysis

In this section, we will first provide some details of the derivation of the reduced scalar equation (1.15), then discuss three important properties that will be used in the proof of the three theorems.

2.1. Derivation of the reduced scalar equation

Denote the Fourier transform of ui+(x,z),i=1,2,3u_{i}^{+}(x,z),i=1,2,3 as u^i+(𝒌),i=1,2,3\hat{u}_{i}^{+}(\bm{k}),i=1,2,3 where 𝒌=(k1,k2),k1,k22π\bm{k}=(k_{1},k_{2}),k_{1}\in\mathbb{R},k_{2}\in 2\pi\mathbb{Z} is the frequency vector. Given 𝒖\bm{u} that satisfies equation (1.13), one can rewrite (σ12+,σ32+)(\sigma_{12}^{+},\sigma_{32}^{+}) on Γ\Gamma^{\prime} as a linear transform of (u1+(𝒌),u3+(𝒌))(u_{1}^{+}(\bm{k}),u_{3}^{+}(\bm{k})) on the Fourier side:

(2.1) (σ^12+(𝒌)σ^32+(𝒌))=𝑨(u^1+(𝒌)u^3+(𝒌)):=2G((k22|𝒌|+11νk12|𝒌|)u^1+(𝒌)+ν1νk1k2|𝒌|u^3+(𝒌)ν1νk1k2|𝒌|u^1+(𝒌)+(k12|𝒌|+11νk22|𝒌|)u^3+(𝒌)).\displaystyle\begin{pmatrix}\ \hat{\sigma}_{12}^{+}(\bm{k})\ \\ \ \hat{\sigma}_{32}^{+}(\bm{k})\ \\ \end{pmatrix}=-\bm{A}\begin{pmatrix}\ \hat{u}_{1}^{+}(\bm{k})\ \\ \ \hat{u}_{3}^{+}(\bm{k})\ \\ \end{pmatrix}:=-2G\begin{pmatrix}\ \left(\dfrac{k_{2}^{2}}{|\bm{k}|}+\dfrac{1}{1-\nu}\dfrac{k_{1}^{2}}{|\bm{k}|}\right)\hat{u}_{1}^{+}(\bm{k})+\dfrac{\nu}{1-\nu}\dfrac{k_{1}k_{2}}{{|\bm{k}|}}\hat{u}_{3}^{+}(\bm{k})\ \\ \ \dfrac{\nu}{1-\nu}\dfrac{k_{1}k_{2}}{|\bm{k}|}\hat{u}_{1}^{+}(\bm{k})+\left(\dfrac{k_{1}^{2}}{|\bm{k}|}+\dfrac{1}{1-\nu}\dfrac{k_{2}^{2}}{|\bm{k}|}\right)\hat{u}_{3}^{+}(\bm{k})\ \\ \end{pmatrix}.

Details of this derivation can be found in Appendix [14].

From equation (2.1), the Euler-Lagrangian equation (1.13) can be rewritten as an equation of u1+(x,z),u3+(x,z)u_{1}^{+}(x,z),u_{3}^{+}(x,z) on Γ\Gamma, i.e.

(2.2) 𝒜(u1+(x,z)u3+(x,z))=(γu1(u1+,u3+)γu3(u1+,u3+)).\displaystyle-\mathcal{A}\begin{pmatrix}u_{1}^{+}(x,z)\\ u_{3}^{+}(x,z)\\ \end{pmatrix}=\begin{pmatrix}\dfrac{\partial\gamma}{\partial u_{1}}(u_{1}^{+},u_{3}^{+})\\ \dfrac{\partial\gamma}{\partial u_{3}}(u_{1}^{+},u_{3}^{+})\\ \end{pmatrix}.

Here 𝒜\mathcal{A} is the nonlocal differential operator with Fourier symbol 𝑨\bm{A}.

A further simplification can be realized on equation (2.2) due to independence of γ\gamma with u3u_{3}, i.e. γu3=0\dfrac{\partial\gamma}{\partial u_{3}}=0. This independence reduces equation (2.2) into an equation of u1u_{1}. On the Fourier side, the second component in (2.2) indicates that we can represent u^3\hat{u}_{3} by u^1\hat{u}_{1}, i.e.

(2.3) u^3(k)=νk1k2(1ν)k12+k22u^1(k).\displaystyle\hat{u}_{3}(k)=-\dfrac{\nu k_{1}k_{2}}{(1-\nu)k_{1}^{2}+k_{2}^{2}}\hat{u}_{1}(k).

Substituting this equality to the first component in (2.1) yields

σ^12(𝒌)\displaystyle\hat{\sigma}_{12}(\bm{k}) =2G[(k22|𝒌|+11νk12|𝒌|)u^1(𝒌)+ν1νk1k2|𝒌|u^3(𝒌)]\displaystyle=-2G\left[\left(\dfrac{k_{2}^{2}}{|\bm{k}|}+\dfrac{1}{1-\nu}\dfrac{k_{1}^{2}}{|\bm{k}|}\right)\hat{u}_{1}(\bm{k})+\dfrac{\nu}{1-\nu}\dfrac{k_{1}k_{2}}{|\bm{k}|}\hat{u}_{3}(\bm{k})\right]
=2G|𝒌|3(1ν)k12+k22u^1(𝒌).\displaystyle=-\dfrac{2G|\bm{k}|^{3}}{(1-\nu)k_{1}^{2}+k_{2}^{2}}\hat{u}_{1}(\bm{k}).

Now denote :H1(×𝕋)L2(×𝕋)L2(×𝕋)\mathcal{L}:H^{1}(\mathbb{R}\times\mathbb{T})\subset L^{2}(\mathbb{R}\times\mathbb{T})\to L^{2}(\mathbb{R}\times\mathbb{T}) the linear operator with Fourier symbol |𝒌|3(1ν)k12+k22\dfrac{|\bm{k}|^{3}}{(1-\nu)k_{1}^{2}+k_{2}^{2}}. Then the first component of equation (2.2)\eqref{eq:gammasys} is in fact an equation of u1u_{1}, i.e. equation (1.15):

(2.4) u1+γ(u1)2G=0.\displaystyle\mathcal{L}u_{1}+\dfrac{\gamma^{\prime}(u_{1})}{2G}=0.

This equation is the reduced scalar equation.

2.2. Properties of \mathcal{L}

Assumption (A) ensures that \mathcal{L} is a self-adjoint operator defined on H1(Ωd)L2(Ωd)H^{1}(\Omega_{d})\subset L^{2}(\Omega_{d}) and maps to L2(Ωd)L^{2}(\Omega_{d}) (see Lemma A.2). By assumption (B) and (C), one can easily conclude that kernel HH satisfies that for any 𝒛𝟎\bm{z}\neq\bm{0},

(2.5) 0<m|𝒛|d+1H(𝒛)\displaystyle 0<\dfrac{m}{|\bm{z}|^{d+1}}\leq{H}(\bm{z}) M|𝒛|d+1.\displaystyle\leq\dfrac{M}{|\bm{z}|^{d+1}}.

Here mm and MM are positive constants. Indeed, for any non-zero 𝒛\bm{z}, we have

H(𝒛)=1|𝒛|n+1H(𝒛|𝒛|)\displaystyle{H}(\bm{z})=\dfrac{1}{|\bm{z}|^{n+1}}{H}\left(\dfrac{\bm{z}}{|\bm{z}|}\right)

and H(𝒛){H}(\bm{z}) has a positive lower bound mm and a positive upper bound MM on the compact set 𝕊d1\mathbb{S}^{d-1}. So (2.5) holds.

Furthermore, we will prove three important properties of the linear operator \mathcal{L} which play critical roles in the proof of Theorem 1, 2 and 3.

First, positivity of HH ensures that if ff attains global maximum at point (x0,𝒚0)Ωd(x_{0},\bm{y}_{0})\in\Omega_{d}, then f|(x0,𝒚0)0\mathcal{L}f|_{(x_{0},\bm{y}_{0})}\geq 0. We call it the maximal principle of operator \mathcal{L}:

Lemma 2.1.

(Maximal principle) Suppose that fH˙1(Ωd)f\in\dot{H}^{1}(\Omega_{d}) attains global maximum at (xM,𝐲M)(x_{M},\bm{y}_{M}) and global minimum at (xm,𝐲m)(x_{m},\bm{y}_{m}) on Ωd\Omega_{d}. Then

f|(xM,𝒚M)0,f|(xm,𝒚m)0.\displaystyle\mathcal{L}f|_{(x_{M},\bm{y}_{M})}\geq 0,\quad\mathcal{L}f|_{(x_{m},\bm{y}_{m})}\leq 0.

The equality holds if and only if ff is constant.

Proof.

By positivity of KK, we know that

f|(xM,𝒚M)\displaystyle\mathcal{L}f|_{(x_{M},\bm{y}_{M})} =Ωd(f(xM,𝒚M)f(x,𝒚))K(xMx,𝒚M𝒚)dxd𝒚0,\displaystyle=\int_{\Omega_{d}}(f(x_{M},\bm{y}_{M})-f(x,\bm{y}))K(x_{M}-x,\bm{y}_{M}-\bm{y})\mathrm{d}x\mathrm{d}\bm{y}\geq 0,
f|(xm,𝒚m)\displaystyle\mathcal{L}f|_{(x_{m},\bm{y}_{m})} =Ωd(f(xm,𝒚m)f(x,𝒚))K(xmx,𝒚m𝒚)dxd𝒚0.\displaystyle=\int_{\Omega_{d}}(f(x_{m},\bm{y}_{m})-f(x,\bm{y}))K(x_{m}-x,\bm{y}_{m}-\bm{y})\mathrm{d}x\mathrm{d}\bm{y}\leq 0.

Thus the inequality holds and the equality holds if and only if f(x,𝒚)f(x,\bm{y}) is constant. ∎

We emphasize that this property of operator \mathcal{L} plays an important role in the proof of the De Giorgi conjecture (see Section 4.2).

Second, homogeneity of HH ensures that if f(x,𝒚)=f(x)f(x,\bm{y})=f(x), i.e. ff is a simple 1D profile independent with variable 𝒚\bm{y}, then there exists constant cc_{\mathcal{L}} such that f|(x,𝒚)=c((xx)1/2f)(x)\mathcal{L}f|_{(x,\bm{y})}=c_{\mathcal{L}}((-\partial_{xx})^{1/2}f)(x).

Lemma 2.2.

Suppose that fH˙1(Ωd)f\in\dot{H}^{1}(\Omega_{d}) satisfies f(x,𝐲)=f(x)f(x,\bm{y})=f(x). Then there exists a constant c>0c_{\mathcal{L}}>0 such that

f|(x,𝒚)=c((xx)1/2f)(x).\displaystyle\mathcal{L}f|(x,\bm{y})=c_{\mathcal{L}}((-\partial_{xx})^{1/2}f)(x).
Proof.

Consider g(x)g(x) which is defined as

g(x)=d1H(x,𝒚)d𝒚.\displaystyle g(x)=\int_{\mathbb{R}^{d-1}}{H}(x,\bm{y})\mathrm{d}\bm{y}.

Then for any x0x\neq 0, a change of variable implies that

g(x)=d1H(x,𝒚)d𝒚=d1|x|d1H(1,𝒚/x)d𝒚=d1|x|2H(1,𝒚)d𝒚=g(1)|x|2.\displaystyle g(x)=\int_{\mathbb{R}^{d-1}}{H}(x,\bm{y})\mathrm{d}\bm{y}=\int_{\mathbb{R}^{d-1}}|x|^{-d-1}{H}(1,\bm{y}/x)\mathrm{d}\bm{y}=\int_{\mathbb{R}^{d-1}}|x|^{-2}{H}(1,\bm{y})\mathrm{d}\bm{y}=\dfrac{g(1)}{|x|^{2}}.

So g(x)=g(1)|x|2g(x)=g(1)|x|^{-2} is the kernel of half Laplacian for one dimension. Therefore by Fubini’s theorem, if f(x,𝒚)=f(x)f(x,\bm{y})=f(x), we have

f|(x,𝒚)\displaystyle\mathcal{L}f|_{(x,\bm{y})} =P.V.𝕋d1(f(x,𝒚)f(x,𝒚))K(xx,𝒚𝒚)d𝒚dx\displaystyle=\mathrm{P.V.}\int_{\mathbb{R}}\int_{\mathbb{T}^{d-1}}(f(x,\bm{y})-f(x^{\prime},\bm{y}^{\prime}))K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}\bm{y}^{\prime}\mathrm{d}x^{\prime}
=P.V.(f(x)f(x))𝕋d1K(xx,𝒚𝒚)d𝒚dx.\displaystyle=\mathrm{P.V.}\int_{\mathbb{R}}(f(x)-f(x^{\prime}))\int_{\mathbb{T}^{d-1}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}\bm{y}^{\prime}\mathrm{d}x^{\prime}.

By (1.25), we have

𝕋d1K(xx,𝒚𝒚)d𝒚\displaystyle\int_{\mathbb{T}^{d-1}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}\bm{y}^{\prime} =𝕋d1𝒋d1H(xx,𝒚𝒚+𝒋)d𝒚\displaystyle=\int_{\mathbb{T}^{d-1}}\sum_{\bm{j}\in\mathbb{Z}^{d-1}}{H}(x-x^{\prime},\bm{y}-\bm{y}^{\prime}+\bm{j})\mathrm{d}\bm{y}^{\prime}
=d1H(xx,𝒚𝒚)d𝒚.\displaystyle=\int_{\mathbb{R}^{d-1}}{H}(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}\bm{y}^{\prime}.

Substituting this back to the formula of f\mathcal{L}f, we have

f|(x,𝒚)\displaystyle\mathcal{L}f|_{(x,\bm{y})} =P.V.(f(x)f(x))d1H(xx,𝒚𝒚)d𝒚dx\displaystyle=\mathrm{P.V.}\int_{\mathbb{R}}(f(x)-f(x^{\prime}))\int_{\mathbb{R}^{d-1}}{H}(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}\bm{y}^{\prime}\mathrm{d}x^{\prime}
=g(1)P.V.f(x)f(x)|xx|2dx\displaystyle=g(1)\cdot\mathrm{P.V.}\int_{\mathbb{R}}\dfrac{f(x)-f(x^{\prime})}{|x-x^{\prime}|^{2}}\mathrm{d}x^{\prime}
=g(1)((xx)1/2f)(x).\displaystyle=g(1)\cdot((-\partial_{xx})^{1/2}f)(x).

So for any ff that is 1D profile, \mathcal{L} acting on ff is just (xx)1/2(-\partial_{xx})^{1/2} acting on ff up to a constant. ∎

Third, Assumption (A) (equation (1.29)) ensures the following equivalence of semi-norms:

Lemma 2.3.

There exist positive constants c1,c2,C1,C2c_{1},c_{2},C_{1},C_{2} such that

(2.6) c1uH˙1(Ωd)uL2(Ωd)C1uH˙1(Ωd),c2uH˙1/2(Ωd)2ΩdΩd|u(𝒘)u(𝒘)|2K(𝒘𝒘)d𝒘d𝒘C2uH˙1/2(Ωd)2.\begin{gathered}c_{1}\|u\|_{\dot{H}^{1}(\Omega_{d})}\leq\|\mathcal{L}u\|_{L^{2}(\Omega_{d})}\leq C_{1}\|u\|_{\dot{H}^{1}(\Omega_{d})},\\ c_{2}\|u\|_{\dot{H}^{1/2}(\Omega_{d})}^{2}\leq\int_{\Omega_{d}}\int_{\Omega_{d}}|u(\bm{w})-u(\bm{w^{\prime}})|^{2}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w^{\prime}}\leq C_{2}\|u\|_{\dot{H}^{1/2}(\Omega_{d})}^{2}.\end{gathered}
Proof.

By Plancherel’s theorem, we have

uL2(Ωd)=u^(𝝂)L2(Ωd)=σ(𝝂)u^(𝝂)L2(Ωd).\displaystyle\|\mathcal{L}u\|_{L^{2}(\Omega_{d})}=\|\widehat{\mathcal{L}u}(\bm{\nu})\|_{L^{2}(\Omega_{d}^{\prime})}=\|\sigma_{\mathcal{L}}(\bm{\nu})\hat{u}(\bm{\nu})\|_{L^{2}(\Omega_{d}^{\prime})}.

Then by (1.29), we know

u^(𝝂)L2(Ωd)C|𝝂|u^(ν)L2(Ωd)=CuH˙1(Ω),\displaystyle\|\widehat{\mathcal{L}u}(\bm{\nu})\|_{L^{2}(\Omega_{d}^{\prime})}\leq C\||\bm{\nu}|\hat{u}(\nu)\|_{L^{2}(\Omega_{d}^{\prime})}=C\|u\|_{\dot{H}^{1}(\Omega)},
u^(𝝂)L2(Ωd)c|𝝂|u^(ν)L2(Ωd)=cuH˙1(Ω).\displaystyle\|\widehat{\mathcal{L}u}(\bm{\nu})\|_{L^{2}(\Omega_{d}^{\prime})}\geq c\||\bm{\nu}|\hat{u}(\nu)\|_{L^{2}(\Omega_{d}^{\prime})}=c\|u\|_{\dot{H}^{1}(\Omega)}.

Here CC and cc are constants in (1.29). So uL2(Ωd)\|\mathcal{L}u\|_{L^{2}(\Omega_{d})} is equivalent to uH˙1(Ωd)\|u\|_{\dot{H}^{1}(\Omega_{d})}. Moreover, by symmetry assumption of KK, we have

u,uL2(Ωd)\displaystyle\langle u,\mathcal{L}u\rangle_{L^{2}(\Omega_{d})} =ΩdΩdu(𝒘)(u(𝒘)u(𝒘))K(𝒘𝒘)d𝒘d𝒘\displaystyle=\int_{\Omega_{d}}\int_{\Omega_{d}}u(\bm{w})(u(\bm{w})-u(\bm{w^{\prime}}))K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w^{\prime}}
=12ΩdΩd|u(𝒘)u(𝒘)|2K(𝒘𝒘)d𝒘d𝒘.\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}|u(\bm{w})-u(\bm{w^{\prime}})|^{2}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w^{\prime}}.

By properties of the Fourier transform, we have

u,uL2(Ωd)=u^,u^L2(Ωd)=u^,σ(𝝂)u^L2(Ωd)\displaystyle\langle u,\mathcal{L}u\rangle_{L^{2}(\Omega_{d})}=\langle\hat{u},\widehat{\mathcal{L}u}\rangle_{L^{2}(\Omega_{d}^{\prime})}=\langle\hat{u},\sigma_{\mathcal{L}}({\bm{\nu}})\hat{u}\rangle_{L^{2}(\Omega_{d}^{\prime})}

hence by (1.29),

u,uL2(Ωd)C|𝝂|1/2u^(𝝂)L2(Ωd)2=CuH˙1/2(Ω)2,\displaystyle\langle u,\mathcal{L}u\rangle_{L^{2}(\Omega_{d})}\leq C\||\bm{\nu}|^{1/2}\hat{u}(\bm{\nu})\|^{2}_{L^{2}(\Omega_{d}^{\prime})}=C\|u\|_{\dot{H}^{1/2}(\Omega)}^{2},
u,uL2(Ωd)c|𝝂|1/2u^(𝝂)L2(Ωd)2=cuH˙1/2(Ω)2.\displaystyle\langle u,\mathcal{L}u\rangle_{L^{2}(\Omega_{d})}\geq c\||\bm{\nu}|^{1/2}\hat{u}(\bm{\nu})\|^{2}_{L^{2}(\Omega_{d}^{\prime})}=c\|u\|_{\dot{H}^{1/2}(\Omega)}^{2}.

Thus (2.6) holds. ∎

3. Existence of minimizers

In this section, we will prove Theorem 1. Recall the energy functional defined in (1.37), i.e.

F(u)\displaystyle F(u) =12ΩdΩd|(u(x,𝒚)u(x,𝒚)|2K(xx,𝒚𝒚)\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(u(x,\bm{y})-u(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})
|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚+Ωdγ(u(x,𝒚))dxd𝒚.\displaystyle-{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y^{\prime}}+\int_{\Omega_{d}}\gamma(u(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}.

We first rewrite this energy functional. In fact, by Lemma 2.2, if we denote v=uηv=u-\eta, then we can rewrite FF as

F(u)\displaystyle F(u) =12ΩdΩd|(η(x,𝒚)+v(x,𝒚)η(x,𝒚)v(x,𝒚))|2K(xx,𝒚𝒚)\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(\eta(x,\bm{y})+v(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})-v(x^{\prime},\bm{y}^{\prime}))|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})
|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚+Ωdγ(u(w))d𝒘\displaystyle-{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y^{\prime}}+\int_{\Omega_{d}}\gamma(u(w))\mathrm{d}\bm{w}
=12ΩdΩd|v(x,𝒚)v(x,𝒚))|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|v(x,\bm{y})-v(x^{\prime},\bm{y}^{\prime}))|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y^{\prime}}
(3.1) +2cΩdv(x,𝒚)(xx)1/2η(x)dxd𝒚+Ωdγ(u(w))d𝒘.\displaystyle+{2}c_{\mathcal{L}}\int_{\Omega_{d}}v(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}+\int_{\Omega_{d}}\gamma(u(w))\mathrm{d}\bm{w}.

Here cc_{\mathcal{L}} is the constant in Lemma 2.2. From (3) we see that subtraction of η\eta in the definition (1.37) ensures that FF is finite if uu is bounded and satisfies uηH1/2(Ω)u-\eta\in H^{1/2}(\Omega). For the sake of convenience, we will switch between (1.37) and (3) when using functional FF.

The main idea in the proof of Theorem 1 is to first slightly modify the minimizing problem on a subset of 𝒜\mathcal{A} denoted as 𝒜I\mathcal{A}_{I}. 𝒜I\mathcal{A}_{I} is defined as

(3.2) 𝒜I:={u𝒜:u=ηonΩdΩdI}.\displaystyle\mathcal{A}_{I}:=\{u\in\mathcal{A}:u=\eta\ \mathrm{on}\ \Omega_{d}\setminus\Omega_{d}^{I}\}.

Here ΩdI=I×𝕋d1,I=(a,b)\Omega_{d}^{I}=I\times\mathbb{T}^{d-1},\ I=(a,b) where a<1a<-1 and b>1b>1 are real numbers. By definition of minimizers, a minimizer uI𝒜Iu_{I}\in\mathcal{A}_{I} solves the following Dirichlet problem in weak sense:

(3.3) {u(x,𝒚)+γ(u(x,𝒚))=0,(x,𝒚)ΩdIu(x,𝒚)=1,x[b,+),u(x,𝒚)=1,x(,a].\displaystyle\begin{cases}\mathcal{L}u(x,\bm{y})+\gamma^{\prime}(u(x,\bm{y}))=0,\ (x,\bm{y})\in\Omega_{d}^{I}\\ u(x,\bm{y})=1,\ x\in[b,+\infty),\\ u(x,\bm{y})=-1,\ x\in(-\infty,a].\\ \end{cases}

For a minimizer uIu_{I}, result similar to Theorem 1 can be proved, which is summarized in the following proposition:

Proposition 1.

Suppose that γC2()\gamma\in C^{2}(\mathbb{R}) is a double-well type potential satisfying condition (1.8). Define function set 𝒜\mathcal{A} as in (1.35) and energy functional FF as in (1.37). Then:

  1. (i)

    (existence) There exists uI𝒜Iu_{I}\in\mathcal{A}_{I} such that F(uI)=minu𝒜IF(u)F(u_{I})=\min\limits_{u\in\mathcal{A}_{I}}F(u). In particular, uIu_{I} is a weak solution to (3.3).

  2. (ii)

    (monotonicity) uIu_{I} in (i)(i) satisfies that for any τ1>0\tau_{1}>0, uI(x+τ1,𝒚)uI(x,𝒚)u_{I}(x+\tau_{1},\bm{y})\geq u_{I}(x,\bm{y}) holds for a.e. (x,𝒚)ΩdI(x,\bm{y})\in\Omega_{d}^{I}, i.e. uI(x,y)u_{I}(x,y) is increasing in xx direction.

