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

Nonlocal half-ball vector operators on bounded domains: Poincaré inequality and its applications

Zhaolong Han Department of Mathematics, University of California, San Diego, CA 92093, United States [email protected]  and  Xiaochuan Tian Department of Mathematics, University of California, San Diego, CA 92093, United States [email protected]
Abstract.

This work contributes to nonlocal vector calculus as an indispensable mathematical tool for the study of nonlocal models that arises in a variety of applications. We define the nonlocal half-ball gradient, divergence and curl operators with general kernel functions (integrable or fractional type with finite or infinite supports) and study the associated nonlocal vector identities. We study the nonlocal function space on bounded domains associated with zero Dirichlet boundary conditions and the half-ball gradient operator and show it is a separable Hilbert space with smooth functions dense in it. A major result is the nonlocal Poincaré inequality, based on which a few applications are discussed, and these include applications to nonlocal convection-diffusion, nonlocal correspondence model of linear elasticity, and nonlocal Helmholtz decomposition on bounded domains.

Key words and phrases:
nonlocal models, nonlocal vector calculus, Poincaré inequality, Helmholtz decomposition, bounded domains, nonlocal half-ball gradient operator, Riesz fractional gradient, Marchaud fractional derivative, peridynamics, correspondence model
2020 Mathematics Subject Classification:
45A05, 26B12, 47G10, 47G30, 35A23, 46E35, 74G65

1. Introduction

In recent decades, nonlocal models that account for interactions occurring at a distance have been increasingly popular in many scientific fields. In particular, they appear widely in applications in continuum mechanics, probability and finance, image processing and population dynamics, and have been shown to more faithfully and effectively model observed phenomena that involve possible discontinuities, singularities and other anomalies [4, 9, 13, 26, 30, 35, 45].

One type of nonlocal problem is featured with generalizing the integer-order scaling laws that appear in PDEs to scaling laws of non-integer orders. This type of problem usually involves integral operators with fractional kernels that are supported in the whole space (i.e., with infinite nonlocal interactions), such as the fractional Laplace operator that models non-standard diffusion of a fractional order [5, 33, 40]. Another type of nonlocal problem focuses on finite range interactions and connects to PDEs by localization of nonlocal interactions [18, 19]. A prominent example is peridynamics, a nonlocal continuum model in solid mechanics, which is shown to be consistent with the classical elasticity theory by localization [42, 48, 50]. Other nonlocal models in this type include nonlocal (convection-)diffusion and nonlocal Stokes equations with finite nonlocal interactions that are inspired by peridynamics. Nonlocal vector calculus is developed in [20] and is used for reformulating nonlocal problems under a more systematic framework analogous to classical vector calculus [19, 29]. See [16, 23] for surveys on connecting the fractional and nonlocal vector calculus.

There are two commonly used frameworks in nonlocal vector calculus [20]; one involves two-point nonlocal (difference) operators and another involves one-point nonlocal (integral) operators. The two-point nonlocal gradient operator and its adjoint operator are used to reformulate nonlocal diffusion and the bond-based peridynamics models [18]. The one-point nonlocal gradient operator, on the other hand, is also used in a variety of applications including nonlocal advection equation, nonlocal Stokes equation and the peridynamics correspondence models [21, 22, 37, 49]. In some sense, the one-point nonlocal operators, including nonlocal gradient, divergence and curl operators, are more convenient to use as modeling tools since they can be directly used in place of their classical counterparts appearing in PDEs. However, the mathematical properties of nonlocal models involving these operators are not readily guaranteed without careful investigation. For example, instability of the peridynamics correspondence model is observed in which nonlocal deformation gradient is used to replace the classical deformation gradient, and is later explained in [21] as a result of lack of conscious choice of the interaction kernels in the nonlocal gradient operators. Singular kernels are proposed in [21] for the remedy of instability which resembles the kernel functions in the Riesz fractional gradient in terms of singularity at origin [8, 46]. Later on, nonlocal gradient operators with hemispherical interaction neighborhoods are used in [38] so that the singularity in kernel functions is no longer a necessity for the corresponding nonlocal Dirichlet energies to be stable. Both [21] and [38] work on functions defined on periodic cells to facilitate Fourier analysis. The starting point of this work is to establish a functional analysis framework that extends the Fourier analysis in [38] and apply it to nonlocal Dirichlet boundary value problems. With a general setting, we work with kernels that include both the Riesz fractional type (with infinite support) and the compactly supported type inspired by peridynamics. We remark that in a recent work [8], the authors consider the truncated Riesz fractional type kernels defined with full spherical support and study the properties of the corresponding function spaces by establishing a nonlocal fundamental theorem of calculus, and no such formula exists for kernels with hemispherical interaction neighborhoods which are our main focus in this work.