  3. (iii)

    (symmetry) uIu_{I} in (i)(i) satisfies that for any 𝝉2d1\bm{\tau}_{2}\in\mathbb{R}^{d-1}, uI(x,𝒚+𝝉2)=uI(x,𝒚)u_{I}(x,\bm{y}+\bm{\tau}_{2})=u_{I}(x,\bm{y}) holds for a.e. (x,𝒚)ΩdI(x,\bm{y})\in\Omega_{d}^{I}, i.e. uI(x,𝒚)=uI(x)u_{I}(x,\bm{y})=u_{I}(x) is a 1D profile.

To prove monotonicity and symmetry, one needs to utilize a critical technique: the energy decreasing rearrangement method in [32] which is based on the rearrangement inequality. After constructing {uI}\{u_{I}\} in Proposition 1, a minimizer of F(u)F(u) on 𝒜\mathcal{A} will be constructed using these minimizers on finite intervals and Theorem 1 is ready to be proved.

Following this logic, we will first carefully introduce the energy decreasing rearrangement method utilized in [32] in Section 3.1. This tool is prepared for the proof of Proposition 1 in Section 3.3 which ensures existence of the minimizer on 𝒜I\mathcal{A}_{I}. Then in Section 3.4, we will introduce several technical lemmas before proving Theorem 1 in Section 3.5.

3.1. Energy decreasing rearrangement

We denote a+=max{a,0}a_{+}=\max\{a,0\} and a=min{a,0}a_{-}=-\min\{a,0\}, i.e., a+a_{+} and aa_{-} represent the positive part and the negative part of aa respectively. In this section, we will introduce the energy decreasing rearrangement method that is used in [32]. In fact, this method relies on the following elementary equality:

Lemma 3.1.

(rearrangement) Suppose that a1,a2,b1,b2a_{1},a_{2},b_{1},b_{2} are four real numbers. Denote a=min{a1,a2},A=max{a1,a2},b=min{b1,b2}a=\min\{a_{1},a_{2}\},A=\max\{a_{1},a_{2}\},b=\min\{b_{1},b_{2}\} and B=max{b1,b2}B=\max\{b_{1},b_{2}\}. Then the following inequality holds:

(3.4) ab+ABa1b1a2b2=(a1a2)+(b1b2)+(a1a2)(b1b2)+0.\displaystyle ab+AB-a_{1}b_{1}-a_{2}b_{2}=(a_{1}-a_{2})_{+}(b_{1}-b_{2})_{-}+(a_{1}-a_{2})_{-}(b_{1}-b_{2})_{+}\geq 0.

In particular, ab+ABa1b1a2b2=0ab+AB-a_{1}b_{1}-a_{2}b_{2}=0 if and only if (a1a2)(b1b2)0(a_{1}-a_{2})(b_{1}-b_{2})\geq 0.

Readers may refer to Appendix B for proof of this inequality. Now we are ready to introduce the energy decreasing rearrangement method.

Lemma 3.2.

(energy decreasing rearrangement in [32]) Suppose that u,vu,v belong to set 𝒜\mathcal{A} which is defined as in (1.35). Define m(𝐰)=min{u(𝐰),v(𝐰)}m(\bm{w})=\min\{u(\bm{w}),v(\bm{w})\} and M(𝐰)=max{u(𝐰),v(𝐰)}M(\bm{w})=\max\{u(\bm{w}),v(\bm{w})\}. Then

(3.5) [F(u(𝒘))+F(v(𝒘))][F(m(𝒘))+F(M(𝒘))]\displaystyle\ \ \ \ [F(u(\bm{w}))+F(v(\bm{w}))]-[F(m(\bm{w}))+F(M(\bm{w}))]
=ΩdΩdK(𝒘𝒘)[(uv)+(𝒘)(uv)(𝒘)+(uv)(𝒘)(uv)+(𝒘)]d𝒘d𝒘.\displaystyle=\int_{\Omega_{d}}\int_{\Omega_{d}}K(\bm{w}-\bm{w}^{\prime})[(u-v)_{+}(\bm{w})(u-v)_{-}(\bm{w}^{\prime})+(u-v)_{-}(\bm{w})(u-v)_{+}(\bm{w}^{\prime})]\mathrm{d}\bm{w}\mathrm{d}\bm{w}^{\prime}.

In particular, F(u(𝐰))+F(v(𝐰))=F(M(𝐰))+F(m(𝐰))F(u(\bm{w}))+F(v(\bm{w}))=F(M(\bm{w}))+F(m(\bm{w})) holds if and only if

(3.6) (u(𝒘)v(𝒘))(u(𝒘)v(𝒘))0\displaystyle(u(\bm{w})-v(\bm{w}))(u(\bm{w}^{\prime})-v(\bm{w}^{\prime}))\geq 0

holds for almost every 𝐰,𝐰\bm{w},\bm{w}^{\prime} in Ωd\Omega_{d}, i.e. either u(𝐰)v(𝐰)u(\bm{w})\geq v(\bm{w}) or u(𝐰)v(𝐰)u(\bm{w})\leq v(\bm{w}) holds a.e. in Ωd\Omega_{d}.

Proof.

Recall energy functional defined in (1.37), i.e.

F(u)\displaystyle F(u) =12ΩdΩd|(u(𝒘)u(𝒘)|2K(𝒘𝒘)\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(u(\bm{w})-u(\bm{w}^{\prime})|^{2}}K(\bm{w}-\bm{w}^{\prime})
|(η(𝒘)η(𝒘)|2K(𝒘𝒘)d𝒘d𝒘+Ωdγ(u(𝒘))d𝒘.\displaystyle-{|(\eta(\bm{w})-\eta(\bm{w}^{\prime})|^{2}}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w^{\prime}}+\int_{\Omega_{d}}\gamma(u(\bm{w}))\mathrm{d}\bm{w}.

since γ(u(𝒘))\gamma(u(\bm{w})) is a local term, we have

Ωdγ(u(𝒘))+γ(v(𝒘))d𝒘=Ωdγ(m(𝒘))+γ(M(𝒘))d𝒘.\displaystyle\int_{\Omega_{d}}\gamma(u(\bm{w}))+\gamma(v(\bm{w}))\mathrm{d}\bm{w}=\int_{\Omega_{d}}\gamma(m(\bm{w}))+\gamma(M(\bm{w}))\mathrm{d}\bm{w}.

So we only need to compare the convolution term. A straightforward calculation implies that

[F(u(𝒘))+F(v(𝒘))][F(m(𝒘))+F(M(𝒘))]\displaystyle\ \ \ \ [F(u(\bm{w}))+F(v(\bm{w}))]-[F(m(\bm{w}))+F(M(\bm{w}))]
=12ΩdΩdK(𝒘𝒘)[|u(𝒘)u(𝒘)|2+|v(𝒘)v(𝒘)|2\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}K(\bm{w}-\bm{w}^{\prime})[|u(\bm{w})-u(\bm{w}^{\prime})|^{2}+|v(\bm{w})-v(\bm{w}^{\prime})|^{2}
|m(𝒘)m(𝒘)|2|M(𝒘)M(𝒘)|2]d𝒘d𝒘.\displaystyle-|m(\bm{w})-m(\bm{w}^{\prime})|^{2}-|M(\bm{w})-M(\bm{w}^{\prime})|^{2}]\mathrm{d}\bm{w}\mathrm{d}\bm{w}^{\prime}.

By the definition of mm and MM, we know that

u(𝒘)2+v(𝒘)2=m(𝒘)2+M(𝒘)2\displaystyle u(\bm{w})^{2}+v(\bm{w})^{2}=m(\bm{w})^{2}+M(\bm{w})^{2}

holds for every 𝒘Ωd\bm{w}\in\Omega_{d}, thus

[F(u(𝒘))+F(v(𝒘))][F(m(𝒘))+F(M(𝒘))]\displaystyle\ \ \ \ [F(u(\bm{w}))+F(v(\bm{w}))]-[F(m(\bm{w}))+F(M(\bm{w}))]
=ΩdΩdK(𝒘𝒘)[M(𝒘)M(𝒘)+m(𝒘)m(𝒘)u(𝒘)u(𝒘)v(𝒘)v(𝒘)]d𝒘d𝒘.\displaystyle=\int_{\Omega_{d}}\int_{\Omega_{d}}K(\bm{w}-\bm{w}^{\prime})[M(\bm{w})M(\bm{w}^{\prime})+m(\bm{w})m(\bm{w}^{\prime})-u(\bm{w})u(\bm{w}^{\prime})-v(\bm{w})v(\bm{w}^{\prime})]\mathrm{d}\bm{w}\mathrm{d}\bm{w}^{\prime}.

Let a1=u(𝒘),a2=v(𝒘),b1=u(𝒘),b2=v(𝒘)a_{1}=u(\bm{w}),\ a_{2}=v(\bm{w}),\ b_{1}=u(\bm{w}^{\prime}),\ b_{2}=v(\bm{w}^{\prime}), then in terms of Lemma 3.1 we know that

m(𝒘)=a,M(𝒘)=A,m(𝒘)=b,M(𝒘)=B.\displaystyle m(\bm{w})=a,\ M(\bm{w})=A,\ m(\bm{w}^{\prime})=b,\ M(\bm{w}^{\prime})=B.

By Lemma 3.1, we have

[F(u(𝒘))+F(v(𝒘))][F(m(𝒘))+F(M(𝒘))]\displaystyle\ \ \ \ [F(u(\bm{w}))+F(v(\bm{w}))]-[F(m(\bm{w}))+F(M(\bm{w}))]
=ΩdΩdK(𝒘𝒘)[(uv)+(𝒘)(uv)(𝒘)+(uv)(𝒘)(uv)+(𝒘)]d𝒘d𝒘.\displaystyle=\int_{\Omega_{d}}\int_{\Omega_{d}}K(\bm{w}-\bm{w}^{\prime})[(u-v)_{+}(\bm{w})(u-v)_{-}(\bm{w}^{\prime})+(u-v)_{-}(\bm{w})(u-v)_{+}(\bm{w}^{\prime})]\mathrm{d}\bm{w}\mathrm{d}\bm{w}^{\prime}.

Thus (3.5) holds. So in terms of integrating 𝒘,𝒘\bm{w},\bm{w}^{\prime}, F(u(𝒘))+F(v(𝒘))=F(m(𝒘))+F(M(𝒘))F(u(\bm{w}))+F(v(\bm{w}))=F(m(\bm{w}))+F(M(\bm{w})) holds if and only if (u(𝒘)v(𝒘))(u(𝒘)v(𝒘))0(u(\bm{w})-v(\bm{w}))(u(\bm{w}^{\prime})-v(\bm{w}^{\prime}))\geq 0 holds a.e. in Ωd\Omega_{d}. This concludes the proof. ∎

Lemma 3.2 ensures that if we are given two functions u,vu,v defined on Ωd\Omega_{d}, we can construct a pair m,Mm,M such that they have a total energy less than F(u)+F(v)F(u)+F(v). Here comes the name of this tool: the energy decreasing property of this construction m,Mm,M is realized by the precedent rearrangement (Lemma 3.1), so we name it as energy decreasing rearrangement method. Now we are ready to prove Proposition 1 using Lemma 3.2.

3.2. Relationship with the increasing rearrangement

Clarification on this rearrangement technique is necessary to help readers distinguish it from other similar tools. Another rearrangement skill broadly utilized in the calculus of variations is the increasing rearrangement, which was first introduced in [33]. Given u:u:\mathbb{R}\to\mathbb{R} satisfying limx±u(x)=±1\lim\limits_{x\to\pm\infty}u(x)=\pm 1, the increasing rearrangement of uu, denoted as uu^{*}, is an increasing function with sublevel sets which are of same volume as those of uu, i.e.,

(3.7) {x:tu(x)}={x:tu(x)}foreveryt(1,1).\displaystyle\{x:t\leq u^{*}(x)\}=\{x:t\leq u(x)\}^{*}\ \mathrm{for}\ \mathrm{every}\ t\in(-1,1).

Here the rearrangement of a Borel set AA\in\mathbb{R}, i.e. AA^{*}, is defined as

(3.8) A:=[c,),c:=b|A[a,b]|foreveryAsatisfying[b,)A[a,).\displaystyle A^{*}:=[c,\infty),\ c:=b-|A\cap[a,b]|\ \mathrm{for}\ \mathrm{every}\ A\ \mathrm{satisfying}\ [b,\infty)\subset A\subset[a,\infty).

The machinery of the increasing rearrangement is exactly the same as that of the cumulative density function (CDF) matching approach. The measure-preserving property maintains local functional energies (e.g., the double-well potential), while the monotonicity reduces the convolution-type non-local functional energy (e.g., the reduction of the elastic energy on the slip plane).

Due to this energy reduction property, the increasing rearrangement is also employed to minimize functional energies that share common structures with FF in (1.37) [3]. Although results derived in [3] are similar to ours, the context of [3] is much different in the sense that the convolution kernel J(h)J(h) satisfies

(3.9) J(h)L1(d),J(h)|h|L1(d),\displaystyle J(h)\in L^{1}(\mathbb{R}^{d}),J(h)|h|\in L^{1}(\mathbb{R}^{d}),

while in the current context we have

(3.10) K(h)|h|d1L1(d),K(h)|h||h|dL1(d).\displaystyle K(h)\sim|h|^{-d-1}\notin L^{1}(\mathbb{R}^{d}),\ K(h)|h|\sim|h|^{-d}\notin L^{1}(\mathbb{R}^{d}).

Therefore, the availability of the increasing rearrangement in our setting is indirect. Application of the increasing rearrangement was also considered in [32] in which the authors commented that it worked for (Δ)α,α(1/2,1)(-\Delta)^{\alpha},\alpha\in(1/2,1) instead of the critical case (Δ)1/2(-\Delta)^{1/2} which is exactly in the PN model.

In contrast, originated in [32], Lemma 3.2 is powerful in this critical case (also other non-critical cases as discussed in [32]) with a much simpler and elementary proof compared to the increasing rearrangement [2]. The main difference between Lemma 3.2 and the increasing rearrangement is that the former rearranges a pair of profiles uu and vv while the latter merely works on a single candidate uu. Therefore, the hidden mechanisms of these two methods are totally different and readers should be aware of this discrepancy.

3.3. Minimizers on finite intervals

Before proving Proposition 1, we introduce the translation-invariant property of the energy functional FF which is applied in the proof.

Lemma 3.3.

(translation invariant) Consider FF in (1.37). Then for any (c1,𝐜𝟐)Ωd(c_{1},\bm{c_{2}})\in\Omega_{d}, we have

F(u(x+c1,𝒚+𝒄𝟐))=F(u(x,𝒚)),\displaystyle F(u(x+c_{1},\bm{y}+\bm{c_{2}}))=F(u(x,\bm{y})),

i.e. FF is invariant under any translation.

The proof of this lemma only relies on some elementary computations of integrals using Lemma 2.2. Readers can refer to Appendix B for detail. We will again use this invariant property later to prove the lower boundedness of FF on 𝒜\mathcal{A}.

Now we are ready to prove Proposition 1 which addresses the minimizer of FF on the set 𝒜I\mathcal{A}_{I} defined in (3.2):

𝒜I={u𝒜:u=ηonΩdΩdI}.\displaystyle\mathcal{A}_{I}=\{u\in\mathcal{A}:u=\eta\ \mathrm{on}\ \Omega_{d}\setminus\Omega_{d}^{I}\}.

We aim to prove that there exists a minimizer uIu_{I} of FF on set 𝒜I\mathcal{A}_{I}.

Proof of Proposition 1.

First we prove statement (i). Notice that for any u𝒜Iu\in\mathcal{A}_{I}, F(u)F(u) is uniformly bounded from below by a constant that depends on η(x)\eta(x), i.e.

F(u)\displaystyle F(u) =12ΩdΩd(|(u(x,𝒚)u(x,𝒚)|2)K(xx,𝒚𝒚)\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}(|(u(x,\bm{y})-u(x^{\prime},\bm{y}^{\prime})|^{2})K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})
|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚+Ωdγ(u(x,𝒚))dxd𝒚\displaystyle-{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}+\int_{\Omega_{d}}\gamma(u(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}
12ΩdIΩdI|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚\displaystyle\geq-\dfrac{1}{2}\int_{\Omega_{d}^{I}}\int_{\Omega_{d}^{I}}{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}
(ΩdI)cΩdI|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxdxdydy\displaystyle-\int_{(\Omega_{d}^{I})^{c}}\int_{\Omega_{d}^{I}}{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}y\mathrm{d}y^{\prime}
=c2II(η(x)η(x))2(xx)2dxdxcIIc(η(x)η(x))2(xx)2dxdx>.\displaystyle=-\dfrac{c_{\mathcal{L}}}{2}\int_{I}\int_{I}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x\mathrm{d}x^{\prime}-c_{\mathcal{L}}\int_{I}\int_{I^{c}}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x>-\infty.

The last equality is by Lemma 2.2. c>0c_{\mathcal{L}}>0 is the constant in Lemma 2.2. Therefore, there exists a minimizing sequence {un}𝒜I\{u_{n}\}\subset\mathcal{A}_{I} such that

(3.11) F(un)CI:=infu𝒜IF(u)asn.\displaystyle F(u_{n})\to C_{I}:=\inf\limits_{u\in\mathcal{A}_{I}}F(u)\ \mathrm{as}\ n\to\infty.

For any u𝒜u\in\mathcal{A}, consider

u~=max{min{u,1},1},\displaystyle\tilde{u}=\max\{\min\{u,1\},-1\},

i.e. u~\tilde{u} is the cut-off of uu from below by 1-1 and from above by 1. Then u~𝒜\tilde{u}\in\mathcal{A} and satisfies

F(u~)F(u)\displaystyle F(\tilde{u})\leq F(u)

by definition of FF. So we can assume |un|1|u_{n}|\leq 1 without loss of generality.

Denote vn=unηv_{n}=u_{n}-\eta. Then vnv_{n} is supported on ΩdI\Omega_{d}^{I} since un𝒜Iu_{n}\in\mathcal{A}_{I}. This indicates that vnL2(Ωd)24(ba)\|v_{n}\|_{L^{2}(\Omega_{d})}^{2}\leq 4(b-a), i.e. {vn}\{v_{n}\} is uniformly bounded in L2(Ωd)L^{2}(\Omega_{d}). Moreover, {vn}\{v_{n}\} is also uniformly bounded in H1/2(Ωd)H^{1/2}(\Omega_{d}) since by Lemma 2.3,

vnH˙1/2(Ωd)2\displaystyle\|v_{n}\|_{\dot{H}^{1/2}(\Omega_{d})}^{2} 1c2ΩΩ|(vn(x,𝒚)vn(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚.\displaystyle\leq\dfrac{1}{c_{2}}\int_{\Omega}\int_{\Omega}{|(v_{n}(x,\bm{y})-v_{n}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}.

Meanwhile, using the definition of FF in (3) and the Cauchy-Schwartz inequality, we have

1c2ΩΩ|(vn(x,𝒚)vn(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚\displaystyle\ \ \ \dfrac{1}{c_{2}}\int_{\Omega}\int_{\Omega}{|(v_{n}(x,\bm{y})-v_{n}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}
=2c2F(un)4cc2Ωdvn(x,𝒚)(xx)1/2η(x)dxd𝒚2c2Ωdγ(u)d𝒘\displaystyle=\dfrac{2}{c_{2}}F(u_{n})-\dfrac{{4c_{\mathcal{L}}}}{c_{2}}\int_{\Omega_{d}}v_{n}(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}-\dfrac{2}{c_{2}}\int_{\Omega_{d}}\gamma(u)\mathrm{d}\bm{w}
2c2F(un)+2cc2vnL2(Ωd)2+2cc2(xx)1/2η(x)L2()2.\displaystyle\leq\dfrac{2}{c_{2}}F(u_{n})+\dfrac{{2c_{\mathcal{L}}}}{c_{2}}\|v_{n}\|^{2}_{L^{2}(\Omega_{d})}+\dfrac{{2c_{\mathcal{L}}}}{c_{2}}\|(-\partial_{xx})^{1/2}\eta(x)\|^{2}_{L^{2}(\mathbb{R})}.
C.\displaystyle\leq C^{\prime}.

Here cc_{\mathcal{L}} is the constant in Lemma 2.2 and c2c_{2} is the constant in Lemma 2.3. CC^{\prime} is a constant that only depends on a,ba,b and η\eta but independent with any certain minimizing sequence. Therefore, {vn}\{v_{n}\} is uniformly bounded in H1/2(Ωd)H^{1/2}(\Omega_{d}).

Now we are ready to prove that uIu_{I} is indeed a minimizer. Uniform boundedness of {vn}\{v_{n}\} in H1/2(Ωd)H^{1/2}(\Omega_{d}) implies that there exists vIH1/2(Ωd)v_{I}\in H^{1/2}(\Omega_{d}) supported on ΩdI\Omega_{d}^{I} such that vnvIv_{n}\rightharpoonup v_{I} in H1/2(Ωd)H^{1/2}(\Omega_{d}). Hence uI:=vI+η𝒜Iu_{I}:=v_{I}+\eta\in\mathcal{A}_{I} and up to a subsequence,

un\displaystyle u_{n} uI,vnvIa.e.inΩd,\displaystyle\to u_{I},\ v_{n}\to v_{I}\quad\mathrm{a.e.}\ \mathrm{in}\ \Omega_{d},
un\displaystyle u_{n} uI,vnvIinL2(ΩdI).\displaystyle\to u_{I},\ v_{n}\to v_{I}\quad\mathrm{in}\ L^{2}(\Omega_{d}^{I}).

Therefore, by Fatou’s lemma, the strong L2L^{2} convergence and the definition of FF in (3), we know that

CI\displaystyle C_{I} =lim infnF(un)\displaystyle=\liminf\limits_{n\to\infty}F(u_{n})
=lim infn12ΩdΩd|(vn(x,𝒚)vn(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚\displaystyle=\liminf\limits_{n\to\infty}\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(v_{n}(x,\bm{y})-v_{n}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}
+2cΩdvn(x,𝒚)(xx)1/2η(x)dxd𝒚+Ωdγ(un)dxd𝒚\displaystyle+{2c_{\mathcal{L}}}\int_{\Omega_{d}}v_{n}(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}+\int_{\Omega_{d}}\gamma(u_{n})\mathrm{d}x\mathrm{d}\bm{y}
12ΩdΩd|(vI(x,𝒚)vI(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚\displaystyle\geq\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(v_{I}(x,\bm{y})-v_{I}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}
+2cΩdvI(x,𝒚)(xx)1/2η(x)dxd𝒚+Ωdγ(uI)dxd𝒚\displaystyle+{2c_{\mathcal{L}}}\int_{\Omega_{d}}v_{I}(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}+\int_{\Omega_{d}}\gamma(u_{I})\mathrm{d}x\mathrm{d}\bm{y}
=F(uI)CI.\displaystyle=F(u_{I})\geq C_{I}.

Thus uI𝒜Iu_{I}\in\mathcal{A}_{I} is indeed a minimizer of FF on set 𝒜I\mathcal{A}_{I}. In particular, it is a weak solution to (3.3) by a simple calculation of the first variation of the energy functional FF. This proves (i).

We will use the energy decreasing rearrangement method (Lemma 3.2) to prove (ii) and (iii). First we prove (ii). For any given τ>0\tau>0, consider v(x,𝒚)=uI(x+τ,𝒚)v(x,\bm{y})=u_{I}(x+\tau,\bm{y}). Denote m(x,𝒚)=min{uI(x,𝒚),v(x,𝒚)}m(x,\bm{y})=\min\{u_{I}(x,\bm{y}),v(x,\bm{y})\} and M(x,𝒚)=max{uI(x,𝒚),v(x,𝒚)}M(x,\bm{y})=\max\{u_{I}(x,\bm{y}),v(x,\bm{y})\}. Then by Lemma 3.2, we know that

F(m)+F(M)F(uI)+F(v).\displaystyle F(m)+F(M)\leq F(u_{I})+F(v).

This inequality is in fact an equality. Notice that M(x,𝒚)=1M(x,\bm{y})=1 if xbτx\geq b-\tau and M(x,𝒚)=1M(x,\bm{y})=-1 if xaτx\leq a-\tau, so M(x,𝒚)𝒜(aτ,bτ)M(x,\bm{y})\in\mathcal{A}_{(a-\tau,b-\tau)}. By the translation invariant property (Lemma 3.3), we know that v(x,𝒚)=uI(x+τ,𝒚)v(x,\bm{y})=u_{I}(x+\tau,\bm{y}) is in fact a minimizer of FF on 𝒜(aτ,bτ).\mathcal{A}_{(a-\tau,b-\tau)}. Thus

F(M)F(v).\displaystyle F(M)\geq F(v).

Note that m(x,𝒚)=1m(x,\bm{y})=1 if xbx\geq b and m(x,𝒚)=1m(x,\bm{y})=-1 if xax\leq a, so m(x,𝒚)𝒜m(x,\bm{y})\in\mathcal{A}. Then by minimality of uIu_{I} we know that

F(m)F(uI).\displaystyle F(m)\geq F(u_{I}).

Therefore, we have

F(v)+F(uI)F(m)+F(M)F(uI)+F(v).\displaystyle F(v)+F(u_{I})\leq F(m)+F(M)\leq F(u_{I})+F(v).

Thus

F(v)+F(uI)=F(m)+F(M).\displaystyle F(v)+F(u_{I})=F(m)+F(M).

By Lemma 3.2, this equality holds if and only if either uI(x,𝒚)v(x,𝒚)u_{I}(x,\bm{y})\geq v(x,\bm{y}) or uI(x,𝒚)v(x,𝒚)u_{I}(x,\bm{y})\leq v(x,\bm{y}) holds almost surely in Ωd\Omega_{d}. By the boundary condition and that |uI|1|u_{I}|\leq 1, we know that the former is true, i.e.

uI(x,𝒚)v(x,𝒚)=uI(x+τ,𝒚).\displaystyle u_{I}(x,\bm{y})\leq v(x,\bm{y})=u_{I}(x+\tau,\bm{y}).

This inequality holds for a.e. (x,𝒚)ΩdI(x,\bm{y})\in\Omega_{d}^{I} for arbitrary τ>0\tau>0. This proves (ii).

Eventually, we prove (iii). Again we will adopt Lemma 3.2, i.e. the energy decreasing rearrangement method. Unlike the case in the proof of (ii) where we only consider translation in xx direction, we consider translation in both xx and 𝒚\bm{y} direction, but with xx direction still positive. For any given (τ1,𝝉𝟐)(\tau_{1},\bm{\tau_{2}}) such that τ1>0,𝝉𝟐𝕋d1\tau_{1}>0,\bm{\tau_{2}}\in\mathbb{T}^{d-1}, consider w(x,𝒚)=uI(x+τ1,𝒚+𝝉𝟐)w(x,\bm{y})=u_{I}(x+\tau_{1},\bm{y}+\bm{\tau_{2}}). As in the proof of (ii), by considering the minimum and maximum of uIu_{I} and ww, we conclude that

(3.12) uI(x+τ1,𝒚+𝝉𝟐)uI(x,𝒚)\displaystyle u_{I}(x+\tau_{1},\bm{y}+\bm{\tau_{2}})\geq u_{I}(x,\bm{y})

holds for almost every (x,𝒚)ΩdI(x,\bm{y})\in\Omega_{d}^{I}.

Now let τ10\tau_{1}\to 0. For any 𝒘Ωd\bm{w}\in\Omega_{d}, denote Sϵ(𝒘)S_{\epsilon}(\bm{w}) the square with length ϵ\epsilon centered at 𝒘\bm{w}. Then for any (x,𝒚)ΩI(x,\bm{y})\in\Omega_{I} (not almost every but every) and ϵ>0\epsilon>0, by inequality (3.12), we have

1ϵdSϵ(x,𝒚)uI(s,𝒕+𝝉𝟐)dsd𝒕\displaystyle\dfrac{1}{\epsilon^{d}}\int_{S_{\epsilon}(x,\bm{y})}u_{I}(s,\bm{t}+\bm{\tau_{2}})\mathrm{d}s\mathrm{d}\bm{t} =limτ10+1ϵdSϵ(x,𝒚)uI(s+τ1,𝒕+𝝉𝟐)dsd𝒕\displaystyle=\lim\limits_{\tau_{1}\to 0^{+}}\dfrac{1}{\epsilon^{d}}\int_{S_{\epsilon}(x,\bm{y})}u_{I}(s+\tau_{1},\bm{t}+\bm{\tau_{2}})\mathrm{d}s\mathrm{d}\bm{t}
1ϵdSϵ(x,𝒚)uI(s,𝒕)dsd𝒕\displaystyle\geq\dfrac{1}{\epsilon^{d}}\int_{S_{\epsilon}(x,\bm{y})}u_{I}(s,\bm{t})\mathrm{d}s\mathrm{d}\bm{t}

Then let ϵ0\epsilon\to 0 and by Lebesgue’s differential theorem, we have

uI(x,𝒚+𝝉𝟐)u(x,𝒚)\displaystyle u_{I}(x,\bm{y}+\bm{\tau_{2}})\geq u(x,\bm{y})

holds for a.e. (x,𝒚)ΩdI(x,\bm{y})\in\Omega_{d}^{I}. This holds for arbitrary 𝝉𝟐𝕋d1\bm{\tau_{2}}\in\mathbb{T}^{d-1} without specific assignment of sign of each component. Then taking both 𝝉𝟐\bm{\tau_{2}} and 𝝉𝟐-\bm{\tau_{2}} in the translation concludes that

uI(x,𝒚+𝝉𝟐)=uI(x,𝒚)\displaystyle u_{I}(x,\bm{y}+\bm{\tau_{2}})=u_{I}(x,\bm{y})

holds for a.e. (x,𝒚)ΩdI(x,\bm{y})\in\Omega_{d}^{I}. This closes the whole proof of Proposition 1. ∎

3.4. Technical lemmas

Before proving the existence theorem, i.e. Theorem 1, we will first provide several technical lemmas whose proofs are attached in Appendix B. These lemmas finally lead to the fact that FF is lower bounded on 𝒜\mathcal{A}. This enables the application of the direct method in calculus of variations in the proof of Theorem 1.

Lemma 3.4 addresses an approximation property:

Lemma 3.4.

For any u𝒜u\in\mathcal{A} such that |u|1|u|\leq 1, there exist a sequence {un}𝒜\{u_{n}\}\subset\mathcal{A} and positive constants {Mn}\{M_{n}\} such that

unηC(Ωd),un=ηon|x|>Mn,\displaystyle u_{n}-\eta\in C^{\infty}(\Omega_{d}),\ u_{n}=\eta\ \mathrm{on}\ |x|>M_{n},

and

F(un)F(u)asn.\displaystyle F(u_{n})\to F(u)\ \mathrm{as}\ n\to\infty.
Proof.

Given u𝒜u\in\mathcal{A} such that |u|1|u|\leq 1, because uηH1/2(Ωd)u-\eta\in H^{1/2}(\Omega_{d}), so standard density argument (see [1, 38]) claims that there exists un{u_{n}} and {Mn}\{M_{n}\} such that unηC(Ωd)u_{n}-\eta\in C^{\infty}(\Omega_{d}), un=ηu_{n}=\eta on |x|>Mn|x|>M_{n} and (uη)(unη)H1/2(Ωd)20.\|(u-\eta)-(u_{n}-\eta)\|_{H^{1/2}(\Omega_{d})}^{2}\to 0. as nn\to\infty. Therefore,

unη\displaystyle u_{n}-\eta uηinH1/2(Ωd),\displaystyle\to u-\eta\quad\mathrm{in}\ H^{1/2}(\Omega_{d}),
unη\displaystyle u_{n}-\eta uηinL2(Ωd).\displaystyle\to u-\eta\quad\mathrm{in}\ L^{2}(\Omega_{d}).

Denote v=uηv=u-\eta, vn=unηv_{n}=u_{n}-\eta. Then by (3), we have

(3.13) F(u)\displaystyle F(u) =12ΩdΩd|v(𝒘)v(𝒘)|2K(𝒘𝒘)d𝒘d𝒘\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}|v(\bm{w})-v(\bm{w^{\prime}})|^{2}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w^{\prime}}
+cΩd(uη)(xx)1/2η(x)dxd𝒚+Ωdγ(u)dxd𝒚,\displaystyle+c_{\mathcal{L}}\int_{\Omega_{d}}(u-\eta)(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}+\int_{\Omega_{d}}\gamma(u)\mathrm{d}x\mathrm{d}\bm{y},

here cc_{\mathcal{L}} is the constant in Lemma 2.2. Then by Lemma 2.3 and convergence in H1/2(Ωd)H^{1/2}(\Omega_{d}) and L2(Ωd)L^{2}(\Omega_{d}), we have

ΩdΩd|vn(𝒘)vn(𝒘)|2K(𝒘𝒘)d𝒘d𝒘\displaystyle\int_{\Omega_{d}}\int_{\Omega_{d}}|v_{n}(\bm{w})-v_{n}(\bm{w^{\prime}})|^{2}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w^{\prime}} ΩdΩd|v(𝒘)v(𝒘)|2K(𝒘𝒘)d𝒘d𝒘\displaystyle\to\int_{\Omega_{d}}\int_{\Omega_{d}}|v(\bm{w})-v(\bm{w^{\prime}})|^{2}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w^{\prime}}
Ωd(unη)(xx)1/2η(x)dxd𝒚\displaystyle\int_{\Omega_{d}}(u_{n}-\eta)(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y} Ωd(uη)(xx)1/2η(x)dxd𝒚\displaystyle\to\int_{\Omega_{d}}(u-\eta)(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}

as nn\to\infty.

For the non-linear potential term in (3.13), the mean value theorem ensures that there exist θ(x,y)\theta(x,y) and θ~(x,y)[0,1]\tilde{\theta}(x,y)\in[0,1] such that

|Ωdγ(un)γ(u)dxd𝒚|\displaystyle\ \ \ \left|\int_{\Omega_{d}}\gamma(u_{n})-\gamma(u)\mathrm{d}x\mathrm{d}\bm{y}\right|
Ωd|γ(θu+(1θ)un)||uun|dxd𝒚\displaystyle\leq\int_{\Omega_{d}}|\gamma^{\prime}(\theta u+(1-\theta)u_{n})||u-u_{n}|\mathrm{d}x\mathrm{d}\bm{y}
Ωd|γ(η+θ(uη)+(1θ)(unη))||uun|dxd𝒚\displaystyle\leq\int_{\Omega_{d}}|\gamma^{\prime}(\eta+\theta(u-\eta)+(1-\theta)(u_{n}-\eta))||u-u_{n}|\mathrm{d}x\mathrm{d}\bm{y}
Ωd|γ(η)||uun|dxd𝒚\displaystyle\leq\int_{\Omega_{d}}|\gamma^{\prime}(\eta)||u-u_{n}|\mathrm{d}x\mathrm{d}\bm{y}
+Ωd|γ′′(η+θ~θ(uη)+θ~(1θ)(unη))||θ(uη)+(1θ)(unη)||uun|dxd𝒚.\displaystyle+\int_{\Omega_{d}}|\gamma^{\prime\prime}(\eta+\tilde{\theta}\theta(u-\eta)+\tilde{\theta}(1-\theta)(u_{n}-\eta))||\theta(u-\eta)+(1-\theta)(u_{n}-\eta)||u-u_{n}|\mathrm{d}x\mathrm{d}\bm{y}.

Because |u|1|u|\leq 1, we can assume |un|1|u_{n}|\leq 1 without loss of generality. Thus

|γ′′(η+θ~θ(uη)+θ~(1θ)(unη))|\displaystyle|\gamma^{\prime\prime}(\eta+\tilde{\theta}\theta(u-\eta)+\tilde{\theta}(1-\theta)(u_{n}-\eta))|

is uniformly bounded in Ωd\Omega_{d}. Also notice that γ(η)=0\gamma^{\prime}(\eta)=0 for |x|>1|x|>1, then by the Cauchy-Schwartz inequality, there exists C>0C>0 that only depends on u,ηu,\eta such that

|Ωdγ(un)γ(u)dxd𝒚|\displaystyle\left|\int_{\Omega_{d}}\gamma(u_{n})-\gamma(u)\mathrm{d}x\mathrm{d}\bm{y}\right| 11𝕋d1|γ(η)||uun|d𝒚dx\displaystyle\leq\int_{-1}^{1}\int_{\mathbb{T}^{d-1}}|\gamma^{\prime}(\eta)||u-u_{n}|\mathrm{d}\bm{y}\mathrm{d}x
+C(uηL2(Ωd)+unηL2(Ωd))uunL2(Ωd)\displaystyle+C(\|u-\eta\|_{L^{2}(\Omega_{d})}+\|u_{n}-\eta\|_{L^{2}(\Omega_{d})})\|u-u_{n}\|_{L^{2}(\Omega_{d})}
CuunL2(Ωd)0.\displaystyle\leq C^{\prime}\|u-u_{n}\|_{L^{2}(\Omega_{d})}\to 0.

Here CC^{\prime} is a constant that only depends on γ,η\gamma,\eta and uu. This closes the proof. ∎

The following lemma claims that we can use the nonlinear potential to control L2L^{2} norm of uηu-\eta.

Lemma 3.5.

Suppose that uu is a non-decreasing function on \mathbb{R} such that u(x)=v(x)+η(x)u(x)=v(x)+\eta(x) is non-decreasing, |u(x)|1|u(x)|\leq 1 for all xx\in\mathbb{R} and u(0)=0u(0)=0. γC2()\gamma\in C^{2}(\mathbb{R}) satisfies (1.8). Then there exist constants C1C_{1} and C2C_{2} such that

γ(u(x))dx+C1C2vL22.\displaystyle\int_{\mathbb{R}}\gamma(u(x))\mathrm{d}x+C_{1}\geq C_{2}\|v\|_{L^{2}}^{2}.

Here C1>0C_{1}>0 and C2>0C_{2}>0 only depend on γ(x)\gamma(x) and are independent with vv.

Proof.

According to (1.8), γ′′(±1)>0\gamma^{\prime\prime}(\pm 1)>0 and γ\gamma attains strict minimum at 1-1 and 11, so there exists C1>0C_{1}>0 such that

γ(x)C1(x1)2,ifx[0,1],\displaystyle\gamma(x)\geq C_{1}(x-1)^{2},\ \mathrm{if}\ x\in[0,1],
γ(x)C1(x+1)2,ifx[1,0].\displaystyle\gamma(x)\geq C_{1}(x+1)^{2},\ \mathrm{if}\ x\in[-1,0].

Remember that u(x)u(x) is non-decreasing, u(0)=0u(0)=0 and 1η(x)1-1\leq\eta(x)\leq 1, so

1\displaystyle-1 v(x)0,ifx1,\displaystyle\leq v(x)\leq 0,\ \mathrm{if}\ x\geq 1,
0\displaystyle 0 v(x)1,ifx1.\displaystyle\leq v(x)\leq 1,\ \mathrm{if}\ x\leq-1.

Therefore, we have

γ(u(x))dx\displaystyle\int_{\mathbb{R}}\gamma(u(x))\mathrm{d}x 1γ(v(x)1)dx+1+γ(v(x)+1)dx\displaystyle\geq\int_{-\infty}^{-1}\gamma(v(x)-1)\mathrm{d}x+\int_{1}^{+\infty}\gamma(v(x)+1)\mathrm{d}x
C11v(x)2dx+C11+v(x)2dx\displaystyle\geq C_{1}\int_{-\infty}^{-1}v(x)^{2}\mathrm{d}x+C_{1}\int_{1}^{+\infty}v(x)^{2}\mathrm{d}x
C1vL222C1.\displaystyle\geq C_{1}\|v\|_{L_{2}}^{2}-2C_{1}.

Using these technical lemmas, we are ready to prove Theorem 1.

3.5. Proof of Theorem 1: existence of the minimizers

As stated in previous sections, we will use the calculus of variations to prove Theorem 1 by minimizing FF on set 𝒜\mathcal{A}. Proved in Proposition 1, a key property of the minimizers uIu_{I} is that for and τ1>0,𝝉2𝕋d1\tau_{1}>0,\ \bm{\tau}_{2}\in\mathbb{T}^{d-1},

(3.14) uI(x+τ1,𝒚)\displaystyle u_{I}(x+\tau_{1},\bm{y}) uI(x,𝒚)a.e.inΩd,\displaystyle\geq u_{I}(x,\bm{y})\ a.e.\ \mathrm{in}\ \Omega_{d},
uI(x,𝒚+𝝉2)\displaystyle u_{I}(x,\bm{y}+\bm{\tau}_{2}) =uI(x,𝒚)a.e.inΩd.\displaystyle=u_{I}(x,\bm{y})\ a.e.\ \mathrm{in}\ \Omega_{d}.

To prove lower boundedness of FF with the help of (3.14), we consider the following subset of 𝒜\mathcal{A} which is much finer than 𝒜\mathcal{A}:

(3.15) :={u𝒜:usatisfies(3.14),|u|1andu(0)=0}.\displaystyle\mathcal{B}:=\left\{u\in\mathcal{A}:\ u\ \mathrm{satisfies}\ \eqref{condition:key},\ |u|\leq 1\ \mathrm{and}\ u(0)=0\right\}.

This definition is inspired by Proposition 1 and preceding technical lemmas: according to Proposition 1, we know that uIu_{I}\in\mathcal{B} if uI(0)=0u_{I}(0)=0. Here uIu_{I} is the minimizer of FF on 𝒜I\mathcal{A}_{I} which is constructed in Proposition 1. Through the bridge of set \mathcal{B}, we will prove that:

Lemma 3.6.

Consider set 𝒜\mathcal{A} in (1.35), set \mathcal{B} in (3.15), and functional energy FF in (1.37). Then:

  1. (i)

    infu𝒜F(u)=infuF(u)\inf\limits_{u\in\mathcal{A}}F(u)=\inf\limits_{u\in\mathcal{B}}F(u).

  2. (ii)

    There exist positive constants C3C_{3} and C4C_{4} that only depend on η\eta and γ\gamma such that for any uu\in\mathcal{B},

    F(u)C3uηH1/2(Ωd)2C4.\displaystyle F(u)\geq C_{3}\|u-\eta\|^{2}_{H^{1/2}(\Omega_{d})}-C_{4}.
Proof.

We first prove (i). Because 𝒜\mathcal{B}\subset\mathcal{A}, so we have infu𝒜F(u)infuF(u)\inf\limits_{u\in\mathcal{A}}F(u)\leq\inf\limits_{u\in\mathcal{B}}F(u). Hence we only need to prove that infu𝒜F(u)infuF(u)\inf\limits_{u\in\mathcal{A}}F(u)\geq\inf\limits_{u\in\mathcal{B}}F(u).

Consider u~=max{min{u,1},1}\tilde{u}=\max\{\min\{u,1\},-1\}, i.e. the cut-off of uu by 1 from above and by 1-1 from below. Then u~\tilde{u} is also in 𝒜\mathcal{A} and satisfies that F(u~)F(u)F(\tilde{u})\leq F(u). So we only need to consider those u𝒜u\in\mathcal{A} such that |u|1|u|\leq 1.

By Lemma 3.4, for any ϵ>0\epsilon>0, there exists u1C(Ω)u_{1}\in C^{\infty}(\Omega) and M>0M>0 such that u1=ηu_{1}=\eta on |x|>M|x|>M and

F(u)>F(u1)ϵ.\displaystyle F(u)>F(u_{1})-\epsilon.

Then according to the definition of 𝒜I\mathcal{A}_{I} in (3.2), we know that u1𝒜Iu_{1}\in\mathcal{A}_{I}. By Proposition 1, we know that F(u1)F(uM)F(u_{1})\geq F(u_{M}) where uMu_{M} is a minimizer of FF on 𝒜I\mathcal{A}_{I} satisfying that uMu_{M} is a 1D profile and increasing in xx direction. Therefore

F(u)F(u1)ϵF(uM)ϵ.\displaystyle F(u)\geq F(u_{1})-\epsilon\geq F(u_{M})-\epsilon.

By the translation-invariant property (Lemma 3.3), we have

F(uM)=F(uM)\displaystyle F(u_{M})=F(u^{*}_{M})

where uMu^{*}_{M} is a translation of uMu_{M} that crosses (0,0)(0,0), i.e.

uM(x)=uM(x+c),uM(0)=0.\displaystyle u^{*}_{M}(x)=u_{M}(x+c),\ u^{*}_{M}(0)=0.

By definition, we know that uMu^{*}_{M}\in\mathcal{B}. Thus for any u𝒜u\in\mathcal{A} and ϵ>0\epsilon>0, there exists uMu^{*}_{M}\in\mathcal{B} such that

F(u)F(u)ϵF(uM)ϵ.\displaystyle F(u)\geq F(u)-\epsilon\geq F(u_{M}^{*})-\epsilon.

Thus infu𝒜F(u)infuF(u)ϵ\inf\limits_{u\in\mathcal{A}}F(u)\geq\inf\limits_{u\in\mathcal{B}}F(u)-\epsilon. By arbitrariness of ϵ\epsilon, we have infu𝒜F(u)infuF(u)\inf\limits_{u\in\mathcal{A}}F(u)\geq\inf\limits_{u\in\mathcal{B}}F(u).

Now we prove (ii). By Lemma 3.5, for any uu\in\mathcal{B}, there exist C1C_{1} and C2C_{2} that only depend on γ\gamma such that

Ωdγ(u)dxd𝒚C1vL2(Ωd)2C2\displaystyle\int_{\Omega_{d}}\gamma(u)\mathrm{d}x\mathrm{d}\bm{y}\geq C_{1}\|v\|_{L^{2}(\Omega_{d})}^{2}-C_{2}

where v=uηv=u-\eta. Therefore, using the expression of FF in (3), the Cauchy-Schwartz inequality, Lemma 2.3 and Lemma 2.2, we have

F(u)\displaystyle F(u) =12ΩdΩd|v(𝒘)v(𝒘)|2K(𝒘𝒘)d𝒘d𝒘+2cΩdv(x,𝒚)(xx)1/2η(x)dxd𝒚\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}|v(\bm{w})-v(\bm{w}^{\prime})|^{2}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w}^{\prime}+{2c_{\mathcal{L}}}\int_{\Omega_{d}}v(x,\bm{y}){(-\partial_{xx})^{1/2}}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}
+Ωdγ(u(x,𝒚))dxd𝒚\displaystyle+\int_{\Omega_{d}}\gamma(u(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}
c22vH˙1/2(Ωd)2+C1vL22C22c2C1(xx)1/2η(x)L2(Ωd)2C12vL22\displaystyle\geq\dfrac{c_{2}}{2}\|v\|^{2}_{\dot{H}^{1/2}(\Omega_{d})}+C_{1}\|v\|_{L^{2}}^{2}-C_{2}-{\dfrac{2c_{\mathcal{L}}^{2}}{C_{1}}}\|(-\partial_{xx})^{1/2}\eta(x)\|^{2}_{L^{2}(\Omega_{d})}-\dfrac{C_{1}}{2}\|v\|_{L^{2}}^{2}
C3vH1/2(Ωd)2C4.\displaystyle\geq C_{3}\|v\|^{2}_{{H}^{1/2}(\Omega_{d})}-C_{4}.

Here c2c_{2} is the constant in Lemma 2.3, cc_{\mathcal{L}} is the constant in Lemma 2.2, and

C3=min{c22,C12},C4=C2+2c2C1(xx)1/2η(x)L2(Ωd)2\displaystyle C_{3}=\min\left\{\dfrac{c_{2}}{2},\dfrac{C_{1}}{2}\right\},C_{4}=C_{2}+{\dfrac{2c_{\mathcal{L}}^{2}}{C_{1}}}\|(-\partial_{xx})^{1/2}\eta(x)\|^{2}_{L^{2}(\Omega_{d})}

are constants that only depend on η\eta, γ\gamma and the operator \mathcal{L}. This concludes the proof. ∎

Lemma 3.6 in fact provides insightful corollaries: first, we have

infu𝒜F(u)=infuF(u)C4.\displaystyle\inf\limits_{u\in\mathcal{A}}F(u)=\inf\limits_{u\in\mathcal{B}}F(u)\geq-C_{4}.

Thus FF is lower bounded in 𝒜\mathcal{A}. Moreover, according to (ii), functional F(u)F(u) can be used to bound H1/2(Ωd)H^{1/2}(\Omega_{d}) norm of uηu-\eta for any uu\in\mathcal{B}. In the proof of Theorem 1, this observation will be used to find an a.e. limit of the minimizing sequence which is proved to be a minimizer of FF on 𝒜\mathcal{A}. Now we are ready to prove Theorem 1.

Proof of Theorem 1.