The major contribution of this work is the study of the functional analysis properties of the nonlocal space associated with the half-ball nonlocal gradient operator defined on bounded domains. We show that the space is a Hilbert space, and more importantly, it is separable with smooth functions dense in it, a property on which many applications are based. Another major result is the Poincaré inequality on functions with zero Dirichlet boundary conditions. We spend two whole sections on the proof it, one for the case of integrable kernels with compact support and another for more general kernels, including non-integrable kernels and kernels with infinite supports. Poincaré inequality is crucial for the study of boundary value problems. Indeed, we illustrate its use in three applications. The first application is the well-posedness of a class of nonlocal convection-diffusion equations defined via nonlocal half-ball gradient and divergence. Secondly, we study the nonlocal correspondence model of linear elasticity, where we also show a nonlocal Korn’s inequality for functions with Dirichlet boundary conditions. Note that the convergence of Galerkin approximations to these equations is natural, although we do not illustrate it in detail due to the length of the paper, as a result of the separability of the associated nonlocal energy spaces. The last application is a nonlocal version of Helmholtz decomposition for vector fields defined on bounded domains, a result of the solvability of the nonlocal Poisson type problem and some nonlocal vector identities involving gradient, divergence and curl which we also established in this paper. We remark that Helmholtz decomposition for one-point nonlocal operators is also studied in [15, 31, 38], but only periodic domains or the whole space are considered in these works.

Outline of the paper. We start with the principal value definition of the nonlocal half-ball gradient, divergence and curl operators for measurable functions in Section 2, and the corresponding distributional gradient, divergence and curl operators are followed. Fourier symbols of these operators are studied for later use and some nonlocal vector identities for smooth functions are established in the section. In Section 3, we define the nonlocal function space associated with the Dirichlet integral defined via the distributional nonlocal half-ball gradient, an analogue of the H01H^{1}_{0} Sobolev space in the local case, and show it is a separable Hilbert space. Ingredients such as closedness under multiplication with smooth functions, continuity of translation and mollification are established to prove the density result. In addition, we show that the distributional divergence and curl are well-defined quantities in the L2L^{2} sense in the nonlocal function space for vector fields, an analogue of the fact that H1H(div)H^{1}\subset H(\text{div}) and H1H(curl)H^{1}\subset H(\text{curl}) in the local case. Thus the vector identities also hold for functions in the nonlocal function spaces. The nonlocal Poincaré inequality is proved for integrable kernels with compact support in Section 4, based on which the nonlocal Poincaré inequality is shown for more general kernels in Section 5. Section 6 contains three applications of our functional analysis framework, including applications to nonlocal convection-diffusion, nonlocal linear elasticity, and nonlocal Helmholtz decomposition on bounded domains. Finally, we conclude in Section 7.