We will first prove (i). By Lemma 3.6, we know that

infu𝒜F(u)=infuF(u)C4>.\displaystyle\inf\limits_{u\in\mathcal{A}}F(u)=\inf\limits_{u\in\mathcal{B}}F(u)\geq-C_{4}>-\infty.

Denote c=infu𝒜F(u)=infuF(u)c=\inf\limits_{u\in\mathcal{A}}F(u)=\inf\limits_{u\in\mathcal{B}}F(u). Then there exists {un}\{u_{n}\}\subset\mathcal{B} such that F(un)cF(u_{n})\to c as nn\to\infty. Again by Lemma 3.6 (ii), we know that unηH1/2(Ωd)\|u_{n}-\eta\|_{H^{1/2}(\Omega_{d})} is uniformly bounded. Thus there exists uu^{*} such that up to a subsequence,

unη\displaystyle u_{n}-\eta uηa.e.inΩd,\displaystyle\to u^{*}-\eta\quad\mathrm{a.e.}\ \mathrm{in}\ \Omega_{d},
unη\displaystyle u_{n}-\eta uηinH1/2(Ωd).\displaystyle\rightharpoonup u^{*}-\eta\quad\mathrm{in}\ H^{1/2}(\Omega_{d}).

Denote vn=unηv_{n}=u_{n}-\eta and v=uηv^{*}=u^{*}-\eta. Then vnvv_{n}\to v^{*} a.e. in Ωd\Omega_{d} and vnvv_{n}\rightharpoonup v^{*} in H1/2(Ωd)H^{1/2}(\Omega_{d}).

In fact, uu^{*} is a minimizer of FF on 𝒜\mathcal{A}. By Fatou’s lemma, we know that

lim infn12ΩdΩd|(vn(x,𝒚)vn(x,𝒚)|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚+Ωdγ(un(x,𝒚))dxd𝒚\displaystyle\liminf\limits_{n\to\infty}\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(v_{n}(x,\bm{y})-v_{n}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y}^{\prime}+\int_{\Omega_{d}}\gamma(u_{n}(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}
(3.16) 12ΩdΩd|(v(x,𝒚)v(x,𝒚)|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚+Ωdγ(u(x,𝒚))dxd𝒚.\displaystyle\geq\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(v^{*}(x,\bm{y})-v^{*}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y}^{\prime}+\int_{\Omega_{d}}\gamma(u^{*}(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}.

Meanwhile, since vnvv_{n}\to v^{*} weakly in H1/2(Ωd)H^{1/2}(\Omega_{d}), hence also converges weakly in L2(Ωd)L^{2}(\Omega_{d}), thus

(3.17) limnΩdvn(x,𝒚)(xx)1/2η(x)dxd𝒚=Ωdv(x,𝒚)(xx)1/2η(x)dxd𝒚.\displaystyle\lim\limits_{n\to\infty}\int_{\Omega_{d}}v_{n}(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}=\int_{\Omega_{d}}v^{*}(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}.

Substituting (3.5) and (3.17) into the following equality, we have

c\displaystyle c =lim infnF(un)\displaystyle=\liminf\limits_{n\to\infty}F(u_{n})
=lim infn12ΩdΩd|(vn(x,𝒚)vn(x,𝒚)|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚\displaystyle=\liminf\limits_{n\to\infty}\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(v_{n}(x,\bm{y})-v_{n}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y}^{\prime}
+2cΩdvn(x,𝒚)(xx)1/2η(x)dxd𝒚+Ωdγ(un(x,𝒚))dxd𝒚\displaystyle+{2c_{\mathcal{L}}}\int_{\Omega_{d}}v_{n}(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}+\int_{\Omega_{d}}\gamma(u_{n}(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}
12ΩdΩd|(v(x,𝒚)v(x,𝒚)|2K(xx,𝒚𝒚)dxd𝒚dxd𝒚\displaystyle\geq\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(v^{*}(x,\bm{y})-v^{*}(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}\bm{y}\mathrm{d}x^{\prime}\mathrm{d}\bm{y}^{\prime}
+2cΩdv(x,𝒚)(xx)1/2η(x)dxd𝒚+Ωdγ(u(x,𝒚))dxd𝒚\displaystyle+{2c_{\mathcal{L}}}\int_{\Omega_{d}}v^{*}(x,\bm{y})(-\partial_{xx})^{1/2}\eta(x)\mathrm{d}x\mathrm{d}\bm{y}+\int_{\Omega_{d}}\gamma(u^{*}(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}
=F(u).\displaystyle=F(u^{*}).

So uu^{*} is in fact a minimizer of FF on 𝒜\mathcal{A}. In particular, it is a weak solution to (1.23). This proves (i).

Now we prove (ii). Because unu_{n}\in\mathcal{B}, so |un|1|u_{n}|\leq 1 and they are all 1D functions and non-decreasing in xx direction, so the a.e. limit uu^{*} is also non-decreasing in xx direction and is a 1D profile satisfying |u|1|u^{*}|\leq 1. Thus (3.14) holds for almost every (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}. To prove that uηH1(Ωd)u^{*}-\eta\in H^{1}(\Omega_{d}), we first show that γ(u)L2(Ω)\gamma^{\prime}(u^{*})\in L^{2}(\Omega). This is true by the mean value theorem and (3.14):

(3.18) Ωd|γ(u)|2dxd𝒚\displaystyle\int_{\Omega_{d}}|\gamma^{\prime}(u^{*})|^{2}\mathrm{d}x\mathrm{d}\bm{y} =|γ(η+uη)|2dx\displaystyle=\int_{\mathbb{R}}|\gamma^{\prime}(\eta+u^{*}-\eta)|^{2}\mathrm{d}x
=|γ(η)+γ′′(η+θ(uη))(uη)|2dx\displaystyle=\int_{\mathbb{R}}|\gamma^{\prime}(\eta)+\gamma^{\prime\prime}(\eta+\theta(u^{*}-\eta))(u^{*}-\eta)|^{2}\mathrm{d}x
2|γ(η)|2dx+2|γ′′(η+θ(uη))|2|uη|2dx\displaystyle\leq 2\int_{\mathbb{R}}|\gamma^{\prime}(\eta)|^{2}\mathrm{d}x+2\int_{\mathbb{R}}|\gamma^{\prime\prime}(\eta+\theta(u^{*}-\eta))|^{2}|u^{*}-\eta|^{2}\mathrm{d}x
C1+C2vL2.\displaystyle\leq C_{1}^{\prime}+C_{2}^{\prime}\|v^{*}\|_{L^{2}}.

So γ(u)L2(Ωd)\gamma^{\prime}(u^{*})\in L^{2}(\Omega_{d}). Remember that u+γ(u)=0\mathcal{L}u^{*}+\gamma^{\prime}(u^{*})=0, so by Lemma 2.3, we have

cuH˙1(Ωd)uL2(Ωd)=γ(u)L2(Ωd).\displaystyle c\|u^{*}\|_{\dot{H}^{1}(\Omega_{d})}\leq\|\mathcal{L}u^{*}\|_{L^{2}(\Omega_{d})}=\|\gamma^{\prime}(u^{*})\|_{L^{2}(\Omega_{d})}.

So uH˙1(Ωd)u^{*}\in\dot{H}^{1}(\Omega_{d}). Moreover, ηL2(Ωd)\mathcal{L}\eta\in L^{2}(\Omega_{d}) by Lemma 2.2, so uηH1(Ωd)u^{*}-\eta\in H^{1}(\Omega_{d}). In particular, uu^{*} solves equation (1.15) in L2L^{2} sense. By (3.14), we know that uu^{*} is a 1D profile, so uηH1(Ωd)u^{*}-\eta\in H^{1}(\Omega_{d}) implies that limx±u(x,𝒚)=±1\lim\limits_{x\to\pm\infty}u^{*}(x,\bm{y})=\pm 1 holds uniformly in 𝒚\bm{y}. So (ii) holds.

Finally, we prove (iii) and (iv). Boundedness of uu^{*} implies that γ′′(u)uL2(Ωd)\gamma^{\prime\prime}(u^{*})\nabla u^{*}\in L^{2}(\Omega_{d}). Thus u=γ(u)H1(Ωd)\mathcal{L}u^{*}=-\gamma^{\prime}(u^{*})\in H^{1}(\Omega_{d}) and uH˙2(Ωd)u^{*}\in\dot{H}^{2}(\Omega_{d}). Remember that uu^{*} is a 1D profile, so by embedding H2()C1()H^{2}(\mathbb{R})\subset C^{1}(\mathbb{R}), we know that (3.14) implies that

𝒚u(x,𝒚)=0,u(x,𝒚)x0\displaystyle\nabla_{\bm{y}}u^{*}(x,\bm{y})=0,\ \dfrac{\partial u^{*}(x,\bm{y})}{\partial x}\geq 0

holds for any (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}. So (iv) is proved and (iii) is partially proved except the strict monotonicity.

To prove the strict monotonicity, suppose that u(x0,𝒚0)x=0\dfrac{\partial u^{*}(x_{0},\bm{y}_{0})}{\partial x}=0 for some (x0,𝒚0)Ωd(x_{0},\bm{y}_{0})\in\Omega_{d}, taking derivative on both sides of (1.23) yields

u(x0,𝒚0)x=γ′′(u)u(x0,𝒚0)x=0.\displaystyle\mathcal{L}\dfrac{\partial u^{*}(x_{0},\bm{y}_{0})}{\partial x}=-\gamma^{\prime\prime}(u^{*})\dfrac{\partial u^{*}(x_{0},\bm{y}_{0})}{\partial x}=0.

Thus ux=0\mathcal{L}\dfrac{\partial u^{*}}{\partial x}=0 at (x0,𝒚0)(x_{0},\bm{y}_{0}). However, since ux0\dfrac{\partial u^{*}}{\partial x}\geq 0, we know that ux\dfrac{\partial u^{*}}{\partial x} attains minimum at (x0,𝒚0)(x_{0},\bm{y}_{0}). Then by Lemma 2.1, we know that ux=0\dfrac{\partial u^{*}}{\partial x}=0, i.e. uu^{*} is a constant. This contradicts with the far field limit of uu^{*}. So u(x,𝒚)x>0\dfrac{\partial u^{*}(x,\bm{y})}{\partial x}>0. This concludes the whole theorem. ∎

4. The De Giorgi Conjecture and uniqueness of solutions

In Theorem 1, we prove that there exists a minimizer uu^{*} of functional FF on set 𝒜\mathcal{A} who satisfies that uηH1(Ωd)u^{*}-\eta\in H^{1}(\Omega_{d}) and for any (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}, we have

𝒚u(x,𝒚)=0,u(x,𝒚)x>0.\displaystyle\nabla_{\bm{y}}u(x,\bm{y})=0,\ \dfrac{\partial u^{*}(x,\bm{y})}{\partial x}>0.

In particular, we have limx±u(x,𝒚)=±1\lim\limits_{x\to\pm\infty}u^{*}(x,\bm{y})=\pm 1. As Definition 2, we keep the same definition of layer solutions for (1.23).

Definition 3.

We call that u:Ωdu:\Omega_{d}\to\mathbb{R} is a layer solution to (1.23), i.e.

u+γ(u)=0,\displaystyle\mathcal{L}u+\gamma^{\prime}(u)=0,

if for any (x,𝐲)Ωd(x,\bm{y})\in\Omega_{d},

(4.1) u(x,𝒚)x>0,limx±u(x,𝒚)=±1.\displaystyle\dfrac{\partial u(x,\bm{y})}{\partial x}>0,\ \lim\limits_{x\to\pm\infty}u(x,\bm{y})=\pm 1.

As far as we know, results parallel to the De Giorgi conjecture that address the vectorial case, i.e. system (1.13), and (1.23) are still wanting and lack of exploration. In this section, we will prove Theorem 2 which fills in this blank: all layer solutions to (1.13) or (1.23) with H1H^{1} regularity are in fact 1D profiles if we further assume γC()\gamma\in C^{\infty}(\mathbb{R}).

In [9] and related literatures on the De Giorgi conjecture, the standard approach to prove this type of symmetry result is to first derive some Schauder estimates for weak solutions and then using Liouville type theorems to prove 1D symmetry. For example, authors in [9] first derived C2,αC^{2,\alpha} regularity for layer solutions by careful application of theories on elliptic PDEs. Then they noticed the following lemma (see also Lemma 2.6 in [9]), a Liouville type lemma:

Lemma 4.1.

(a Liouville type theorem) Let φLloc(+d¯)\varphi\in L_{loc}^{\infty}(\overline{\mathbb{R}^{d}_{+}}) be a positive function, not necessarily bounded on all of +d\mathbb{R}^{d}_{+}. Suppose that σHloc1(+d¯)\sigma\in H_{loc}^{1}(\overline{\mathbb{R}^{d}_{+}}) satisfies

{σdiv(φ2σ)0in+d,σσn0on+d\displaystyle\begin{cases}-\sigma\mathrm{div}(\varphi^{2}\nabla\sigma)\leq 0&\ \mathrm{in\ }\mathbb{R}^{d}_{+},\\ \sigma\dfrac{\partial\sigma}{\partial n}\leq 0&\ \mathrm{on\ }\partial\mathbb{R}^{d}_{+}\end{cases}

in the weak sense. Assume that, for every R>1R>1, we have

BR+(φσ)2dxCR2\displaystyle\int_{B^{+}_{R}}(\varphi\sigma)^{2}\mathrm{d}x\leq CR^{2}

for some constant CC independent of RR. Then σ\sigma is a constant.

Applying this lemma to function σ=uyi/ux,i=1,2,,d1\sigma=u_{y_{i}}/u_{x},\ i=1,2,...,d-1, where xx direction is the monotone direction for the layer solution uu, they proved the following lemma (see also Lemma 4.2 in [9]):

Lemma 4.2.

Suppose that γC2,α()\gamma\in C^{2,\alpha}(\mathbb{R}) is a double-well potential satisfying (1.8)\eqref{condition:potential}. Assume that d3d\leq 3 and that uu is a bounded solution of

{Δu=0in+d,un=γ(u)on+d.\displaystyle\begin{cases}\Delta u=0&\ \mathrm{in\ }\mathbb{R}^{d}_{+},\\ \dfrac{\partial u}{\partial n}=\gamma^{\prime}(u)&\ \mathrm{on\ }\partial\mathbb{R}^{d}_{+}.\end{cases}

Then there exists a function φCloc1(+d¯)C2(+d)\varphi\in C_{\text{loc}}^{1}(\overline{\mathbb{R}^{d}_{+}})\bigcap C^{2}(\mathbb{R}^{d}_{+}) with φ>0{\varphi>0} in +d¯\overline{\mathbb{R}^{d}_{+}} and such that for every i=1,2,,d1i=1,2,...,d-1,

uyi=ciφin+d\displaystyle\dfrac{\partial u}{\partial y_{i}}=c_{i}\varphi\quad\mathrm{in\ }\mathbb{R}^{d}_{+}

for some constant cic_{i}.

As a straightforward corollary, the one-dimensional symmetry of solutions to (Δ)1/2u+γ(u)=0(-\Delta)^{1/2}u+\gamma^{\prime}(u)=0 is also established.

Instead of adopting any Liouville type theorem to prove Theorem 2, we will develop a new approach that is first utilized in our previous work [18] to prove 1D symmetry of layer solutions to (1.15). Although Liouville type theorem is not employed, we found that the insightful observation provided by Lemma 4.2 in [9] is significant: as long as one can prove that there exist constants ci,i=1,2,d1c_{i},i=1,2,...d-1 such that

uyi=ciux\displaystyle u_{y_{i}}=c_{i}u_{x}

holds, then the profile uu is a 1D profile. Remember the discussion in Section 1, uxu_{x} and uyi,i=1,2,,d1u_{y_{i}},i=1,2,...,d-1 are eigenfunctions of eigenvalue 0 for the linear operator

(4.2) L:H1(Ωd)L2(Ωd)L2(Ωd),Lϕ=ϕ+γ′′(u)ϕ.\displaystyle L:H^{1}(\Omega_{d})\subset L^{2}(\Omega_{d})\to L^{2}(\Omega_{d}),\ L\phi=\mathcal{L}\phi+\gamma^{\prime\prime}(u)\phi.

Therefore, as long as we can prove that 0 is a simple eigenvalue of LL, i.e. the eigenspace of 0 is only 1 dimension, then we prove that uxu_{x} and uyi,i=1,2,,d1u_{y_{i}},i=1,2,...,d-1 are in fact linearly dependent, which indicates 1D symmetry. This is the main idea and approach we will utilize to prove Theorem 2. Following this logic, we will first establish proper regularity results for layer solutions uu in Section 4.1 and then prove Theorem 2 in Section 4.2.

4.1. Regularity results

In this section, we will derive some regularity results for layer solutions to equation (1.23) and some properties of elements in the kernel of LL. Two main results will be derived in this section under assumption γC()\gamma\in C^{\infty}(\mathbb{R}). First, any layer solution of equation (1.23) is in H˙n(Ωd)\dot{H}^{n}(\Omega_{d}) for any n>0n>0 (see Lemma 4.3) and in particular, uu is smooth with bounded derivatives of any order. Second, eigenfunctions of LL with eigenvalue 0 are in Hn(Ωd)H^{n}(\Omega_{d}) for any n>0n>0 and in particular, they decay to 0 uniformly in 𝒚\bm{y} as |x||x|\to\infty (see Lemma 4.4).

As a reminder, we assume γC()\gamma\in C^{\infty}(\mathbb{R}) in this section. Even though this is stronger than C2C^{2} assumption which is generally considered, this setting indeed covers many important cases. For instance, γ(u)=1π2(cos(πu)+1)\gamma(u)=\dfrac{1}{\pi^{2}}(\cos(\pi u)+1) in the PN model and γ(u)=(1u2)2\gamma(u)=(1-u^{2})^{2} in the Allen-Cahn equation [5].

Now we begin to prove these two lemmas. All these lemmas only require that uu is bounded which is ensured by being a layer solution. Using the Gagliardo-Nirenberg interpolation inequality [31] and ideas in [29] (see Proposition 3.9), we will prove that:

Lemma 4.3.

Suppose that γC()\gamma\in C^{\infty}(\mathbb{R}) is a double-well potential satisfying (1.8) and \mathcal{L} is the linear operator defined in (1.24) satisfying assumption (A), (B), (C) and (D). For any dimension d1d\geq 1, if uH˙1(Ωd)u\in\dot{H}^{1}(\Omega_{d}) satisfying uηH1(Ωd)u-\eta\in H^{1}(\Omega_{d}) is a bounded solution to equation (1.23), i.e.

u+γ(u)=0,\displaystyle\mathcal{L}u+\gamma^{\prime}(u)=0,

then uηHn(Ωd)u-\eta\in{H}^{n}(\Omega_{d}) for any n>0n>0. In particular, uu is in H˙n(Ω)\dot{H}^{n}(\Omega) for any n>0n>0 and smooth with bounded derivatives of any order.

Proof.

Taking derivative on both sides of the equation yields

ux+γ′′(u)ux=0.\displaystyle\mathcal{L}u_{x}+\gamma^{\prime\prime}(u)u_{x}=0.

Remember that uu is bounded, so is γ′′(u)\gamma^{\prime\prime}(u) by continuity of γ′′\gamma^{\prime\prime}. Thus γ′′(u)uxL2(Ωd)\gamma^{\prime\prime}(u)u_{x}\in L^{2}(\Omega_{d}) which implies that uxL2(Ωd)\mathcal{L}u_{x}\in L^{2}(\Omega_{d}). Thus by Lemma 2.3, uxH1(Ωd)u_{x}\in H^{1}(\Omega_{d}). This also holds for 𝒚u(x,𝒚)\nabla_{\bm{y}}u(x,\bm{y}). Thus uH˙2(Ωd)u\in\dot{H}^{2}(\Omega_{d}) and uηH2(Ωd)u-\eta\in H^{2}(\Omega_{d}).

Now we prove by induction that uηHn(Ωd)u-\eta\in{H}^{n}(\Omega_{d}) for any positive integer n3n\geq 3. Suppose that uηHm(Ωd)u-\eta\in{H}^{m}(\Omega_{d}), then by the Gagliardo-Nirenberg interpolation inequality, we know that for any 1jm1\leq j\leq m, we have

Dj(uη)Lp(Ω)C(uη)Hm(Ω)auηL(Ω)1a\displaystyle\|D^{j}(u-\eta)\|_{L^{p}(\Omega)}\leq C\|(u-\eta)\|^{a}_{H^{m}(\Omega)}\|u-\eta\|_{L^{\infty}(\Omega)}^{1-a}

Here pp and j/mα1j/m\leq\alpha\leq 1 satisfy

1p=jd+a(12md).\displaystyle\dfrac{1}{p}=\dfrac{j}{d}+a\left(\dfrac{1}{2}-\dfrac{m}{d}\right).

Take α=j/m\alpha=j/m, then we have p=2m/jp=2m/j and

Dj(uη)L2m/j(Ωd).\displaystyle D^{j}(u-\eta)\in L^{2m/j}(\Omega_{d}).

Notice that Dη=0D\eta=0 if |x|>1|x|>1 and ηC(Ωd)\eta\in C^{\infty}(\Omega_{d}), so DjηL2m/j(Ωd)D^{j}\eta\in L^{2m/j}(\Omega_{d}), hence

DjuL2m/j(Ωd)\displaystyle D^{j}u\in L^{2m/j}(\Omega_{d})

Chain rule implies that for any multi-index α\alpha that satisfies |α|=m|\alpha|=m, we have

Dαγ(u)=β1++βk=αCβu(β1)u(β2)u(βk)γ(k+1)(u).\displaystyle D^{\alpha}\gamma^{\prime}(u)=\sum_{\beta_{1}+...+\beta_{k}=\alpha}C_{\beta}u^{(\beta_{1})}u^{(\beta_{2})}...u^{(\beta_{k})}\gamma^{(k+1)}(u).

Here CβC_{\beta} are constants depending on β=(β1,β2,,βk)\beta=(\beta_{1},\beta_{2},...,\beta_{k}). Boundedness of uu and smoothness of γ\gamma ensure that γ(k+1)(u)\gamma^{(k+1)}(u) is also bounded. Remember that DjuL2m/j(Ωd)D^{j}u\in L^{2m/j}(\Omega_{d}), so we have u(βj)L2m/|βj|(Ωd)u^{(\beta_{j})}\in L^{2m/|\beta_{j}|}(\Omega_{d}) for all j=1,2,,kj=1,2,...,k. Thus by Hölder’s inequality, we have

u(β1)u(β2)u(βk)Lq(Ωd)j=1ku(βj)L2m/|βj|(Ωd)\displaystyle\|u^{(\beta_{1})}u^{(\beta_{2})}...u^{(\beta_{k})}\|_{L^{q}(\Omega_{d})}\leq\prod_{j=1}^{k}\|u^{(\beta_{j})}\|_{L^{2m/|\beta_{j}|}(\Omega_{d})}

where qq satisfies

1q=j=1k|βj|2m=|α|2m=12.\displaystyle\dfrac{1}{q}=\sum_{j=1}^{k}\dfrac{|\beta_{j}|}{2m}=\dfrac{|\alpha|}{2m}=\dfrac{1}{2}.

Thus q=2q=2 and u(β1)u(β2)u(βk)L2(Ωd)u^{(\beta_{1})}u^{(\beta_{2})}...u^{(\beta_{k})}\in L^{2}(\Omega_{d}). Thus Dαγ(u)L2(Ωd)D^{\alpha}\gamma^{\prime}(u)\in L^{2}(\Omega_{d}) for any multi-index α\alpha that satisfies |α|=m|\alpha|=m, so Dmγ(u)L2(Ωd)D^{m}\gamma^{\prime}(u)\in L^{2}(\Omega_{d}). Therefore,

Dm(u)L2(Ωd)=Dmγ(u)L2(Ωd)<.\displaystyle\|D^{m}(\mathcal{L}u)\|_{L^{2}(\Omega_{d})}=\|D^{m}\gamma^{\prime}(u)\|_{L^{2}(\Omega_{d})}<\infty.

Thus Dm(u)L2(Ωd)D^{m}(\mathcal{L}u)\in L^{2}(\Omega_{d}). Then by Lemma 2.3 and Assumption (A), we have uηHm+1(Ωd)u-\eta\in H^{m+1}(\Omega_{d}) and uH˙m+1(Ωd)u\in\dot{H}^{m+1}(\Omega_{d}). Thus by induction, uηHn(Ωd)u-\eta\in H^{n}(\Omega_{d}) for any n>0n>0. In particular, this indicates that uu is smooth with bounded derivatives of any order. ∎

Remember that γ\gamma is smooth, so Lemma 4.3 also ensures that γ′′(u)\gamma^{\prime\prime}(u) is smooth and bounded with bounded derivatives of any order. Recall that uxu_{x} is a 0 eigenfunction of operator LL defined in (1.38), so by ellipticity of \mathcal{L} and regularity of γ′′(u)\gamma^{\prime\prime}(u), we can prove that uxu_{x}, or more generally, any 0 eigenfunction of LL should attain Hk(Ωd)H^{k}(\Omega_{d}) regularity for any k>0k>0. As a direct corollary of Lemma 4.3, we have

Lemma 4.4.

Suppose that γC()\gamma\in C^{\infty}(\mathbb{R}) is a double-well potential satisfying (1.8) and \mathcal{L} is the linear operator defined in (1.24) satisfying assumption (1.3). For any dimension d1d\geq 1, if gH1(Ωd)g\in H^{1}(\Omega_{d}) satisfies

g+γ′′(u)g=0,\displaystyle\mathcal{L}g+\gamma^{\prime\prime}(u)g=0,

where uu is a bounded solution of (1.23) as in Lemma 4.3. Then gHn(Ωd)g\in H^{n}(\Omega_{d}) for any n>0n>0. In particular, gg is smooth and

lim|x|g(x,𝒚)=0\displaystyle\lim\limits_{|x|\to\infty}g(x,\bm{y})=0

holds uniformly in 𝐲\bm{y}.

Proof.

By Lemma 4.3, we know that γ′′(u)\gamma^{\prime\prime}(u) is smooth with bounded derivatives of any order. Suppose that gHk(Ωd)g\in H^{k}(\Omega_{d}) for some k1k\geq 1, then

Dk(g)L2(Ωd)=Dk(γ′′(u)g)L2(Ωd)CkDkgL2(Ωd)<.\displaystyle\|D^{k}(\mathcal{L}g)\|_{L^{2}(\Omega_{d})}=\|D^{k}(\gamma^{\prime\prime}(u)g)\|_{L^{2}(\Omega_{d})}\leq C_{k}\|D^{k}g\|_{L^{2}(\Omega_{d})}<\infty.

Here CkC_{k} is a constant that only depends on k,uk,u and γ\gamma. Thus Dk(g)L2(Ωd)D^{k}(\mathcal{L}g)\in L^{2}(\Omega_{d}), hence by Lemma 2.3 and Assumption (A), we know gHk+1(Ωd)g\in H^{k+1}(\Omega_{d}). So by induction, gHn(Ωd)g\in H^{n}(\Omega_{d}) for any n>0n>0. In particular, gg is smooth and satisfies that lim|x|g(x,𝒚)=0\lim\limits_{|x|\to\infty}g(x,\bm{y})=0 holds uniformly in 𝒚\bm{y}. ∎

Remark 2.

Although we assume γC(Ωd)\gamma\in C^{\infty}(\Omega_{d}), for a given dimension dd, γCd+3(Ωd)\gamma\in C^{d+3}(\Omega_{d}) is sufficient to ensure that layer solutions uu and 0 eigenfunctions gg of operator LL are continuous and lim|x|g(x,𝐲)=0,limx±u(x,𝐲)=±1\lim\limits_{|x|\to\infty}g(x,\bm{y})=0,\lim\limits_{x\to\pm\infty}u(x,\bm{y})=\pm 1 hold uniformly in 𝐲\bm{y}. These are the properties we need to prove Theorem 2.

Finishing proving these two lemmas, we are ready to prove Theorem 2.

4.2. Proof of Theorem 2: the De Giorgi conjecture

As discussed in the beginning of this section, we will prove Theorem 2 by proving that the Ker(L)\mathrm{Ker}(L) is only 1 dimension. Here LL is the operator defined in (1.38). Similar to the proof in [18], Lemma 2.1, i.e. the maximal property plays a critical role in concluding linear dependence of uxu_{x} and any other function gg in the Ker(L)\mathrm{Ker}(L).

Proof of Theorem 2.

We will prove that if a non-trivial gH1(Ωd)g\in H^{1}(\Omega_{d}) satisfies g+γ′′(u)g=0\mathcal{L}g+\gamma^{\prime\prime}(u)g=0, then there exists a constant cc such that g=cux(x,𝒚)g=cu_{x}(x,\bm{y}).

According to Lemma 4.3, uH˙n(Ωd)u\in\dot{H}^{n}(\Omega_{d}) for any n>0n>0, so uu is continuous. By definition of layer solution (see Definition 3), we know that limx±u(x,𝒚)=±1\lim\limits_{x\to\pm\infty}u(x,\bm{y})=\pm 1.

This limit actually holds uniformly in 𝒚\bm{y} by continuity of uu. To prove uniformness, by strict monotonicity of uu, for any a(1,1)a\in(-1,1) and 𝒚𝕋d1\bm{y}\in\mathbb{T}^{d-1}, there exists a unique xx\in\mathbb{R} such that u(x,𝒚)=au(x,\bm{y})=a. Therefore, for any a(1,1)a\in(-1,1), we consider function

fa(𝒚):𝕋d1,fa(𝒚)={x:u(x,𝒚)=a}.\displaystyle f_{a}(\bm{y}):\mathbb{T}^{d-1}\to\mathbb{R},\ f_{a}(\bm{y})=\{x:u(x,\bm{y})=a\}.

We prove that fa(𝒚)f_{a}(\bm{y}) is continuous. Given 𝒚𝕋d1\bm{y}\in\mathbb{T}^{d-1} and ϵ>0\epsilon>0 sufficiently small, since u(fa(𝒚),𝒚)=au(f_{a}(\bm{y}),\bm{y})=a, by strict monotonicity of uu w.r.t. xx, we know that

a1:=u(fa(𝒚)+ϵ,𝒚)>a>a2:=u(fa(𝒚)ϵ,𝒚).\displaystyle a_{1}:=u(f_{a}(\bm{y})+\epsilon,\bm{y})>a>a_{2}:=u(f_{a}(\bm{y})-\epsilon,\bm{y}).

Then there exists δ>0\delta>0 such that

u(x,𝒚)>aif(x,𝒚)Sδ(fa(𝒚)+ϵ,𝒚),\displaystyle u(x,\bm{y})>a\ \mathrm{if}\ (x,\bm{y})\in S_{\delta}(f_{a}(\bm{y})+\epsilon,\bm{y}),
u(x,𝒚)<aif(x,𝒚)Sδ(fa(𝒚)ϵ,𝒚).\displaystyle u(x,\bm{y})<a\ \mathrm{if}\ (x,\bm{y})\in S_{\delta}(f_{a}(\bm{y})-\epsilon,\bm{y}).

Here Sδ(𝒘)S_{\delta}(\bm{w}) is the square centered at 𝒘\bm{w} with width δ\delta. Then by definition of faf_{a} and monotonicity of uu, we know that for any 𝒚1𝕋d1{\bm{y}_{1}}\in\mathbb{T}^{d-1} such that |𝒚1𝒚|<δ2|{\bm{y}_{1}-\bm{y}}|<\dfrac{\delta}{2}, we have |fa(𝒚1)fa(𝒚)|<ϵ|f_{a}(\bm{y}_{1})-f_{a}(\bm{y})|<\epsilon. Thus fa(𝒚):𝕋d1f_{a}(\bm{y}):\mathbb{T}^{d-1}\to\mathbb{R} is a continuous function for any a(1,1)a\in(-1,1). So by compactness of 𝕋d1\mathbb{T}^{d-1}, there exist real numbers xax_{a} and XaX_{a} such that

xa<fa(𝒚)<Xa.\displaystyle x_{a}<f_{a}(\bm{y})<X_{a}.

So by monotonicity, we know that for any 𝒚Ωd\bm{y}\in\Omega_{d},

u(x,𝒚)\displaystyle u(x,\bm{y}) aifxXa,\displaystyle\geq a\ \mathrm{if}\ x\geq X_{a},
u(x,𝒚)\displaystyle u(x,\bm{y}) aifxxa.\displaystyle\leq a\ \mathrm{if}\ x\leq x_{a}.

Thus limit limx±u(x,𝒚)=±1\lim\limits_{x\to\pm\infty}u(x,\bm{y})=\pm 1 holds uniformly in 𝒚\bm{y}.

By Lemma 4.4, we know that

limx±ux(x,𝒚)=0,limx±g(x,𝒚)=0\displaystyle\lim\limits_{x\to\pm\infty}u_{x}(x,\bm{y})=0,\ \lim\limits_{x\to\pm\infty}g(x,\bm{y})=0

hold uniformly in 𝒚\bm{y}. Consider ϕβ=ux+βg\phi_{\beta}=u_{x}+\beta g and define set

(4.3) D1:={β<0:ϕβ(𝝃)<0forsome𝝃Ωd}.\displaystyle D_{1}:=\{\beta<0:\ \phi_{\beta}(\bm{\xi})<0\ \mathrm{for\ some}\ \bm{\xi}\in\Omega_{d}\}.

Because gg is non-trivial, we assume that g(x0,𝒚0)>0g(x_{0},\bm{y}_{0})>0 for some (x0,𝒚0)Ωd(x_{0},\bm{y}_{0})\in\Omega_{d} without loss of generality. Then D1D_{1} is non-empty because

β1:=2ux(x0,𝒚0)/g(x0,𝒚0)D1.\displaystyle\beta_{1}:=-2u_{x}(x_{0},\bm{y}_{0})/g(x_{0},\bm{y}_{0})\in D_{1}.

Here we use the positivity of uxu_{x} in the definition of layer solutions. Therefore,

β¯:=supD1\displaystyle\overline{\beta}:=\sup D_{1}

is well-defined and satisfies β¯[β1,0]\overline{\beta}\in[\beta_{1},0].

We can also prove that for any βD1\beta\in D_{1}, there exists 𝝃βΩd\bm{\xi}_{\beta}\in\Omega_{d} such that ϕβ(𝝃β)\phi_{\beta}(\bm{\xi}_{\beta}) attains a negative minimum. By construction of D1D_{1} and Lemma 4.4, we know that

lim|x|ϕβ(x,𝒚)=0\displaystyle\lim\limits_{|x|\to\infty}\phi_{\beta}(x,\bm{y})=0

holds uniformly in 𝒚\bm{y}. Meanwhile, ϕβ\phi_{\beta} attains a negative minimum. Therefore, there exists 𝝃β=(xβ,𝒚β)\bm{\xi}_{\beta}=(x_{\beta},\bm{y}_{\beta}) such that ϕβ\phi_{\beta} attains minimum at 𝝃β\bm{\xi}_{\beta} by continuity of gg and uxu_{x}, which is ensured by Lemma 4.4 and Lemma 4.3.

Moreover, there exists X0X_{0}\in\mathbb{R} that only depends on γ\gamma and uu such that |xβ|X0|x_{\beta}|\leq X_{0} for any βD1\beta\in D_{1}. Notice that ϕβ\phi_{\beta} satisfies ϕβ+γ′′(u)ϕβ=0\mathcal{L}\phi_{\beta}+\gamma^{\prime\prime}(u)\phi_{\beta}=0 since both gg and uxu_{x} are so, thus

γ′′(u(xβ,𝒚β))ϕβ(xβ,𝒚β)=ϕβ|(xβ,𝒚β)>0\displaystyle\gamma^{\prime\prime}(u(x_{\beta},\bm{y}_{\beta}))\phi_{\beta}(x_{\beta},\bm{y}_{\beta})=-\mathcal{L}\phi_{\beta}|_{(x_{\beta},\bm{y}_{\beta})}>0

holds by minimality of ϕβ\phi_{\beta} and Lemma 2.1. Because ϕβ(xβ,𝒚β)<0\phi_{\beta}(x_{\beta},\bm{y}_{\beta})<0, so γ′′(u(xβ,𝒚β))<0\gamma^{\prime\prime}(u(x_{\beta},\bm{y}_{\beta}))<0. However, since limx±u(x,𝒚)=±1\lim\limits_{x\to\pm\infty}u(x,\bm{y})=\pm 1 uniformly in 𝒚\bm{y} and γ′′(±1)>0\gamma^{\prime\prime}(\pm 1)>0, so there exists a constant X0>0X_{0}>0 such that if |x|X0|x|\geq X_{0}, then γ′′(u(x,𝒚))0\gamma^{\prime\prime}(u(x,\bm{y}))\geq 0. Because γ′′(u(xβ,𝒚β))<0\gamma^{\prime\prime}(u(x_{\beta},\bm{y}_{\beta}))<0, so |xβ|<X0|x_{\beta}|<X_{0}.

Therefore, we know that {𝝃β}βD1\{\bm{\xi}_{\beta}\}_{\beta\in D_{1}} is a compact set in Ωd\Omega_{d}. So there exists a subsequence of β\beta in D1D_{1} such that ββ¯\beta\to\overline{\beta}, i.e. the supremum of set D1D_{1}, and

𝝃β𝝃¯β\displaystyle\bm{\xi}_{\beta}\to\overline{\bm{\xi}}_{\beta}

for some 𝝃¯βΩd\overline{\bm{\xi}}_{\beta}\in\Omega_{d} . Because ϕβ(𝝃β)<0\phi_{\beta}(\bm{\xi}_{\beta})<0, so

ϕβ¯(𝝃¯β)0\displaystyle\phi_{\overline{\beta}}(\overline{\bm{\xi}}_{\beta})\leq 0

by passing the limit ββ¯\beta\to\overline{\beta} and continuity of gg and uxu_{x}. However, by the definition of β¯\overline{\beta}, we have ϕβ¯(𝝃)0\phi_{\overline{\beta}}(\bm{\xi})\geq 0 for any 𝝃Ωd\bm{\xi}\in\Omega_{d} otherwise β¯\overline{\beta} should not be the supremum of D1D_{1}. Thus ϕβ¯(𝝃¯β)=0\phi_{\overline{\beta}}(\overline{\bm{\xi}}_{\beta})=0.

This ensures ϕβ¯0\phi_{\overline{\beta}}\equiv 0. Because ϕβ¯0\phi_{\overline{\beta}}\geq 0, so ϕβ¯\phi_{\overline{\beta}} attains minimum at 𝝃¯β\overline{\bm{\xi}}_{\beta}. However, since ϕβ¯\phi_{\overline{\beta}} is also in the kernel of LL, we have

ϕβ¯|𝝃¯β=γ′′(u(𝝃¯β))ϕβ¯(𝝃¯β)=0.\displaystyle\mathcal{L}\phi_{\overline{\beta}}|_{\overline{\bm{\xi}}_{\beta}}=-\gamma^{\prime\prime}(u(\overline{\bm{\xi}}_{\beta}))\phi_{\overline{\beta}}(\overline{\bm{\xi}}_{\beta})=0.

Then by Lemma 2.1 and minimality of 𝝃¯β\overline{\bm{\xi}}_{\beta}, we have ϕβ¯0\phi_{\overline{\beta}}\equiv 0. Thus

ux+β¯g=0,\displaystyle u_{x}+\overline{\beta}g=0,

i.e. gg and uxu_{x} are linearly dependent. Thus the kernel of LL is only 1 dimension. Notice that every partial derivative of uu belongs to kernel of LL, so there exist constants ci(i=1,2,,d1)c_{i}\ (i=1,2,...,d-1) such that

uyi+ciux=0,i=1,2,,d1\displaystyle u_{y_{i}}+c_{i}u_{x}=0,\ i=1,2,...,d-1

for any yi𝕋d1y_{i}\in\mathbb{T}^{d-1}.

To close the proof, we prove that in fact ci=0,i=1,2,,d1c_{i}=0,i=1,2,...,d-1. Otherwise, we assume ci>0c_{i}>0 without loss of generality. For any given (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}, by periodicity and the far end limit assumption (1.9), we have

u(x,𝒚)\displaystyle u(x,\bm{y}) =limn+u(x+cin,𝒚n𝒆i)=limn+u(x+cin,𝒚)=1,\displaystyle=\lim\limits_{n\to+\infty}u(x+c_{i}n,\bm{y}-n\bm{e}_{i})=\lim\limits_{n\to+\infty}u(x+c_{i}n,\bm{y})=1,
u(x,𝒚)\displaystyle u(x,\bm{y}) =limn+u(xcin,𝒚+n𝒆i)=limn+u(xcin,𝒚)=1.\displaystyle=\lim\limits_{n\to+\infty}u(x-c_{i}n,\bm{y}+n\bm{e}_{i})=\lim\limits_{n\to+\infty}u(x-c_{i}n,\bm{y})=-1.

Here nn are positive integers and 𝒆i=(0,,0,1,0,0),i=1,2,,d1\bm{e}_{i}=(0,...,0,1,0,...0),i=1,2,...,d-1 form the canonical orthogonal basis in d1\mathbb{R}^{d-1} with 1 only at the ii th component, and 0 for others. This yields contradiction. So ci=0,i=1,2,,d1c_{i}=0,i=1,2,...,d-1, i.e. 𝒚u(x,𝒚)=0\nabla_{\bm{y}}u(x,\bm{y})=0 and uu is a 1D profile that only depends on xx. ∎

4.3. Proof of Theorem 3: uniqueness up to translations

To completely understand all layer solutions to (1.23) and minimizers of functional FF on set 𝒜\mathcal{A}, we prove the following lemma:

Lemma 4.5.

(minimizers are layer solutions) For any dimension d1d\geq 1, suppose that γC()\gamma\in C^{\infty}(\mathbb{R}) is a double-well type potential satisfying (1.8). Consider functional energy FF in (1.37), set 𝒜\mathcal{A} in (1.35), and set 𝒜\mathcal{A}_{\ell}, 𝒜m\mathcal{A}_{m} in (1.39). Then

𝒜m𝒜.\displaystyle\mathcal{A}_{m}\subset\mathcal{A}_{\ell}.
Proof.

Let u𝒜mu^{*}\in\mathcal{A}_{m}. First of all, uu^{*} is a weak solution to equation (1.23). Then as in the proof of Theorem 1, we know that it solves (1.23) L2L^{2} sense, i.e. u=γ(u)L2(Ωd)\mathcal{L}u^{*}=-\gamma^{\prime}(u^{*})\in L^{2}(\Omega_{d}) (see calculation (3.18)). Then by Lemma 2.3, we know that uηH1(Ωd)u^{*}-\eta\in H^{1}(\Omega_{d}).

Because uu^{*} is a minimizer, so |u|1|u^{*}|\leq 1. Otherwise

(4.4) u~=max{1,min{1,u}}\displaystyle\tilde{u}=\max\{1,\min\{-1,u^{*}\}\}

is also in 𝒜\mathcal{A} and satisfies F(u~)<F(u)F(\tilde{u})<F(u^{*}) by definition of FF in (1.37). This contradicts with the minimality of uu^{*}. So uu^{*} is bounded. Then by Lemma 4.3, uηHn(Ωd)u^{*}-\eta\in H^{n}(\Omega_{d}) for any n>0n>0. Therefore, we have

(4.5) limx±u(x,𝒚)=±1.\displaystyle\lim\limits_{x\to\pm\infty}u^{*}(x,\bm{y})=\pm 1.

Now it is left to prove the strict monotonicity of uu^{*}. Again, this is realized by the energy decreasing rearrangement method (Lemma 3.2). For any τ>0\tau>0, consider the translation of uu^{*}, i.e.

uτ(x,𝒚)=u(x+τ,𝒚).\displaystyle u_{\tau}(x,\bm{y})=u^{*}(x+\tau,\bm{y}).

Define

m(𝒘):=min{uτ(𝒘),u(𝒘)},M(𝒘):=max{uτ(𝒘),u(𝒘)}.\displaystyle m(\bm{w}):=\min\{u_{\tau}(\bm{w}),\ u^{*}(\bm{w})\},\ M(\bm{w}):=\max\{u_{\tau}(\bm{w}),\ u^{*}(\bm{w})\}.

Then by Lemma 3.2, we know that

F(m)+F(M)F(uτ)+F(u).\displaystyle F(m)+F(M)\leq F(u_{\tau})+F(u^{*}).

By translation-invariance (Lemma 3.3), we know F(uτ)=F(u)F(u_{\tau})=F(u^{*}). Thus both uu^{*} and uτu_{\tau} are minimizers. So by minimality of uu^{*} and uτu_{\tau}, we have

F(m)+F(M)=F(uτ)+F(u).\displaystyle F(m)+F(M)=F(u_{\tau})+F(u^{*}).

Again, by Lemma 3.2, this equality holds if and only if either uτ(𝒘)u(𝒘)u_{\tau}(\bm{w})\geq u^{*}(\bm{w}) or uτ(𝒘)u(𝒘)u_{\tau}(\bm{w})\leq u^{*}(\bm{w}). Then by the limit condition (4.5), we know uτ(𝒘)u(𝒘)u_{\tau}(\bm{w})\geq u^{*}(\bm{w}). Thus uu^{*} is non-decreasing.

Finally, as in the proof of Theorem 1, the fact that uu is non-decreasing implies strict monotonicity. Suppose that u(x0,𝒚0)x=0\dfrac{\partial u^{*}(x_{0},\bm{y}_{0})}{\partial x}=0 for some (x0,𝒚0)Ωd(x_{0},\bm{y}_{0})\in\Omega_{d}, then taking derivative on both sides of (1.23) yields

u(x0,𝒚0)x=γ′′(u)u(x0,𝒚0)x=0.\displaystyle\mathcal{L}\dfrac{\partial u^{*}(x_{0},\bm{y}_{0})}{\partial x}=-\gamma^{\prime\prime}(u^{*})\dfrac{\partial u^{*}(x_{0},\bm{y}_{0})}{\partial x}=0.

Thus ux=0\mathcal{L}\dfrac{\partial u^{*}}{\partial x}=0 at (x0,𝒚0)(x_{0},\bm{y}_{0}). However, since ux0\dfrac{\partial u^{*}}{\partial x}\geq 0, we know that ux\dfrac{\partial u^{*}}{\partial x} attains minimum at (x0,𝒚0)(x_{0},\bm{y}_{0}). Then by Lemma 2.1, we know that ux=0\dfrac{\partial u^{*}}{\partial x}=0, i.e. uu^{*} is a constant. This contradicts with the far field limit of uu^{*}. So u(x,𝒚)x>0\dfrac{\partial u^{*}(x,\bm{y})}{\partial x}>0 holds for any (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}. Thus uu^{*} is a layer solution. ∎

Therefore, all minimizers of FF on set 𝒜\mathcal{A} are layer solutions. Recall that Theorem 2 claims that all layer solutions with H1H^{1} regularity have one-dimensional symmetry if the double-well potential γ\gamma is smooth, so all these minimizers are also exactly 1D profiles.

Moreover, these 1D profiles are unique up to translations. According to [9], if γC2,α()\gamma\in C^{2,\alpha}(\mathbb{R}) is a double-well potential, then layer solutions to

(4.6) (xx)1/2u(x)+γ(u(x))=0,x.\displaystyle(-\partial_{xx})^{1/2}u(x)+\gamma^{\prime}(u(x))=0,\ x\in\mathbb{R}.

is unique up to translations (see Theorem 1.2 in [9]). Remember that Lemma 2.2 ensures that f=c(xx)1/2f\mathcal{L}f=c_{\mathcal{L}}(-\partial_{xx})^{1/2}f if f(x,𝒚)=f(x)f(x,\bm{y})=f(x) is a 1D profile, therefore, both layer solutions and minimizers are unique up to translations.

Proof of Theorem 3.

By Theorem 2, we know that for any u𝒜u\in\mathcal{A}_{\ell}, u is a 1D profile to solution (1.23), i.e.

u(x,𝒚)+γ(u(x,𝒚))=0.\displaystyle\mathcal{L}u(x,\bm{y})+\gamma^{\prime}(u(x,\bm{y}))=0.

By Lemma 2.2, for any (x,𝒚)Ωd(x,\bm{y})\in\Omega_{d}, we have

u(x,𝒚)=c(xx)1/2u(x).\displaystyle\mathcal{L}u(x,\bm{y})=c_{\mathcal{L}}(-\partial_{xx})^{1/2}u(x).

Thus viewed as a 1D profile u(x)u(x), a layer solution u(x,𝒚)u(x,\bm{y}) satisfies

(4.7) c(xx)1/2u(x)+γ(u(x))=0.\displaystyle c_{\mathcal{L}}(-\partial_{xx})^{1/2}u(x)+\gamma^{\prime}(u(x))=0.