2. Nonlocal half-ball vector operators

We introduce the nonlocal half-ball vector operators in this section and discuss their properties. In the following, we let 𝝂d\bm{\nu}\in\mathbb{R}^{d} be a fixed unit vector. Denote by χ𝝂(𝒛)\chi_{\bm{\nu}}(\bm{z}) the characteristic function of the half-space 𝝂:={𝒛d:𝒛𝝂0}\mathcal{H}_{\bm{\nu}}:=\{\bm{z}\in\mathbb{R}^{d}:\bm{z}\cdot\bm{\nu}\geq 0\} parameterized by the unit vector 𝝂\bm{\nu}.

Throughout the paper, we adopt the following notations in linear algebra. For two column vectors 𝒂,𝒃dd×1\bm{a},\bm{b}\in\mathbb{R}^{d}\cong\mathbb{R}^{d\times 1}, 𝒂𝒃\bm{a}\cdot\bm{b} is the dot product and 𝒂×𝒃\bm{a}\times\bm{b} is the cross product if d=3d=3. If 𝒂d\bm{a}\in\mathbb{R}^{d} and 𝒃N\bm{b}\in\mathbb{R}^{N}, then the tensor product of 𝒂\bm{a} and 𝒃\bm{b} is a d×Nd\times N matrix, given by 𝒂𝒃=(aibj)1id,1jN\bm{a}\otimes\bm{b}=(a_{i}b_{j})_{1\leq i\leq d,1\leq j\leq N}. For two matrices A,Bm,n()A,B\in\mathcal{M}_{m,n}(\mathbb{R}), we define A:B=i=1mj=1naijbijA:B=\sum_{i=1}^{m}\sum_{j=1}^{n}a_{ij}b_{ij}.

2.1. Definitions and integration by parts

Following the notion of nonlocal nonsymmetric operator defined in [38], we define the nonlocal half-ball vector operators as follows.

Throughout the paper, we assume that ww satisfies the following conditions:

(1) {wLloc1(d\{𝟎}),w0,wis radial;there existsϵ0(0,1) such that w(𝒙)>0 for 0<|𝒙|ϵ0;dmin(1,|𝒙|)w(𝒙)d𝒙=|𝒙|1w(𝒙)|𝒙|d𝒙+|𝒙|>1w(𝒙)d𝒙=:Mw1+Mw2<.\left\{\begin{aligned} &w\in L^{1}_{\mathrm{loc}}(\mathbb{R}^{d}\backslash\{\bm{0}\}),\ w\geq 0,\ w\ \text{is radial};\\ &\text{there exists}\ \epsilon_{0}\in(0,1)\text{ such that }w(\bm{x})>0\text{ for }0<|\bm{x}|\leq\epsilon_{0};\\ &\int_{\mathbb{R}^{d}}\min(1,|\bm{x}|)w(\bm{x})d\bm{x}=\int_{|\bm{x}|\leq 1}w(\bm{x})|\bm{x}|d\bm{x}+\int_{|\bm{x}|>1}w(\bm{x})d\bm{x}=:M_{w}^{1}+M_{w}^{2}<\infty.\end{aligned}\right.
Remark 2.1.

There are two typical types of kernels used in the literature that satisfy eq. 1. One type of kernel is those with compact supports, e.g., suppwBδ(𝟎)\text{supp}\ w\subset B_{\delta}(\bm{0}) for some δ>0\delta>0, where δ\delta represents the finite length of nonlocal interactions. Compactly supported kernels are used in peridynamics and the related studies, see e.g., [8, 18, 38, 49]. Another type of kernel has non-compact supports, e.g., w(𝐱)=C|𝐱|dαw(\bm{x})=C|\bm{x}|^{-d-\alpha} for α(0,1)\alpha\in(0,1), which relates to the Riesz fractional derivatives studied in [8, 46, 47]. For d=1d=1 with w(𝐱)=C|𝐱|1αw(\bm{x})=C|\bm{x}|^{-1-\alpha}, the nonlocal half-ball operators in this work directly relate to the Marchaud one-sided derivatives studied in [2, 34, 51]. Tempered fractional operators are discussed in [16, 44] where w(𝐱)=Ceλ|𝐱||𝐱|dαw(\bm{x})=Ce^{-\lambda|\bm{x}|}|\bm{x}|^{-d-\alpha} for λ>0\lambda>0 and α(0,1)\alpha\in(0,1).