Then by Theorem 1.2 in [9], we know

(4.8) 𝒜={u:u(x,𝒚)=u(x+x0)forsomex0}.\displaystyle\mathcal{A}_{\ell}=\{u:\ u(x,\bm{y})=u^{*}(x+x_{0})\ \mathrm{for\ some}\ x_{0}\in\mathbb{R}\}.

Here uu^{*} is the unique solution to (1.41).

By Theorem 1 and Lemma 4.5, we know that 𝒜m\mathcal{A}_{m} is non-empty and 𝒜m𝒜\mathcal{A}_{m}\subset\mathcal{A}_{\ell}. Moreover, by Lemma 3.3, i.e. the translation-invariant property, we know that u(x)𝒜mu(x)\in\mathcal{A}_{m} if and only if u(x+x0)𝒜mu(x+x_{0})\in\mathcal{A}_{m}. Notice that 𝒜\mathcal{A}_{\ell} itself is also unique up to translations, so we have (1.40), i.e.

𝒜m=𝒜={u:u(x,𝒚)=u(x+x0)forsomex0}.\displaystyle\mathcal{A}_{m}=\mathcal{A}_{\ell}=\{u:\ u(x,\bm{y})=u^{*}(x+x_{0})\ \mathrm{for\ some}\ x_{0}\in\mathbb{R}\}.

This concludes the uniqueness (up to translations) of layer solutions to equation (1.23) and minimizers of FF on set 𝒜\mathcal{A}. ∎

4.4. Proof of Theorem 4: implication on the PN model

As a direct application of previous results on the existence and rigidity, now we can prove Theorem 4.

Proof of Theorem 4.

As a minimizer of E~\tilde{E} in (1.42) in the perturbed sense, we know that 𝒖\bm{u} is a weak solution to (1.13) by Lemma 1.1. A calculation (see [10]) involving the Dirichlet and Neumann map implies that if 𝒖\bm{u} satisfies (1.13), the elastic energy in the bulk can be expressed by u1+(x,z)u_{1}^{+}(x,z) which is defined on the slip plane:

(4.9) Eels(𝒖)=Γu1+(𝒘)u1+(𝒘)d𝒘=12Ω2Ω2|u1+(𝒘)u1+(𝒘)|2K(𝒘𝒘)d𝒘d𝒘.\displaystyle E_{\mathrm{els}}(\bm{u})=\int_{\Gamma^{\prime}}\mathcal{L}u_{1}^{+}(\bm{w})u_{1}^{+}(\bm{w})\mathrm{d}\bm{w}=\dfrac{1}{2}\int_{\Omega_{2}}\int_{\Omega_{2}}|u_{1}^{+}(\bm{w})-u_{1}^{+}(\bm{w}^{\prime})|^{2}K(\bm{w}-\bm{w}^{\prime})\mathrm{d}\bm{w}\mathrm{d}\bm{w}^{\prime}.

Here \mathcal{L} is the linear operator defined in (1.15) and KK is the corresponding convolution kernel which satisfies Assumption (A)-(D). Therefore, u1+u_{1}^{+} is the minimizer of FF defined in (1.37). Then by Theorem 2 and Theorem 3, we know that statement (ii) hold. Therefore, u1,u2u_{1},u_{2} only depend on xx and yy, satisfying the following reduced system of (1.13) in two dimensions:

(4.10) {Δ𝒖+112ν(𝒖)=0,in2Γ1,σ12++σ12=γu1(u1+),onΓ1,σ22+=σ22,onΓ1.\displaystyle\begin{cases}\Delta\bm{u}+\dfrac{1}{1-2\nu}\nabla(\nabla\cdot\bm{u})=0,&\ \mathrm{in\ }\mathbb{R}^{2}\setminus\Gamma_{1},\\ \sigma^{+}_{12}+\sigma^{-}_{12}=\dfrac{\partial\gamma}{\partial u_{1}}(u_{1}^{+}),&\ \mathrm{on\ }\Gamma_{1},\\ \sigma^{+}_{22}=\sigma^{-}_{22},&\ \mathrm{on\ }\Gamma_{1}.\end{cases}

Here Γ1={(x,y)2:y=0}\Gamma_{1}=\{(x,y)\in\mathbb{R}^{2}:\ y=0\}. Thus smoothness of u1+u_{1}^{+} implies that 𝒖\bm{u} is smooth in 2×𝕋Γ\mathbb{R}^{2}\times\mathbb{T}\setminus\Gamma^{\prime}, so (i) holds. Finally, by Lemma 2.3 in [19], we know that (iii), (iv) and (v) are true and hold point-wisely in 2×𝕋Γ\mathbb{R}^{2}\times\mathbb{T}\setminus\Gamma^{\prime} by smoothness. ∎

5. Spectral analysis of LL

In Theorem 2, we prove that if uu is a layer solution to (1.23), then the operator in (1.38), i.e.

L:H1(Ω)L2(Ω)L2(Ω),Lϕ=ϕ+γ′′(u)ϕ\displaystyle L:H^{1}(\Omega)\subset L^{2}(\Omega)\to L^{2}(\Omega),\ L\phi=\mathcal{L}\phi+\gamma^{\prime\prime}(u)\phi

has one dimensional kernel which is exactly span{ux}\mathrm{span}\{u_{x}\}. In this section, we proceed to prove that LL is positively semi-definite and 0 is an isolated point spectrum. Denote the spectrum, the point spectrum, the residual spectrum and the continuous spectrum of a linear operator LL as σ(L),σp(L),σr(L)\sigma(L),\sigma_{p}(L),\sigma_{r}(L) and σc(L)\sigma_{c}(L) respectively.

First of all, according to [41], since LL is self-adjoint (see Lemma A.2), we know σr(L)=\sigma_{r}(L)=\emptyset. Meanwhile, since

limx±u(x,𝒚)=±1\displaystyle\lim\limits_{x\to\pm\infty}u(x,\bm{y})=\pm 1

holds uniformly in the 𝒚\bm{y} direction and γ′′(±1)>0\gamma^{\prime\prime}(\pm 1)>0, we know that γ′′(u)\gamma^{\prime\prime}(u) is lower bounded and can only be negative on a compact set in Ωd\Omega_{d}. Therefore, there exists a finite lower bound of the spectrum of LL, i.e.

Lemma 5.1.

σ(L)=σp(L)σc(L)[λ1,)\sigma(L)=\sigma_{p}(L)\cup\sigma_{c}(L)\subset\left[-\lambda_{1},\infty\right). Here λ1>0\lambda_{1}>0 is a constant.

Employing the perturbation theory of self-adjoint operators [25], we can characterize the essential spectrum of LL by viewing LL as a self-adjoint perturbation of \mathcal{L}. Remember that the continuous spectrum is a subset of the essential spectrum, we have the following lemma:

Lemma 5.2.

σc(L)[λ2,)\sigma_{c}(L)\subset\left[\lambda_{2},\infty\right). Here λ2>0\lambda_{2}>0 is a constant.

Therefore, the spectrum of LL that belongs to (λ1,λ2)(-\lambda_{1},\lambda_{2}) is a subset of σp(L)\sigma_{p}(L) with finite dimensional eigenspaces. Moreover, they are isolated points in σ(L)\sigma(L). To finish the spectral analysis of L2L_{2}, we finally prove the positive semi-definiteness of LL.

Lemma 5.3.

σp(L)[0,)\sigma_{p}(L)\subset[0,\infty).

We will only prove Lemma 5.3 in this section and the proof of the fact that LL is self-adjoint, Lemma 5.1 and Lemma 5.2 is attached in Appendix A. Similar to the proof of Theorem 2, the proof of Lemma 5.3 adopts an argument of contradiction and relies on the maximal principle of \mathcal{L} in Lemma 2.1.

Proof of Lemma 5.3.

We will prove σp(L)[0,)\sigma_{p}(L)\subset[0,\infty) by contradiction. Suppose that there exist λ<0\lambda<0 and non-zero gH1(Ωd)g\in H^{1}(\Omega_{d}) s.t.

Lg=λg.\displaystyle Lg=\lambda g.

similar to the proof of Lemma 4.4, we can prove that gHn(Ωd)g\in H^{n}(\Omega_{d}) for any n>0n>0 and

lim|x|+g(x,𝒚)=0\displaystyle\lim\limits_{|x|\to+\infty}g(x,\bm{y})=0

holds uniformly in 𝒚\bm{y} direction.

Consider L|g|L|g|. By Assumption B, i.e. positivity of kernel KK, and the fact that for any 𝒘,𝒘Ωd\bm{w},\bm{w}^{\prime}\in\Omega_{d},

|g(𝒘)|sgn(g)(𝒘)g(𝒘),\displaystyle|g(\bm{w}^{\prime})|\geq\mathrm{sgn}(g)(\bm{w})g(\bm{w}^{\prime}),

we have

L|g|(x,𝒚)\displaystyle L|g|(x,\bm{y}) =|g|(x,𝒚)+γ′′(u)|g|(x,𝒚)\displaystyle=\mathcal{L}|g|(x,\bm{y})+\gamma^{\prime\prime}(u)|g|(x,\bm{y})
=Ωd(|g(x,𝒚)||g(x,𝒚)|)K(xx,𝒚𝒚)dxd𝒚+γ′′(u)|g|(x,𝒚)\displaystyle=\int_{\Omega_{d}}(|g(x,\bm{y})|-|g(x^{\prime},\bm{y}^{\prime})|)K(x-x,\bm{y}-\bm{y}^{\prime})\mathrm{d}x^{\prime}\mathrm{d}\bm{y}^{\prime}+\gamma^{\prime\prime}(u)|g|(x,\bm{y})
=Ωd(sgn(g)g(x,𝒚)|g(x,𝒚)|)K(xx,𝒚𝒚)dxd𝒚+sgn(g)(x,𝒚)γ′′(u)g(x,𝒚)\displaystyle=\int_{\Omega_{d}}(\mathrm{sgn}(g)g(x,\bm{y})-|g(x^{\prime},\bm{y}^{\prime})|)K(x-x,\bm{y}-\bm{y}^{\prime})\mathrm{d}x^{\prime}\mathrm{d}\bm{y}^{\prime}+\mathrm{sgn}(g)(x,\bm{y})\gamma^{\prime\prime}(u)g(x,\bm{y})
sgn(g)(x,𝒚)[Ωd(g(x,𝒚)g(x,𝒚))K(xx,𝒚𝒚)dxd𝒚+γ′′(u)g(x,𝒚)]\displaystyle\leq\mathrm{sgn}(g)(x,\bm{y})\left[\int_{\Omega_{d}}(g(x,\bm{y})-g(x^{\prime},\bm{y}^{\prime}))K(x-x,\bm{y}-\bm{y}^{\prime})\mathrm{d}x^{\prime}\mathrm{d}\bm{y}^{\prime}+\gamma^{\prime\prime}(u)g(x,\bm{y})\right]
sgn(g)(x,𝒚)Lg(x,𝒚)\displaystyle\leq\mathrm{sgn}(g)(x,\bm{y})\cdot Lg(x,\bm{y})
=λ|g|(x,𝒚)0.\displaystyle=\lambda|g|(x,\bm{y})\leq 0.

Thus |g||g| satisfies

(5.1) L|g|λ|g|0.\displaystyle L|g|\leq\lambda|g|\leq 0.

Define ϕβ=ux+β|g|\phi_{\beta}=u_{x}+\beta|g| for real number β\beta. Consider the following set of β\beta:

D:={β<0|ϕβ(𝝃)<0forsome𝝃Ωd}.\displaystyle D:=\{\beta<0\ |\ \phi_{\beta}(\bm{\xi})<0\ \mathrm{for\ some\ }\bm{\xi}\in\Omega_{d}\}.

DD is nonempty because

β1=2ux(x0,𝒚0)|g|(x0,𝒚0)D\displaystyle\beta_{1}=\dfrac{-2u_{x}(x_{0},\bm{y}_{0})}{|g|(x_{0},\bm{y}_{0})}\in D

for (x0,𝒚0)(x_{0},\bm{y}_{0}) satisfying |g|(x0,𝒚0)>0|g|(x_{0},\bm{y}_{0})>0. Therefore

β¯:=supD\displaystyle\overline{\beta}:=\mathrm{sup}D

is a well-defined finite number that lies in [β1,0][\beta_{1},0].

Now for any βD\beta\in D, we will prove that there exists (xβ,𝒚β)Ωd(x_{\beta},\bm{y}_{\beta})\in\Omega_{d} such that ϕβ(xβ,𝒚β)\phi_{\beta}(x_{\beta},\bm{y}_{\beta}) attains a negative minimum at 𝝃β=(xβ,𝒚β)\bm{\xi}_{\beta}=(x_{\beta},\bm{y}_{\beta}). First of all, by the definition of DD, we know that ϕβ\phi_{\beta} is non-zero and attains a negative infimum. Remember that

lim|x|g(x,𝒚)=0,lim|x|ux(x,𝒚)=0\displaystyle\lim\limits_{|x|\to\infty}g(x,\bm{y})=0,\quad\lim\limits_{|x|\to\infty}u_{x}(x,\bm{y})=0

holds uniformly in 𝒚\bm{y}, so

lim|x|ϕβ(x,𝒚)=0\displaystyle\lim\limits_{|x|\to\infty}\phi_{\beta}(x,\bm{y})=0

holds uniformly in 𝒚\bm{y}. Recall that ϕβ\phi_{\beta} attains a negative infimum, so continuity of ϕβ\phi_{\beta} implies that this infimum is indeed a minimum that is attained for some (xβ,𝒚β)(x_{\beta},\bm{y}_{\beta}).

Moreover, {𝝃β}β\{\bm{\xi}_{\beta}\}_{\beta} is bounded in Ωd\Omega_{d}. Notice that by (5.1) and β<0\beta<0,

Lϕβ\displaystyle L\phi_{\beta} =Lux+βL|g|βλ|g|0.\displaystyle=Lu_{x}+\beta L|g|\geq\beta\lambda|g|\geq 0.

Thus Lϕβ|(x,𝒚)=𝝃β0L\phi_{\beta}|_{(x,\bm{y})=\bm{\xi}_{\beta}}\geq 0. By maximal principle (Lemma 2.1), we know that

ϕβ|(x,𝒚)=𝝃β0\displaystyle\mathcal{L}\phi_{\beta}|_{(x,\bm{y})=\bm{\xi}_{\beta}}\leq 0

since ϕβ\phi_{\beta} attains minimum at 𝝃β\bm{\xi}_{\beta}. Therefore, we have

γ′′(u(xβ,𝒚β))ϕβ(xβ,𝒚β)=Lϕβ|(x,𝒚)=𝝃βϕβ|(x,𝒚)=𝝃β0.\displaystyle\gamma^{\prime\prime}(u(x_{\beta},\bm{y}_{\beta}))\phi_{\beta}(x_{\beta},\bm{y}_{\beta})=L\phi_{\beta}|_{(x,\bm{y})=\bm{\xi}_{\beta}}-\mathcal{L}\phi_{\beta}|_{(x,\bm{y})=\bm{\xi}_{\beta}}\geq 0.

Because ϕβ(xβ,𝒚β)0\phi_{\beta}(x_{\beta},\bm{y}_{\beta})\leq 0 by definition of (xβ,𝒚β)(x_{\beta},\bm{y}_{\beta}), so γ′′(u(xβ,𝒚β))0\gamma^{\prime\prime}(u(x_{\beta},\bm{y}_{\beta}))\geq 0. So there exists X>0X>0 that only depends on uu and γ\gamma such that |xβ|X|x_{\beta}|\leq X. Since gg is periodic in 𝒚\bm{y}, we know that {𝝃β}β\{\bm{\xi}_{\beta}\}_{\beta} is bounded in Ωd\Omega_{d}.

Given the boundedness of sequence {𝝃β}β\{\bm{\xi}_{\beta}\}_{\beta}, we can now take a subsequence of β\beta (still denoted as β\beta) such that ββ¯\beta\to\overline{\beta}, the supremum of DD, and 𝝃βξ\bm{\xi}_{\beta}\to\xi^{*} as ββ¯\beta\to\overline{\beta}. As the supremum of DD, β¯\overline{\beta} satisfies that ϕβ¯(x,𝒚)0\phi_{\overline{\beta}}(x,\bm{y})\geq 0. However, since ϕβ(𝝃β)0\phi_{\beta}(\bm{\xi}_{\beta})\leq 0, passing the limit in β\beta gives that

ϕβ¯(𝝃)=limββ¯ϕβ(𝝃β)0.\displaystyle\phi_{\overline{\beta}}(\bm{\xi}^{*})=\lim\limits_{\beta\to\overline{\beta}}\phi_{\beta}(\bm{\xi}_{\beta})\leq 0.

So ϕβ¯(𝝃)=0\phi_{\overline{\beta}}(\bm{\xi}^{*})=0, which means that ϕβ¯\phi_{\overline{\beta}} attains minimum 0 at 𝝃\bm{\xi}^{*}.

This in fact ensures that ϕβ¯0.\phi_{\overline{\beta}}\equiv 0. By Lemma 2.1, we know

ϕβ¯|(x,𝒚)=𝝃0.\displaystyle\mathcal{L}\phi_{\overline{\beta}}|_{(x,\bm{y})=\bm{\xi}^{*}}\leq 0.

However, we also have

Lϕβ¯=β¯L|g|β¯λ|g|0.\displaystyle L\phi_{\overline{\beta}}=\overline{\beta}L|g|\geq\overline{\beta}\lambda|g|\geq 0.

Remember ϕβ¯(𝝃)=0\phi_{\overline{\beta}}(\bm{\xi}^{*})=0, so

0\displaystyle 0 Lϕβ¯(𝝃)\displaystyle\leq L\phi_{\overline{\beta}}(\bm{\xi}^{*})
=ϕβ¯|(x,𝒚)=𝝃+γ′′(u(x,𝒚))ϕβ¯(x,𝒚)|(x,𝒚)=𝝃\displaystyle=\mathcal{L}\phi_{\overline{\beta}}|_{(x,\bm{y})=\bm{\xi}^{*}}+\gamma^{\prime\prime}(u(x,\bm{y}))\phi_{\overline{\beta}}(x,\bm{y})|_{(x,\bm{y})=\bm{\xi}^{*}}
=ϕβ¯|(x,𝒚)=𝝃\displaystyle=\mathcal{L}\phi_{\overline{\beta}}|_{(x,\bm{y})=\bm{\xi}^{*}}
0.\displaystyle\leq 0.

So all these inequalities are in fact equalities, i.e. ϕβ¯|(x,𝒚)=𝝃=0\mathcal{L}\phi_{\overline{\beta}}|_{(x,\bm{y})=\bm{\xi}^{*}}=0. By Lemma 2.1, we know that ϕβ¯0\phi_{\overline{\beta}}\equiv 0, which gives L|g|=0L|g|=0 and 0L|g|λ|g|0,0\leq L|g|\leq\lambda|g|\leq 0, hence |g|=0|g|=0, contradiction! Thus LL has no negative point spectrum. ∎

Acknowledgement

Jian-Guo Liu was supported in part by the National Science Foundation (NSF) under award DMS-1812573 and DMS-2106988.

References

  • [1] Robert A Adams and John JF Fournier. Sobolev spaces. Elsevier, 2003.
  • [2] Giovanni Alberti. Some remarks about a notion of rearrangement. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 29(2):457–472, 2000.
  • [3] Giovanni Alberti and Giovanni Bellettini. A nonlocal anisotropic model for phase transitions. Mathematische Annalen, 310(3):527–560, 1998.
  • [4] Giovanni Alberti, Guy Bouchitté, and Pierre Seppecher. Phase transition with the line-tension effect. Archive for Rational Mechanics and Analysis, 144(1):1–46, Nov 1998.
  • [5] Samuel Miller Allen and John W Cahn. Ground state structures in ordered binary alloys with second neighbor interactions. Acta Metallurgica, 20(3):423–433, 1972.
  • [6] Peter M Anderson, John P Hirth, and Jens Lothe. Theory of dislocations. Cambridge University Press, 2017.
  • [7] T Brooke Benjamin. Internal waves of permanent form in fluids of great depth. Journal of Fluid Mechanics, 29(3):559–592, 1967.
  • [8] Timothy Blass, Irene Fonseca, Giovanni Leoni, and Marco Morandotti. Dynamics for systems of screw dislocations. SIAM Journal on Applied Mathematics, 75(2):393–419, 2015.
  • [9] Xavier Cabré and Joan Solà-Morales. Layer solutions in a half-space for boundary reactions. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 58(12):1678–1732, 2005.
  • [10] Luis Caffarelli and Luis Silvestre. An extension problem related to the fractional laplacian. Communications in Partial Differential Equations, 32(7-9), 2006.
  • [11] Edward Norman Dancer. New solutions of equations on rnr^{n}. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 30(3-4):535–563, 2001.
  • [12] Ennio De Giorgi. Convergence problems for functionals and operators. Proc.int.meeting on Recent Methods in Nonlinear Analysis, pages 131–188, 1978.
  • [13] Manuel Del Pino, Michal Kowalczyk, and Juncheng Wei. On De Giorgi’s conjecture in dimension n \geq 9. Annals of Mathematics, pages 1485–1569, 2011.
  • [14] Hongjie Dong and Yuan Gao. Existence and uniqueness of bounded stable solutions to the peierls–nabarro model for curved dislocations. Calculus of Variations and Partial Differential Equations, 60(2):1–26, 2021.
  • [15] Alessio Figalli and Joaquim Serra. On stable solutions for boundary reactions: a De Giorgi-type result in dimension 4+ 1. Inventiones mathematicae, 219(1):153–177, 2020.
  • [16] Irene Fonseca, Janusz Ginster, and Stephan Wojtowytsch. On the motion of curved dislocations in three dimensions: Simplified linearized elasticity. SIAM Journal on Mathematical Analysis, 53(2):2373–2426, 2021.
  • [17] JA Frenkel. Zur theorie der elastizitätsgrenze und der festigkeit kristallinischer körper. Zeitschrift für Physik, 37(7-8):572–609, 1926.
  • [18] Yuan Gao and Jian-Guo Liu. Long time behavior of dynamic solution to Peierls–Nabarro dislocation model. Methods and Applications of Analysis, 27:7–51, 2020.
  • [19] Yuan Gao, Jian-Guo Liu, Tao Luo, and Yang Xiang. Mathematical validation of the Peierls–Nabarro model for edge dislocations. DCDS, series B, 26:3177–3207, 2021.
  • [20] A. Garroni and S. Müller. γ\gamma-limit of a phase-field model of dislocations. SIAM Journal on Mathematical Analysis, 36(6):1943–1964, Jan 2005.
  • [21] Nassif Ghoussoub and Changfeng Gui. On a conjecture of De Giorgi and some related problems. Mathematische Annalen, 311(3):481–491, 1998.
  • [22] Radu Ignat and Antonin Monteil. A De Giorgi–type conjecture for minimal solutions to a nonlinear Stokes equation. Communications on Pure and Applied Mathematics, 73(4):771–854, 2020.
  • [23] Cyril Imbert and Panagiotis E. Souganidis. Phasefield theory for fractional diffusion-reaction equations and applications. arXiv:0907.5524, Jul 2009.
  • [24] Tianpeng Jiang, Yang Xiang, and Luchan Zhang. Stochastic peierls–nabarro model for dislocations in high entropy alloys. SIAM Journal on Applied Mathematics, 80(6):2496–2517, 2020.
  • [25] Tosio Kato. Perturbation theory for linear operators, volume 132. Springer Science & Business Media, 2013.
  • [26] Mateusz Kwaśnicki. Ten equivalent definitions of the fractional Laplace operator. Fractional Calculus and Applied Analysis, 20(1):7–51, 2017.
  • [27] Gang Lu. The Peierls—Nabarro model of dislocations: a venerable theory and its current development. In Handbook of materials modeling, pages 793–811. Springer, 2005.
  • [28] Gang Lu, Nicholas Kioussis, Vasily V Bulatov, and Efthimios Kaxiras. Generalized-stacking-fault energy surface and dislocation properties of aluminum. Physical Review B, 62(5):3099, 2000.
  • [29] E Taylor Michael and ME Taylor. Partial differential equations. iii. Applied Mathematical Sciences, 117, 1999.
  • [30] Vincent Millot, Yannick Sire, and Kelei Wang. Asymptotics for the fractional Allen–Cahn equation and stationary nonlocal minimal surfaces. Archive for Rational Mechanics and Analysis, 231(2):1129–1216, Feb 2019.
  • [31] Louis Nirenberg. On elliptic partial differential equations. In Il principio di minimo e sue applicazioni alle equazioni funzionali, pages 1–48. Springer, 2011.
  • [32] Giampiero Palatucci, Ovidiu Savin, and Enrico Valdinoci. Local and global minimizers for a variational energy involving a fractional norm. Annali di matematica pura ed applicata, 192(4):673–718, 2013.
  • [33] Frédéric Riesz. Sur une inégalité intégarale. Journal of the London Mathematical Society, 1(3):162–168, 1930.
  • [34] Ovidiu Savin. Rigidity of minimizers in nonlocal phase transitions. Anal. PDE, 11(8):1881–1900, 2018.
  • [35] Ovidiu Savin. Rigidity of minimizers in nonlocal phase transitions ii. Analysis in Theory and Applications, (1):1, 2019.
  • [36] Gunther Schoeck. The Peierls energy revisited. Philosophical Magazine A, 79(11):2629–2636, 1999.
  • [37] Yannick Sire and Enrico Valdinoci. Fractional Laplacian phase transitions and boundary reactions: a geometric inequality and a symmetry result. Journal of Functional Analysis, 256(6):1842–1864, 2009.
  • [38] Elias M Stein. Singular integrals and differentiability properties of functions, volume 2. Princeton university press, 1970.
  • [39] WM Stobbs and CH Sworn. The weak beam technique as applied to the determination of the stacking-fault energy of copper. Philosophical Magazine, 24(192):1365–1381, 1971.
  • [40] Yang Xiang, He Wei, Pingbing Ming, and Weinan E. A generalized Peierls–Nabarro model for curved dislocations and core structures of dislocation loops in Al and Cu. Acta materialia, 56(7):1447–1460, 2008.
  • [41] Kösaku Yosida. Functional analysis. Springer Science & Business Media, 2012.
  • [42] Jonathan A Zimmerman, Huajian Gao, and Farid F Abraham. Generalized stacking fault energies for embedded atom FCC metals. Modelling and Simulation in Materials Science and Engineering, 8(2):103, 2000.

Appendix A Review of functional analysis

For the sake of completeness, we prove the fact that operator LL defined in (1.38) and \mathcal{L} in (1.24) are both self-adjoint in Lemma 5.1 and Lemma 5.2 which address the spectrum of LL. Let us recall that linear operator LL is given by

L:H1(Ω)L2(Ω)L2(Ω),Lϕ=ϕ+γ′′(u)ϕ.\displaystyle L:H^{1}(\Omega)\subset L^{2}(\Omega)\to L^{2}(\Omega),\ L\phi=\mathcal{L}\phi+\gamma^{\prime\prime}(u)\phi.

Here uu is a layer solution to equation (1.15). For theorems and definitions in functional analysis, one can refer to [41]. For perturbation theory of self-adjoint operators, one can refer to [25].

In fact, the fact that LL is self-adjoint is a corollary of Kato-Rellich’s theorem (see [25]). We still repeat the proof for readers’ convenience. The proof here needs an equivalent criterion for self-adjoint operators:

Lemma A.1.

Suppose that HH is a complex Hilbert space with inner product ,H\langle\cdot,\cdot\rangle_{H} and A:HHA:H\to H is a symmetry operator on HH. Then AA is self-adjoint if and only if

(A.1) Ran(A±i)=H.\displaystyle\mathrm{Ran}(A\pm i)=H.
Proof.

\implies: Suppose that AA is self-adjoint, we prove that Ran(A±i)=H\mathrm{Ran}(A\pm i)=H. To prove this, notice that for any wHw\in H, we have

(A±i)w2=Aw2+w2w2,\displaystyle\|(A\pm i)w\|^{2}=\|Aw\|^{2}+\|w\|^{2}\geq\|w\|^{2},

so by the closed image theorem, we know that Ran(A±i)\mathrm{Ran}(A\pm i) is closed and Ker(A±i)={0}\mathrm{Ker}(A\pm i)=\{0\}. Also notice that

Ran(A±i)=Ker(Ai)={0},\displaystyle\mathrm{Ran}(A\pm i)^{\perp}=\mathrm{Ker}(A\mp i)=\{0\},

so Ran(A±i)\mathrm{Ran}(A\pm i) are dense in HH by the Hahn-Banach theorem. Remember that they are also closed, so Ran(A±i)=H\mathrm{Ran}(A\pm i)=H.

\impliedby: Suppose that Ran(A±i)=H\mathrm{Ran}(A\pm i)=H. We prove that AA is self-adjoint. Because AA is symmetry, so we only need to prove that dom(A)dom(A)\mathrm{dom}(A^{*})\subset\mathrm{dom}(A) since dom(A)dom(A)\mathrm{dom}(A)\subset\mathrm{dom}(A^{*}) holds for any symmetry operator. Notice that

Ker(Ai)=Ran(A±i)={0},\displaystyle\mathrm{Ker}(A^{*}\mp i)=\mathrm{Ran}(A\pm i)^{\perp}=\{0\},

so Ker(Ai)={0}\mathrm{Ker}(A^{*}\mp i)=\{0\}. Remember that Ran(A±i)=H\mathrm{Ran}(A\pm i)=H, so for any xdom(A)x\in\mathrm{dom}(A^{*}), there exists zdom(A)z\in\mathrm{dom}(A) such that

(A±i)x=(A±i)z.\displaystyle(A^{*}\pm i)x=(A\pm i)z.

So (A±i)(xz)=0(A^{*}\pm i)(x-z)=0. Here we used that for zdom(A)z\in\mathrm{dom}(A), Az=AzAz=A^{*}z. Thus xzKer(A±i)={0}x-z\in\mathrm{Ker}(A^{*}\pm i)=\{0\}. So x=zx=z and A=AA=A^{*}. So AA is self-adjoint. ∎

Lemma A.2.

Operator LL defined in (1.38) and operator \mathcal{L} defined in (1.24) are self-adjoint.

Proof.

First of all, both LL and \mathcal{L} are symmetric by Assumption A. For any w,vH1(Ωd)w,v\in H^{1}(\Omega_{d}), we have

w,vL2(Ωd)\displaystyle\langle w,\mathcal{L}v\rangle_{L^{2}(\Omega_{d})} =w^,σL(𝝂)v^L2(Ωd)=σL(𝝂)w^,v^L2(Ωd)=w,vL2(Ωd)\displaystyle=\langle\hat{w},\sigma_{L}(\bm{\nu})\hat{v}\rangle_{L^{2}(\Omega_{d}^{\prime})}=\langle\sigma_{L}(\bm{\nu})\hat{w},\hat{v}\rangle_{L^{2}(\Omega_{d}^{\prime})}=\langle\mathcal{L}w,v\rangle_{L^{2}(\Omega_{d})}
w,LvL2(Ωd)\displaystyle\langle w,Lv\rangle_{L^{2}(\Omega_{d})} =w,v+γ′′(u)vL2(Ωd)=w+γ′′(u)w,vL2(Ωd)=Lw,vL2(Ωd).\displaystyle=\langle w,\mathcal{L}v+\gamma^{\prime\prime}(u)v\rangle_{L^{2}(\Omega_{d})}=\langle\mathcal{L}w+\gamma^{\prime\prime}(u)w,v\rangle_{L^{2}(\Omega_{d})}=\langle Lw,v\rangle_{L^{2}(\Omega_{d})}.

So they are all symmetric.

Then we prove that \mathcal{L} in (1.24) is self-adjoint. By Lemma A.1, we only need to prove that Ran(±i)=L2(Ωd)\mathrm{Ran}(\mathcal{L}\pm i)=L^{2}(\Omega_{d}). We prove for only +i\mathcal{L}+i, the other side direction is just the same.

To prove this, we only need to prove that for any vL2(Ωd)v\in L^{2}(\Omega_{d}), there exists uH1(Ωd)u\in H^{1}(\Omega_{d}) such that

(+i)u=v.\displaystyle(\mathcal{L}+i)u=v.

One can rewrite this equality on the Fourier side as

(σL(𝝂)+i)u^(𝝂)=v^(𝝂).\displaystyle(\sigma_{L}(\bm{\nu})+i)\hat{u}(\bm{\nu})=\hat{v}(\bm{\nu}).

Thus

(A.2) u^(𝝂)=v^(𝝂)σL(𝝂)+i.\displaystyle\hat{u}(\bm{\nu})=\dfrac{\hat{v}(\bm{\nu})}{\sigma_{L}(\bm{\nu})+i}.

So we only need to prove that for any vL2(Ωd)v\in L^{2}(\Omega_{d}), uu in (A.2) is in H1(Ωd)H^{1}(\Omega_{d}). This is true by Assumption A which assumes that σ(𝝂)\sigma(\bm{\nu}) is real and with same order as |𝝂||\bm{\nu}|:

uH1(Ωd)2\displaystyle\|u\|_{H^{1}(\Omega_{d})}^{2} =u^(𝝂),u^(𝝂)L2(Ωd)+|𝝂|u^(𝝂),|𝝂|u^(𝝂)L2(Ωd)\displaystyle=\langle\hat{u}(\bm{\nu}),\hat{u}(\bm{\nu})\rangle_{L^{2}(\Omega_{d}^{\prime})}+\langle|\bm{\nu}|\hat{u}(\bm{\nu}),|\bm{\nu}|\hat{u}(\bm{\nu})\rangle_{L^{2}(\Omega_{d}^{\prime})}
=v^(𝝂),|𝝂|2+1σL2(𝝂)+1v^(𝝂)L2(Ωd)\displaystyle=\left\langle\hat{v}(\bm{\nu}),\dfrac{|\bm{\nu}|^{2}+1}{\sigma_{L}^{2}(\bm{\nu})+1}\hat{v}(\bm{\nu})\right\rangle_{L^{2}(\Omega_{d}^{\prime})}
1c2vL2(Ω)2.\displaystyle\leq\dfrac{1}{c^{2}}\|v\|_{L^{2}(\Omega)}^{2}.

Here c>0c>0 is the constant in Assumption A. So by Lemma A.1, we know that \mathcal{L} is self-adjoint.

Finally, we prove that LL in (1.38) is self-adjoint. Denote A=A=\mathcal{L} and B=γ′′(u)B=\gamma^{\prime\prime}(u) who is understood as a multiplier, then L=A+BL=A+B. First, because AA is self-adjoint, so by Lemma A.1, Ran(A±μi)=L2(Ωd)\mathrm{Ran}(A\pm\mu i)=L^{2}(\Omega_{d}) for any real number μ>0\mu>0.

Moreover, there also exists μ>0\mu>0 such that Ran(A+B±μi)=L2(Ωd)\mathrm{Ran}(A+B\pm\mu i)=L^{2}(\Omega_{d}). To prove this, notice that for any yH1(Ωd)y\in H^{1}(\Omega_{d}), we have

(A.3) (A±μi)y2=Ay2+μ2y2.\displaystyle\|(A\pm\mu i)y\|^{2}=\|Ay\|^{2}+\mu^{2}\|y\|^{2}.

Then take y=(A±μi)1xy=(A\pm\mu i)^{-1}x for any xL2(Ωd)x\in L^{2}(\Omega_{d}), we have

A(A±μi)1x2\displaystyle\|A(A\pm\mu i)^{-1}x\|^{2} =(A±μi)(A±μi)1x2μ2(A±μi)1x2x2,\displaystyle=\|(A\pm\mu i)(A\pm\mu i)^{-1}x\|^{2}-\mu^{2}\|(A\pm\mu i)^{-1}x\|^{2}\leq\|x\|^{2},
μ2(A±μi)1x2\displaystyle\mu^{2}\|(A\pm\mu i)^{-1}x\|^{2} =(A±μi)(A±μi)1x2A(A±μi)1x2x2.\displaystyle=\|(A\pm\mu i)(A\pm\mu i)^{-1}x\|^{2}-\|A(A\pm\mu i)^{-1}x\|^{2}\leq\|x\|^{2}.

so A(A±i)11\|A(A\pm i)^{-1}\|\leq 1 and (A±μi)11μ\|(A\pm\mu i)^{-1}\|\leq\dfrac{1}{\mu}. Notice that for sufficiently large μ\mu, we have B(A±μi)1<1\|B(A\pm\mu i)^{-1}\|<1 since BB is a bounded linear operator and

B(A±μi)1xb(A±μi)1xbμx.\displaystyle\|B(A\pm\mu i)^{-1}x\|\leq{b}\|(A\pm\mu i)^{-1}x\|\leq\dfrac{b}{\mu}\|x\|.

So by choosing sufficiently large μ\mu, we have B(A±μi)1<1\|B(A\pm\mu i)^{-1}\|<1. This gives that B(A±μi)1+IB(A\pm\mu i)^{-1}+I are invertible. Notice that

A+B±μi=[B(A±μi)1+I](A±μi)\displaystyle A+B\pm\mu i=[B(A\pm\mu i)^{-1}+I](A\pm\mu i)

so Ran(A+B±μi)=L2(Ωd)\mathrm{Ran}(A+B\pm\mu i)=L^{2}(\Omega_{d}) since Ran(A±μi)=L2(Ωd)\mathrm{Ran}(A\pm\mu i)=L^{2}(\Omega_{d}) and B(A±μi)1+IB(A\pm\mu i)^{-1}+I is invertible. Then by Lemma A.1, L=A+BL=A+B is self-adjoint. ∎

Now we prove Lemma 5.1. See 5.1

Proof.

Notice that LL is self-adjoint, so σ(L)=σp(L)σc(L)\sigma(L)=\sigma_{p}(L)\cup\sigma_{c}(L). Because limx±u(x,𝒚)=±1\lim\limits_{x\to\pm\infty}u(x,\bm{y})=\pm 1 holds uniformly in 𝒚\bm{y} and γ′′(±1)>0\gamma^{\prime\prime}(\pm 1)>0, so there exists λ1>0\lambda_{1}>0 such that γ′′(u(x,y))>λ1\gamma^{\prime\prime}(u(x,y))>-\lambda_{1} holds by continuity of uu and γ\gamma. Now we prove that for any λ[λ1,+)\lambda\in\mathbb{C}\setminus[-\lambda_{1},+\infty), λIL\lambda I-L has a bounded inverse. This directly shows σ(L)[λ1,+)\sigma(L)\subset[-\lambda_{1},+\infty).

First, Ran(λIL)\mathrm{Ran}(\lambda I-L) is closed. Let λ=a+bi\lambda=a+bi. For any wH1(Ωd)w\in H^{1}(\Omega_{d}), if b0b\neq 0, we have

(λIL)w2=(a2+b2)w2+Lw22aw,Lwb2w2.\displaystyle\|(\lambda I-L)w\|^{2}=(a^{2}+b^{2})\|w\|^{2}+\|Lw\|^{2}-2a\langle w,Lw\rangle\geq b^{2}\|w\|^{2}.

If b=0b=0 but a<λ1a<-\lambda_{1}, we have

(λIL)w2=(a+λ1)2w2+(L+λ1)w22(a+λ1)w,(L+λ1)w(a+λ1)2w2.\displaystyle\|(\lambda I-L)w\|^{2}=(a+\lambda_{1})^{2}\|w\|^{2}+\|(L+\lambda_{1})w\|^{2}-2(a+\lambda_{1})\langle w,(L+\lambda_{1})w\rangle\geq(a+\lambda_{1})^{2}\|w\|^{2}.

This is because

w,(L+λ1)w0\displaystyle\langle w,(L+\lambda_{1})w\rangle\geq 0

since γ′′(u)>λ1\gamma^{\prime\prime}(u)>-\lambda_{1} and \mathcal{L} is positively semi-definite. Thus for any λ[λ1,+)\lambda\in\mathbb{C}\setminus[-\lambda_{1},+\infty), there exists c>0c>0 such that (λIL)wcw\|(\lambda I-L)w\|\geq c\|w\| for any wH1(Ωd)w\in H^{1}(\Omega_{d}).

Therefore, by the closed image theorem, Ran(λIL)\mathrm{Ran}(\lambda I-L) is closed and Ker(λIL)={0}\mathrm{Ker}(\lambda I-L)=\{0\}. So λIL\lambda I-L is injective. Moreover, we have

Ran(λIL)=Ker(λIL)={0}\displaystyle\mathrm{Ran}(\lambda I-L)^{\perp}=\mathrm{Ker}(\lambda^{*}I-L)=\{0\}

since λ\lambda^{*} also belongs to [λ1,+)\mathbb{C}\setminus[-\lambda_{1},+\infty). So by the Hahn-Banach theorem, Ran(λIL)¯=L2(Ωd)\overline{\mathrm{Ran}(\lambda I-L)}=L^{2}(\Omega_{d}). Remember that Ran(λIL)\mathrm{Ran}(\lambda I-L) is closed, so Ran(λIL)=L2(Ωd)\mathrm{Ran}(\lambda I-L)=L^{2}(\Omega_{d}). Thus λIL\lambda I-L is a bijection. Because LL is self-adjoint, so λIL\lambda I-L is closed, so is (λIL)1(\lambda I-L)^{-1}. Thus by the closed graph theorem, (λIL)1(\lambda I-L)^{-1} is bounded. Therefore, λ\lambda is not in the spectrum of LL. ∎

Finally, we prove Lemma 5.2. To prove this lemma, we need to employ Weyl’s theorem on perturbation of self-adjoint operators.

Lemma A.3.

(Weyl’s theorem [25]) Suppose that HH is a Hilbert space, AA is a self-adjoint operator on HH and BB is a symmetric operator on HH. Then if BB is relatively compact with respect to AA, then σess(A+B)=σess(A)\sigma_{ess}(A+B)=\sigma_{ess}(A).

See 5.2

Proof.

Define function f:Ωdf:\Omega_{d}\to\mathbb{R} as

f(x,𝒚)={γ′′(1),ifx>0,γ′′(1),ifx0.\displaystyle f(x,\bm{y})=\begin{cases}\gamma^{\prime\prime}(1),\ &\mathrm{if}\ x>0,\\ \gamma^{\prime\prime}(-1),\ &\mathrm{if}\ x\leq 0.\end{cases}

Notice that γ′′(±1)>0\gamma^{\prime\prime}(\pm 1)>0 and u(x,𝒚)±u(x,\bm{y})\to\pm as xx\to\infty holds uniformly in 𝒚\bm{y}, so there exists c>0c>0 such that f>cf>c and

lim|x|γ′′(u(x,𝒚))f(x,𝒚)=0\displaystyle\lim\limits_{|x|\to\infty}\gamma^{\prime\prime}(u(x,\bm{y}))-f(x,\bm{y})=0

holds uniformly in 𝒚\bm{y} direction. Now we rewrite operator LL as

L=A+B,A=+f(x,y),B=γ′′(u)f(x,y).\displaystyle L=A+B,\ A=\mathcal{L}+f(x,y),\ B=\gamma^{\prime\prime}(u)-f(x,y).

BB is understood as a multiplier. We will prove that BB is relatively compact with respect to AA. Suppose that {uj}L2(Ωd)\{u_{j}\}\subset L^{2}(\Omega_{d}) is bounded. We only need to prove that {B(A+i)1uj}j\{B(A+i)^{-1}u_{j}\}_{j} is compact in L2(Ωd)L^{2}(\Omega_{d}).

Denote wj=(A+i)1ujw_{j}=(A+i)^{-1}u_{j}. We only need to prove that for any ϵ>0\epsilon>0, there exists a subsequence of {wj}\{w_{j}\} such that Bwj,nBwj,mϵ\|Bw_{j,n}-Bw_{j,m}\|\leq\epsilon. First of all, because wj=(A+i)1ujw_{j}=(A+i)^{-1}u_{j}, thus by Lemma 2.2, we have

|wj,uj|\displaystyle|\langle w_{j},u_{j}\rangle| =|wj,(A+i)wj|\displaystyle=|\langle w_{j},(A+i)w_{j}\rangle|
|w,w+wj,f(x,𝒚)wj+iwj2|\displaystyle\geq|\langle w,\mathcal{L}w\rangle+\langle w_{j},f(x,\bm{y})w_{j}\rangle+i\|w_{j}\|^{2}|
cwjH1/2(Ωd)2.\displaystyle\geq c\|w_{j}\|^{2}_{H^{1/2}(\Omega_{d})}.

Here c>0c>0 is a constant that only depends on \mathcal{L}. Then by the Cauchy-Schwartz inequality, we know that

12cuj2+c2wj2\displaystyle\dfrac{1}{2c}\|u_{j}\|^{2}+\dfrac{c}{2}\|w_{j}\|^{2} |wj,uj|cwjH1/2(Ωd)2.\displaystyle\geq|\langle w_{j},u_{j}\rangle|\geq c\|w_{j}\|^{2}_{H^{1/2}(\Omega_{d})}.

Thus there exists c>0c^{\prime}>0 such that wjH1/2(Ωd)2cuj2\|w_{j}\|^{2}_{H^{1/2}(\Omega_{d})}\leq c^{\prime}\|u_{j}\|^{2}, thus {wj}\{w_{j}\} is bounded in H1/2(Ωd)H^{1/2}(\Omega_{d}). Moreover, for any ϵ1\epsilon_{1} sufficiently small, there exists R>0R>0 such that

|γ′′(u(x,𝒚))f(x,𝒚)|ϵ1\displaystyle|\gamma^{\prime\prime}(u(x,\bm{y}))-f(x,\bm{y})|\leq\epsilon_{1}

for (x,𝒚)[R,R]c×𝕋d1(x,\bm{y})\in[-R,R]^{c}\times\mathbb{T}^{d-1}. Therefore,

BwjBwkL2([R,R]c×𝕋d1)2\displaystyle\|Bw_{j}-Bw_{k}\|^{2}_{L^{2}([-R,R]^{c}\times\mathbb{T}^{d-1})} =[R,R]c×𝕋d1|BwjBwk|2dxd𝒚\displaystyle=\int_{[-R,R]^{c}\times\mathbb{T}^{d-1}}|Bw_{j}-Bw_{k}|^{2}\mathrm{d}x\mathrm{d}\bm{y}
ϵ12wjwkL2(Ωd)2<ϵ2\displaystyle\leq\epsilon_{1}^{2}\|w_{j}-w_{k}\|^{2}_{L^{2}(\Omega_{d})}<\dfrac{\epsilon}{2}

by selecting ϵ1\epsilon_{1} sufficiently small. Then by compact embedding of H1/2([R,R]×𝕋d1)L2([R,R]×𝕋d1)H^{1/2}([-R,R]\times\mathbb{T}^{d-1})\subset L^{2}([-R,R]\times\mathbb{T}^{d-1}) and boundedness of BB, we know that there exists a subsequent of wj{w_{j}} (still denoted as wjw_{j}) such that BwjBwkL2([R,R]×𝕋d1)2ϵ2\|Bw_{j}-Bw_{k}\|^{2}_{L^{2}([-R,R]\times\mathbb{T}^{d-1})}\leq\dfrac{\epsilon}{2}. Then for this subsequence, we have

BwjBwkL2(Ωd)2=BwjBwkL2([R,R]×𝕋d1)2+BwjBwkL2([R,R]c×𝕋d1)2ϵ.\displaystyle\|Bw_{j}-Bw_{k}\|^{2}_{L^{2}(\Omega_{d})}=\|Bw_{j}-Bw_{k}\|^{2}_{L^{2}([-R,R]\times\mathbb{T}^{d-1})}+\|Bw_{j}-Bw_{k}\|^{2}_{L^{2}([-R,R]^{c}\times\mathbb{T}^{d-1})}\leq{\epsilon}.

This proves that σess(L)=σess(A)\sigma_{ess}(L)=\sigma_{ess}(A). However, since f(x,𝒚)>c>0f(x,\bm{y})>c>0 is uniformly bounded from below, so A=+f(x,𝒚)A=\mathcal{L}+f(x,\bm{y}) is positively definite and σ(A)[c,+)\sigma(A)\subset[c,+\infty). Thus σc(L)σess(L)[c,+)\sigma_{c}(L)\subset\sigma_{ess}(L)\subset[c,+\infty). Taking λ2=c\lambda_{2}=c closes the proof. ∎

Appendix B Proof of lemmas

Proof of Lemma 1.1.