Definition 2.1.

Given a measurable vector-valued function 𝐮:dN\bm{u}:\mathbb{R}^{d}\to\mathbb{R}^{N}, the action of nonlocal half-ball gradient operator 𝒢w𝛎\mathcal{G}^{\bm{\nu}}_{w} on 𝐮\bm{u} is defined as

(2) 𝒢w𝝂𝒖(𝒙):=limϵ0d\Bϵ(𝒙)χ𝝂(𝒚𝒙)𝒚𝒙|𝒚𝒙|(𝒖(𝒚)𝒖(𝒙))w(𝒚𝒙)𝑑𝒚,𝒙d,\mathcal{G}^{\bm{\nu}}_{w}\bm{u}(\bm{x}):=\lim_{\epsilon\to 0}\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\chi_{\bm{\nu}}(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes(\bm{u}(\bm{y})-\bm{u}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y},\quad\bm{x}\in\mathbb{R}^{d},

where 𝒢w𝛎𝐮:dd×N\mathcal{G}^{\bm{\nu}}_{w}\bm{u}:\mathbb{R}^{d}\to\mathbb{R}^{d\times N}. Given a measurable matrix-valued function 𝐯:dd×N\bm{v}:\mathbb{R}^{d}\to\mathbb{R}^{d\times N}, the action of nonlocal half-ball divergence operator 𝒟w𝛎\mathcal{D}^{\bm{\nu}}_{w} on 𝐯\bm{v} is defined as

(3) 𝒟w𝝂𝒗(𝒙):=limϵ0d\Bϵ(𝒙)χ𝝂(𝒚𝒙)[𝒚T𝒙T|𝒚𝒙|(𝒗(𝒚)𝒗(𝒙))]Tw(𝒚𝒙)𝑑𝒚,𝒙d,\mathcal{D}_{w}^{\bm{\nu}}\bm{v}(\bm{x}):=\lim_{\epsilon\to 0}\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\chi_{\bm{\nu}}(\bm{y}-\bm{x})\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\bm{v}(\bm{y})-\bm{v}(\bm{x}))\right]^{T}w(\bm{y}-\bm{x})d\bm{y},\quad\bm{x}\in\mathbb{R}^{d},

where 𝒟w𝛎𝐯:dN\mathcal{D}^{\bm{\nu}}_{w}\bm{v}:\mathbb{R}^{d}\to\mathbb{R}^{N}. If d=3d=3 and 𝐯:33\bm{v}:\mathbb{R}^{3}\to\mathbb{R}^{3}, then the action of nonlocal half-ball curl operator 𝒞w𝛎\mathcal{C}^{\bm{\nu}}_{w} on 𝐯\bm{v} is defined as

(4) 𝒞w𝝂𝒗(𝒙):=limϵ03\Bϵ(𝒙)χ𝝂(𝒚𝒙)𝒚𝒙|𝒚𝒙|×(𝒗(𝒚)𝒗(𝒙))w(𝒚𝒙)𝑑𝒚,𝒙3,\mathcal{C}_{w}^{\bm{\nu}}\bm{v}(\bm{x}):=\lim_{\epsilon\to 0}\int_{\mathbb{R}^{3}\backslash B_{\epsilon}(\bm{x})}\chi_{\bm{\nu}}(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\times(\bm{v}(\bm{y})-\bm{v}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y},\quad\bm{x}\in\mathbb{R}^{3},

where 𝒞w𝛎𝐯:33\mathcal{C}^{\bm{\nu}}_{w}\bm{v}:\mathbb{R}^{3}\to\mathbb{R}^{3}.

Remark 2.2.