From Definition 1 of minimizers, we calculate the variation of energy in terms of a perturbation with compact support in an arbitrary ball B(R)3B(R)\subset\mathbb{R}^{3} which is centered at 𝟎\bm{0} with radius RR. For any 𝒗C(B(R)\Γ)\bm{v}\in C^{\infty}(B(R)\backslash\Gamma) such that 𝒗\bm{v} has compact support in B(R)B(R) and satisfies (1.11), we consider the perturbation δ𝒗\delta\bm{v} where δ\delta is a small real number. We denote ε:=ε(𝒖)\varepsilon:=\varepsilon(\bm{u}), σ:=σ(𝒖)\sigma:=\sigma(\bm{u}) and ε1:=ε(𝒗)\varepsilon_{1}:=\varepsilon(\bm{v}), σ1:=σ(𝒗)\sigma_{1}:=\sigma(\bm{v}). Then we have that

(B.1) limδ01δ(E(𝒖+δ𝒗)E(𝒖))\displaystyle\lim_{\delta\to 0}\frac{1}{\delta}(E(\bm{u}+\delta\bm{v})-E(\bm{u}))
=\displaystyle= B(R)\Γ12(σ1:ε+σ:ε1)dxdydz+B(R)Γu1γ(u1+,u3+)v1++u3γ(u1+,u3+)v3+dxdz\displaystyle\int_{B(R)\backslash\Gamma}\frac{1}{2}(\sigma_{1}:\varepsilon+\sigma:\varepsilon_{1})\,\mathrm{d}x\,\mathrm{d}y\,\mathrm{d}z+\int_{B(R)\cap\Gamma}\partial_{u_{1}}\gamma(u_{1}^{+},u_{3}^{+})v_{1}^{+}+\partial_{u_{3}}\gamma(u_{1}^{+},u_{3}^{+})v_{3}^{+}\,\mathrm{d}x\,\mathrm{d}z
=\displaystyle= B(R)\Γσ:ε1dxdydz+B(R)Γu1γ(u1+,u3+)v1++u3γ(u1+,u3+)v3+dxdz\displaystyle\int_{B(R)\backslash\Gamma}\sigma:\varepsilon_{1}\,\mathrm{d}x\,\mathrm{d}y\,\mathrm{d}z+\int_{B(R)\cap\Gamma}\partial_{u_{1}}\gamma(u_{1}^{+},u_{3}^{+})v_{1}^{+}+\partial_{u_{3}}\gamma(u_{1}^{+},u_{3}^{+})v_{3}^{+}\,\mathrm{d}x\,\mathrm{d}z
=\displaystyle= B(R)\Γσ:𝒗dxdydz+B(R)Γu1γ(u1+,u3+)v1++u3γ(u1+,u3+)v3+dxdz\displaystyle\int_{B(R)\backslash\Gamma}\sigma:\nabla\bm{v}\,\mathrm{d}x\,\mathrm{d}y\,\mathrm{d}z+\int_{B(R)\cap\Gamma}\partial_{u_{1}}\gamma(u_{1}^{+},u_{3}^{+})v_{1}^{+}+\partial_{u_{3}}\gamma(u_{1}^{+},u_{3}^{+})v_{3}^{+}\,\mathrm{d}x\,\mathrm{d}z
=\displaystyle= B(R)\Γjσijvidxdydz+B(R){y=0+}σij+nj+vi+dxdz\displaystyle-\int_{B(R)\backslash\Gamma}\partial_{j}\sigma_{ij}v_{i}\,\mathrm{d}x\,\mathrm{d}y\,\mathrm{d}z+\int_{B(R)\cap\{y=0^{+}\}}\sigma_{ij}^{+}n_{j}^{+}v_{i}^{+}\,\mathrm{d}x\,\mathrm{d}z
+B(R){y=0}σijnjvidxdz+B(R)Γu1γ(u1+,u3+)v1++u3γ(u1+,u3+)v3+dxdz0\displaystyle\quad+\int_{B(R)\cap\{y=0^{-}\}}\sigma_{ij}^{-}n_{j}^{-}v_{i}^{-}\,\mathrm{d}x\,\mathrm{d}z+\int_{B(R)\cap\Gamma}\partial_{u_{1}}\gamma(u_{1}^{+},u_{3}^{+})v_{1}^{+}+\partial_{u_{3}}\gamma(u_{1}^{+},u_{3}^{+})v_{3}^{+}\,\mathrm{d}x\,\mathrm{d}z\geq 0

where we used the property that σ\sigma and σ\nabla\cdot\sigma are locally integrable in {y>0}{y<0}\{y>0\}\cup\{y<0\} when carrying out the integration by parts, and the outer normal vector of the boundary Γ\Gamma is 𝐧+\mathbf{n}^{+} (resp. the 𝐧\mathbf{n}^{-}) for the upper half-plane (resp. lower half-plane). Similarly, taking perturbation as 𝒗-\bm{v} and notice that that 𝐧+=(0,1,0)\mathbf{n}^{+}=(0,-1,0) and 𝐧=(0,1,0)\mathbf{n}^{-}=(0,1,0), we have

(B.2) {y=0+}σij+nj+vi+dxdz+{y=0}σijnjvidxdz\displaystyle\int_{\{y=0^{+}\}}\sigma_{ij}^{+}n_{j}^{+}v_{i}^{+}\,\mathrm{d}x\,\mathrm{d}z+\int_{\{y=0^{-}\}}\sigma_{ij}^{-}n_{j}^{-}v_{i}^{-}\,\mathrm{d}x\,\mathrm{d}z
=\displaystyle= {y=0+}σ22+v2+dxdz+{y=0}σ22v2dxdz+{y=0+}σ12+v1+dxdz+{y=0}σ12v1dxdz\displaystyle\int_{\{y=0^{+}\}}-\sigma_{22}^{+}v_{2}^{+}\,\mathrm{d}x\,\mathrm{d}z+\int_{\{y=0^{-}\}}\sigma_{22}^{-}v_{2}^{-}\,\mathrm{d}x\,\mathrm{d}z+\int_{\{y=0^{+}\}}-\sigma_{12}^{+}v_{1}^{+}\,\mathrm{d}x\,\mathrm{d}z+\int_{\{y=0^{-}\}}\sigma_{12}^{-}v_{1}^{-}\,\mathrm{d}x\,\mathrm{d}z
+{y=0+}σ32+v3+dxdz+{y=0}σ32v3dxdz\displaystyle+\int_{\{y=0^{+}\}}-\sigma_{32}^{+}v_{3}^{+}\,\mathrm{d}x\,\mathrm{d}z+\int_{\{y=0^{-}\}}\sigma_{32}^{-}v_{3}^{-}\,\mathrm{d}x\,\mathrm{d}z

Since v1+(x,z)=v1(x,z)v_{1}^{+}(x,z)=-v_{1}^{-}(x,z), v3+(x,z)=v3(x,z)v_{3}^{+}(x,z)=-v_{3}^{-}(x,z) and v2+(x,z)=v2(x,z)v_{2}^{+}(x,z)=v_{2}^{-}(x,z). Hence due to the arbitrariness of RR, we conclude that the minimizer 𝒖\bm{u} must satisfy

(B.3) Γ[σ12++σ12u1γ(u1+,u3+)]v1+dxdz=0,\displaystyle\int_{\Gamma}\left[\sigma_{12}^{+}+\sigma_{12}^{-}-\partial_{u_{1}}\gamma(u_{1}^{+},u_{3}^{+})\right]v_{1}^{+}\,\mathrm{d}x\,\mathrm{d}z=0,
Γ[σ32++σ32u3γ(u1+,u3+)]v3+dxdz=0,\displaystyle\int_{\Gamma}\left[\sigma_{32}^{+}+\sigma_{32}^{-}-\partial_{u_{3}}\gamma(u_{1}^{+},u_{3}^{+})\right]v_{3}^{+}\,\mathrm{d}x\,\mathrm{d}z=0,
Γ(σ22+σ22)v2+dxdz=0,\displaystyle\int_{\Gamma}\left(\sigma_{22}^{+}-\sigma_{22}^{-}\right)v_{2}^{+}\,\mathrm{d}x\,\mathrm{d}z=0,
3\Γ(σ)𝒗dxdydz=0\displaystyle\int_{\mathbb{R}^{3}\backslash\Gamma}(\nabla\cdot\sigma)\cdot\bm{v}~{}\,\mathrm{d}x\,\mathrm{d}y\,\mathrm{d}z=0

for any 𝒗C(B(R)\Γ)\bm{v}\in C^{\infty}(B(R)\backslash\Gamma) and 𝒗\bm{v} has compact support in B(R)B(R), which leads to the Euler–Lagrange equation (1.13). Here we have written the equation σ=0\nabla\cdot\sigma=0 in 3\Γ\mathbb{R}^{3}\backslash\Gamma as the first equation of (1.13) in terms of the displacement 𝒖\bm{u}. ∎

Proof of Lemma 3.1.

If a1=a2a_{1}=a_{2} or b1=b2b_{1}=b_{2} holds, then the equality holds. So we will focus on cases where a1a2a_{1}\neq a_{2} and b1b2b_{1}\neq b_{2}. By enumeration of all possible orders, we have:

  1. (i)

    If a1>a2a_{1}>a_{2} and b1>b2b_{1}>b_{2}, then a=a2,A=a1,b=b2a=a_{2},A=a_{1},b=b_{2} and B=b1B=b_{1}. So

    ab+ABa1b1a2b2=ab+ABABab=0.\displaystyle ab+AB-a_{1}b_{1}-a_{2}b_{2}=ab+AB-AB-ab=0.

    The equality in (3.4) holds.

  2. (ii)

    If a1>a2a_{1}>a_{2} and b1<b2b_{1}<b_{2}, then a=a2,A=a1,b=b1a=a_{2},A=a_{1},b=b_{1} and B=b2B=b_{2}. So

    ab+ABa1b1a2b2=ab+ABAbaB=(aA)(bB)>0.\displaystyle ab+AB-a_{1}b_{1}-a_{2}b_{2}=ab+AB-Ab-aB=(a-A)(b-B)>0.

    The ’>>’ in (3.4) holds.

  3. (iii)

    If a1<a2a_{1}<a_{2} and b1>b2b_{1}>b_{2}, then a=a1,A=a2,b=b2a=a_{1},A=a_{2},b=b_{2} and B=b1B=b_{1}. So

    ab+ABa1b1a2b2=ab+ABaBAb=(aA)(bB)>0.\displaystyle ab+AB-a_{1}b_{1}-a_{2}b_{2}=ab+AB-aB-Ab=(a-A)(b-B)>0.

    The ’>>’ in (3.4) holds.

  4. (iv)

    If a1<a2a_{1}<a_{2} and b1<b2b_{1}<b_{2}, then a=a1,A=a2,b=b1a=a_{1},A=a_{2},b=b_{1} and B=b2B=b_{2}. So

    ab+ABa1b1a2b2=ab+ABabAB=0.\displaystyle ab+AB-a_{1}b_{1}-a_{2}b_{2}=ab+AB-ab-AB=0.

    The equality in (3.4) holds.

Therefore, the inequality holds. The equality is attained if and only if a1=a2a_{1}=a_{2} or b1=b2b_{1}=b_{2} or the order is preserved, i.e. a1<a2,b1<b2a_{1}<a_{2},b_{1}<b_{2} or a1>a2,b1>b2a_{1}>a_{2},b_{1}>b_{2}. These conditions are equivalent to the following clear inequality: (a1a2)(b1b2)0(a_{1}-a_{2})(b_{1}-b_{2})\geq 0. This concludes the proof. ∎

Proof of Lemma 3.3.

In fact, by change of variables, we have

F(u(x+c1,𝒚+𝒄2))\displaystyle F(u(x+c_{1},\bm{y}+\bm{c}_{2})) =12ΩdΩd|(u(x+c1,𝒚+𝒄2)u(x+c1,𝒚+𝒄2)|2K(xx,𝒚𝒚)\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(u(x+c_{1},\bm{y}+\bm{c}_{2})-u(x^{\prime}+c_{1},\bm{y}^{\prime}+\bm{c}_{2})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})
|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚\displaystyle-{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}
+Ωdγ(u(x+c1,𝒚+𝒄2))dxd𝒚\displaystyle+\int_{\Omega_{d}}\gamma(u(x+c_{1},\bm{y}+\bm{c}_{2}))\mathrm{d}x\mathrm{d}\bm{y}
=12ΩdΩd|(u(x,𝒚)u(x,𝒚)|2K(xx,𝒚𝒚)\displaystyle=\dfrac{1}{2}\int_{\Omega_{d}}\int_{\Omega_{d}}{|(u(x,\bm{y})-u(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})
|(η(xc1,𝒚𝒄2)η(xc1,𝒚𝒄2)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚\displaystyle-{|(\eta(x-c_{1},\bm{y}-\bm{c}_{2})-\eta(x^{\prime}-c_{1},\bm{y}^{\prime}-\bm{c}_{2})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}
+Ωdγ(u(x,𝒚))dxd𝒚.\displaystyle+\int_{\Omega_{d}}\gamma(u(x,\bm{y}))\mathrm{d}x\mathrm{d}\bm{y}.

Thus by Lemma 2.2, we have

F(u(x,y))F(u(x+c1,𝒚+𝒄2))\displaystyle F(u(x,y))-F(u(x+c_{1},\bm{y}+\bm{c}_{2})) =12ΩΩ|η(x+c1,𝒚+𝒄2)η(x+c1,𝒚+𝒄2)|2K(xx,𝒚𝒚)\displaystyle=\dfrac{1}{2}\int_{\Omega}\int_{\Omega}{|\eta(x+c_{1},\bm{y}+\bm{c}_{2})-\eta(x^{\prime}+c_{1},\bm{y}^{\prime}+\bm{c}_{2})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})
|(η(x,𝒚)η(x,𝒚)|2K(xx,𝒚𝒚)dxdxd𝒚d𝒚\displaystyle-{|(\eta(x,\bm{y})-\eta(x^{\prime},\bm{y}^{\prime})|^{2}}K(x-x^{\prime},\bm{y}-\bm{y}^{\prime})\mathrm{d}x\mathrm{d}x^{\prime}\mathrm{d}\bm{y}\mathrm{d}\bm{y}^{\prime}
=A2(η(x+c1)η(x+c1))2(xx)2(η(x)η(x))2(xx)2dxdx.\displaystyle=\dfrac{A}{2}\int_{\mathbb{R}}\dfrac{(\eta(x+c_{1})-\eta(x^{\prime}+c_{1}))^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x\mathrm{d}x^{\prime}.

Here AA is the constant in Lemma 2.2. So we only need to prove that

(η(x+c)η(x+c))2(xx)2(η(x)η(x))2(xx)2dxdx=0\displaystyle\int_{\mathbb{R}}\int_{\mathbb{R}}\dfrac{(\eta(x+c)-\eta(x^{\prime}+c))^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x=0

for any cc\in\mathbb{R}. Without loss of generality, we assume that c>0c>0. Then for x1x\geq 1, we know that η(x)=η(x+c)=1\eta(x)=\eta(x+c)=1 and for x1cx\leq-1-c, we have η(x)=η(x+c)=1\eta(x)=\eta(x+c)=-1. Denote J=[1c,1]J=[-1-c,1], and we separate the integral into 3 different parts, i.e. integral on J×JJ\times J (denoted as I1I_{1}), J×JcJ\times J^{c} (denoted as I2I_{2}) and Jc×JcJ^{c}\times J^{c} (denoted as I3I_{3}). Since η(x)=η(x+c)\eta(x)=\eta(x+c) on JcJ^{c}, we know that

I3=JcJc(η(x+c)η(x+c))2(xx)2(η(x)η(x))2(xx)2dxdx=0.\displaystyle I_{3}=\int_{J^{c}}\int_{J^{c}}\dfrac{(\eta(x+c)-\eta(x^{\prime}+c))^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x=0.

On J×JJ\times J, we have

I1\displaystyle I_{1} =1c11c1(η(x+c)η(x+c))2(xx)2(η(x)η(x))2(xx)2dxdx\displaystyle=\int_{-1-c}^{1}\int_{-1-c}^{1}\dfrac{(\eta(x+c)-\eta(x^{\prime}+c))^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
=11+c11+c(η(x)η(x))2(xx)2dxdx1c11c1(η(x)η(x))2(xx)2dxdx\displaystyle=\int_{-1}^{1+c}\int_{-1}^{1+c}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x-\int_{-1-c}^{1}\int_{-1-c}^{1}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
=11+c11+c(η(x)η(x))2(xx)2dxdx+21111+c(η(x)η(x))2(xx)2dxdx\displaystyle=\int_{1}^{1+c}\int_{1}^{1+c}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x+2\int_{-1}^{1}\int_{1}^{1+c}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
1c11c1(η(x)η(x))2(xx)2dxdx2111c1(η(x)η(x))2(xx)2dxdx.\displaystyle-\int_{-1-c}^{-1}\int_{-1-c}^{-1}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x-2\int_{-1}^{1}\int_{-1-c}^{-1}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x.

Because η(x)=η(x)\eta(x)=\eta(x^{\prime}) if x,x1x,x^{\prime}\geq 1 or x,x1x,x^{\prime}\leq-1, so integral vanishes on [1,1+c]×[1,1+c][1,1+c]\times[1,1+c] or [1c,1]×[1c,1][-1-c,-1]\times[-1-c,-1]. Thus

I1\displaystyle I_{1} =21111+c(η(x)η(x))2(xx)2dxdx2111c1(η(x)η(x))2(xx)2dxdx\displaystyle=2\int_{-1}^{1}\int_{1}^{1+c}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x-2\int_{-1}^{1}\int_{-1-c}^{-1}\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
=21111+c(η(x)1)2(xx)2dxdx2111c1(η(x)+1)2(xx)2dxdx\displaystyle=2\int_{-1}^{1}\int_{1}^{1+c}\dfrac{(\eta(x)-1)^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x-2\int_{-1}^{1}\int_{-1-c}^{-1}\dfrac{(\eta(x)+1)^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
(B.4) =211(η(x)1)2x1c(η(x)1)2x1dx211(η(x)+1)2x+1(η(x)+1)2x+1+cdx.\displaystyle=2\int_{-1}^{1}\dfrac{(\eta(x)-1)^{2}}{x-1-c}-\dfrac{(\eta(x)-1)^{2}}{x-1}\mathrm{d}x-2\int_{-1}^{1}\dfrac{(\eta(x)+1)^{2}}{x+1}-\dfrac{(\eta(x)+1)^{2}}{x+1+c}\mathrm{d}x.

On J×JcJ\times J^{c}, we have

I2\displaystyle I_{2} =1c11+(η(x+c)η(x+c))2(xx)2(η(x)η(x))2(xx)2dxdx\displaystyle=\int_{-1-c}^{1}\int_{1}^{+\infty}\dfrac{(\eta(x+c)-\eta(x^{\prime}+c))^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
+1c11c(η(x+c)η(x+c))2(xx)2(η(x)η(x))2(xx)2dxdx\displaystyle+\int_{-1-c}^{1}\int_{-\infty}^{-1-c}\dfrac{(\eta(x+c)-\eta(x^{\prime}+c))^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
=1c11+(η(x+c)1)2(xx)2(η(x)1)2(xx)2dxdx\displaystyle=\int_{-1-c}^{1}\int_{1}^{+\infty}\dfrac{(\eta(x+c)-1)^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-1)^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
+1c11c(η(x+c)+1)2(xx)2(η(x)+1)2(xx)2dxdx\displaystyle+\int_{-1-c}^{1}\int_{-\infty}^{-1-c}\dfrac{(\eta(x+c)+1)^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)+1)^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x
=1c1(η(x+c)1)2(η(x)1)21xdx+1c1(η(x+c)+1)2(η(x)+1)2x+1+cdx\displaystyle=\int_{-1-c}^{1}\dfrac{(\eta(x+c)-1)^{2}-(\eta(x)-1)^{2}}{1-x}\mathrm{d}x+\int_{-1-c}^{1}\dfrac{(\eta(x+c)+1)^{2}-(\eta(x)+1)^{2}}{x+1+c}\mathrm{d}x

Notice that η(x+c)=1\eta(x+c)=1 for x[1c,1]x\in[1-c,1] and η(x)=1\eta(x)=-1 for x[1c,1]x\in[-1-c,-1], so we have

1c1(η(x+c)1)2(η(x)1)21xdx\displaystyle\ \ \ \ \int_{-1-c}^{1}\dfrac{(\eta(x+c)-1)^{2}-(\eta(x)-1)^{2}}{1-x}\mathrm{d}x
=1c1c(η(x+c)1)21xdx11(η(x)1)21xdx1c141xdx\displaystyle=\int_{-1-c}^{1-c}\dfrac{(\eta(x+c)-1)^{2}}{1-x}\mathrm{d}x-\int_{-1}^{1}\dfrac{(\eta(x)-1)^{2}}{1-x}\mathrm{d}x-\int_{-1-c}^{-1}\dfrac{4}{1-x}\mathrm{d}x
=11(η(x)1)21+cx(η(x)1)21xdx+4ln24ln(2+c)\displaystyle=\int_{-1}^{1}\dfrac{(\eta(x)-1)^{2}}{1+c-x}-\dfrac{(\eta(x)-1)^{2}}{1-x}\mathrm{d}x+4\ln 2-4\ln(2+c)

and

1c1(η(x+c)+1)2(η(x)+1)2x+1+cdx=11(η(x)+1)2x+1(η(x)+1)2x+1+cdx+4ln(2+c)4ln2.\displaystyle\int_{-1-c}^{1}\dfrac{(\eta(x+c)+1)^{2}-(\eta(x)+1)^{2}}{x+1+c}\mathrm{d}x=\int_{-1}^{1}\dfrac{(\eta(x)+1)^{2}}{x+1}-\dfrac{(\eta(x)+1)^{2}}{x+1+c}\mathrm{d}x+4\ln(2+c)-4\ln 2.

Then substituting these two formulas into I2I_{2}, we have

I2\displaystyle I_{2} =1c1(η(x+c)1)2(η(x)1)21xdx+1c1(η(x+c)+1)2(η(x)+1)2x+1+cdx\displaystyle=\int_{-1-c}^{1}\dfrac{(\eta(x+c)-1)^{2}-(\eta(x)-1)^{2}}{1-x}\mathrm{d}x+\int_{-1-c}^{1}\dfrac{(\eta(x+c)+1)^{2}-(\eta(x)+1)^{2}}{x+1+c}\mathrm{d}x
=11(η(x)1)21+cx(η(x)1)21xdx+4ln24ln(2+c)\displaystyle=\int_{-1}^{1}\dfrac{(\eta(x)-1)^{2}}{1+c-x}-\dfrac{(\eta(x)-1)^{2}}{1-x}\mathrm{d}x+4\ln 2-4\ln(2+c)
+11(η(x)+1)2x+1(η(x)+1)2x+1+cdx+4ln(2+c)4ln2\displaystyle+\int_{-1}^{1}\dfrac{(\eta(x)+1)^{2}}{x+1}-\dfrac{(\eta(x)+1)^{2}}{x+1+c}\mathrm{d}x+4\ln(2+c)-4\ln 2
(B.5) =11(η(x)1)21+cx(η(x)1)21xdx+11(η(x)+1)2x+1(η(x)+1)2x+1+cdx\displaystyle=\int_{-1}^{1}\dfrac{(\eta(x)-1)^{2}}{1+c-x}-\dfrac{(\eta(x)-1)^{2}}{1-x}\mathrm{d}x+\int_{-1}^{1}\dfrac{(\eta(x)+1)^{2}}{x+1}-\dfrac{(\eta(x)+1)^{2}}{x+1+c}\mathrm{d}x

A careful comparison of equation (B) and (B) shows that I1+2I2=0I_{1}+2I_{2}=0. Thus

(η(x+c)η(x+c))2(xx)2(η(x)η(x))2(xx)2dxdx=I1+2I2+I3=0.\displaystyle\int_{\mathbb{R}}\int_{\mathbb{R}}\dfrac{(\eta(x+c)-\eta(x^{\prime}+c))^{2}}{(x-x^{\prime})^{2}}-\dfrac{(\eta(x)-\eta(x^{\prime}))^{2}}{(x-x^{\prime})^{2}}\mathrm{d}x^{\prime}\mathrm{d}x=I_{1}+2I_{2}+I_{3}=0.