Suppose suppwB1(𝟎)¯\mathrm{supp}\ w\subset\overline{B_{1}(\bm{0})}. For the affine function 𝐮(𝐱)=A𝐱+𝐛\bm{u}(\bm{x})=A\bm{x}+\bm{b} where AN×dA\in\mathbb{R}^{N\times d} and 𝐛N\bm{b}\in\mathbb{R}^{N}, it follows that

𝒢w𝝂𝒖(𝒙)=B1(𝟎)χ𝝂(𝒛)𝒛|𝒛|(A𝒛)w(𝒛)𝑑𝒛=(B1(𝟎)χ𝝂(𝒛)|𝒛|w(𝒛)𝒛|𝒛|𝒛|𝒛|𝑑𝒛)AT=Mw12dAT,𝒙d,\begin{split}\mathcal{G}^{\bm{\nu}}_{w}\bm{u}(\bm{x})&=\int_{B_{1}(\bm{0})}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}\otimes(A\bm{z})w(\bm{z})d\bm{z}\\ &=\left(\int_{B_{1}(\bm{0})}\chi_{\bm{\nu}}(\bm{z})|\bm{z}|w(\bm{z})\frac{\bm{z}}{|\bm{z}|}\otimes\frac{\bm{z}}{|\bm{z}|}d\bm{z}\right)A^{T}\\ &=\frac{M_{w}^{1}}{2d}A^{T},\quad\forall\bm{x}\in\mathbb{R}^{d},\end{split}

where we used

B1(𝟎)χ𝝂(𝒛)|𝒛|w(𝒛)𝒛|𝒛|𝒛|𝒛|𝑑𝒛=B1(𝟎)χ𝝂(𝒛)|𝒛|w(𝒛)𝒛|𝒛|𝒛|𝒛|𝑑𝒛(change of variable𝒛=𝒛)=12B1(𝟎)(χ𝝂(𝒛)+χ𝝂(𝒛))|𝒛|w(𝒛)𝒛|𝒛|𝒛|𝒛|𝑑𝒛=12B1(𝟎)|𝒛|w(𝒛)𝒛|𝒛|𝒛|𝒛|𝑑𝒛(χ𝝂(𝒛)+χ𝝂(𝒛)=1)=1201𝕊d1rdw(r)𝜼𝜼𝑑𝜼𝑑r=12(01rdw(r)𝑑r)1dωd1Id=Mw12dId.\begin{split}&\int_{B_{1}(\bm{0})}\chi_{\bm{\nu}}(\bm{z})|\bm{z}|w(\bm{z})\frac{\bm{z}}{|\bm{z}|}\otimes\frac{\bm{z}}{|\bm{z}|}d\bm{z}\\ =&\int_{B_{1}(\bm{0})}\chi_{-\bm{\nu}}(\bm{z})|\bm{z}|w(\bm{z})\frac{\bm{z}}{|\bm{z}|}\otimes\frac{\bm{z}}{|\bm{z}|}d\bm{z}\qquad(\text{change of variable}\ \bm{z}^{\prime}=-\bm{z})\\ =&\frac{1}{2}\int_{B_{1}(\bm{0})}\left(\chi_{\bm{\nu}}(\bm{z})+\chi_{-\bm{\nu}}(\bm{z})\right)|\bm{z}|w(\bm{z})\frac{\bm{z}}{|\bm{z}|}\otimes\frac{\bm{z}}{|\bm{z}|}d\bm{z}\\ =&\frac{1}{2}\int_{B_{1}(\bm{0})}|\bm{z}|w(\bm{z})\frac{\bm{z}}{|\bm{z}|}\otimes\frac{\bm{z}}{|\bm{z}|}d\bm{z}\qquad(\chi_{\bm{\nu}}(\bm{z})+\chi_{-\bm{\nu}}(\bm{z})=1)\\ =&\frac{1}{2}\int_{0}^{1}\int_{\mathbb{S}^{d-1}}r^{d}w(r)\bm{\eta}\otimes\bm{\eta}d\bm{\eta}dr\\ =&\frac{1}{2}\left(\int_{0}^{1}r^{d}w(r)dr\right)\frac{1}{d}\omega_{d-1}I_{d}\\ =&\frac{M_{w}^{1}}{2d}I_{d}.\end{split}

Here ωd1\omega_{d-1} is the surface area of (d1)(d-1)-sphere 𝕊d1\mathbb{S}^{d-1} and IdI_{d} is the d×dd\times d identity matrix.

One may further show that the localizations of these nonlocal operators are their local counterparts multiplied by a constant Mw12d\frac{M_{w}^{1}}{2d}, which justifies this definition. Specially, let wδ(𝐱)=1δd+1w(𝐱δ)w_{\delta}(\bm{x})=\frac{1}{\delta^{d+1}}w(\frac{\bm{x}}{\delta}) and uCc2(d)u\in C^{2}_{c}(\mathbb{R}^{d}), then by Taylor expansion one can prove that

𝒢wδ𝝂u(𝒙)Mw12du(𝒙),δ0,𝒙d,\mathcal{G}^{\bm{\nu}}_{w_{\delta}}u(\bm{x})\to\frac{M_{w}^{1}}{2d}\nabla u(\bm{x}),\quad\delta\to 0,\quad\forall\bm{x}\in\mathbb{R}^{d},

where u(𝐱)=(ujxi(𝐱))1id, 1jN\nabla u(\bm{x})=\left(\frac{\partial u_{j}}{\partial x_{i}}(\bm{x})\right)_{1\leq i\leq d,\,1\leq j\leq N} is the gradient matrix of uu, i.e., the transpose of the Jacobian matrix of uu. Similarly, for 𝐯Cc2(d;N)\bm{v}\in C^{2}_{c}(\mathbb{R}^{d};\mathbb{R}^{N}),

𝒟wδ𝝂𝒗(𝒙)Mw12d𝒗(𝒙),δ0,𝒙d,\mathcal{D}^{\bm{\nu}}_{w_{\delta}}\bm{v}(\bm{x})\to\frac{M_{w}^{1}}{2d}\nabla\cdot\bm{v}(\bm{x}),\quad\delta\to 0,\quad\forall\bm{x}\in\mathbb{R}^{d},

where the divergence vector of 𝐯\bm{v} given by 𝐯(𝐱)=(j=1dvjixj)1iN\nabla\cdot\bm{v}(\bm{x})=\left(\sum_{j=1}^{d}\frac{\partial v_{ji}}{\partial x_{j}}\right)_{1\leq i\leq N} is a column vector in N\mathbb{R}^{N}, and if d=N=3d=N=3,

𝒞wδ𝝂𝒗(𝒙)Mw12dcurl𝒗(𝒙),δ0,𝒙3.\mathcal{C}^{\bm{\nu}}_{w_{\delta}}\bm{v}(\bm{x})\to\frac{M_{w}^{1}}{2d}\mathrm{curl}\,\bm{v}(\bm{x}),\quad\delta\to 0,\quad\forall\bm{x}\in\mathbb{R}^{3}.

Note that in 2.1, the integrals are understood in the principal value sense. For smooth functions with compact support, the above integrals are just Lebesgue integrals, and moreover, the action of nonlocal operators yields smooth functions whose derivatives are LpL^{p} functions for 1p1\leq p\leq\infty. We summarize these results in the following lemma. The proof is similar to that of Proposition 1 in [15] and hence omitted.

Lemma 2.1.

Suppose that uCc(d)u\in C^{\infty}_{c}(\mathbb{R}^{d}) and 𝐯Cc(d;d)\bm{v}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}). Then 𝒢w𝛎u\mathcal{G}^{\bm{\nu}}_{w}u, 𝒟w𝛎𝐯\mathcal{D}_{w}^{\bm{\nu}}\bm{v} and 𝒞w𝛎𝐯\mathcal{C}_{w}^{\bm{\nu}}\bm{v} (d=3d=3) are CC^{\infty} functions with

(5) 𝒢w𝝂u(𝒙)=dχ𝝂(𝒚𝒙)𝒚𝒙|𝒚𝒙|(u(𝒚)u(𝒙))w(𝒚𝒙)𝑑𝒚,𝒙d,\mathcal{G}^{\bm{\nu}}_{w}u(\bm{x})=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}(u(\bm{y})-u(\bm{x}))w(\bm{y}-\bm{x})d\bm{y},\quad\bm{x}\in\mathbb{R}^{d},
(6) 𝒟w𝝂𝒗(𝒙)=dχ𝝂(𝒚𝒙)𝒚𝒙|𝒚𝒙|(𝒗(𝒚)𝒗(𝒙))w(𝒚𝒙)𝑑𝒚,𝒙d,\mathcal{D}_{w}^{\bm{\nu}}\bm{v}(\bm{x})=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\bm{v}(\bm{y})-\bm{v}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y},\quad\bm{x}\in\mathbb{R}^{d},

and if d=3d=3,

(7) 𝒞w𝝂𝒗(𝒙)=3χ𝝂(𝒚𝒙)𝒚𝒙|𝒚𝒙|×(𝒗(𝒚)𝒗(𝒙))w(𝒚𝒙)𝑑𝒚,𝒙3.\mathcal{C}_{w}^{\bm{\nu}}\bm{v}(\bm{x})=\int_{\mathbb{R}^{3}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\times(\bm{v}(\bm{y})-\bm{v}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y},\quad\bm{x}\in\mathbb{R}^{3}.

For p[1,]p\in[1,\infty] and multi-index αd\alpha\in\mathbb{N}^{d}, there is a constant CC depending on pp such that the following estimates hold:

(8) Dα𝒢w𝝂uLp(d;d)C(Mw1DαuLp(d;d)+Mw2DαuLp(d)),\|D^{\alpha}\mathcal{G}_{w}^{\bm{\nu}}u\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}\leq C\left(M_{w}^{1}\|\nabla D^{\alpha}u\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}+M_{w}^{2}\|D^{\alpha}u\|_{L^{p}(\mathbb{R}^{d})}\right),
(9) Dα𝒟w𝝂𝒗Lp(d)C(Mw1Dα𝒗Lp(d;d×d)+Mw2Dα𝒗Lp(d;d)),\|D^{\alpha}\mathcal{D}_{w}^{\bm{\nu}}\bm{v}\|_{L^{p}(\mathbb{R}^{d})}\leq C\left(M_{w}^{1}\|\nabla D^{\alpha}\bm{v}\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}+M_{w}^{2}\|D^{\alpha}\bm{v}\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}\right),

and if d=3d=3,

(10) Dα𝒞w𝝂𝒗Lp(3;3)C(Mw1Dα𝒗Lp(3;3×3)+Mw2Dα𝒗Lp(3;3)).\|D^{\alpha}\mathcal{C}_{w}^{\bm{\nu}}\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3})}\leq C\left(M_{w}^{1}\|\nabla D^{\alpha}\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3\times 3})}+M_{w}^{2}\|D^{\alpha}\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3})}\right).

If we replace smooth functions with compact support by W1,pW^{1,p} functions, then the action of nonlocal operators still yield LpL^{p} functions and the equalities (5)-(7) hold for a.e. xdx\in\mathbb{R}^{d}. The proof uses some ideas of Proposition 2.1(2) in [43] and is left to the appendix.

Lemma 2.2.

Let p[1,]p\in[1,\infty]. Then 𝒢w𝛎:W1,p(d)Lp(d;d)\mathcal{G}^{\bm{\nu}}_{w}:W^{1,p}(\mathbb{R}^{d})\to L^{p}(\mathbb{R}^{d};\mathbb{R}^{d}), 𝒟w𝛎:W1,p(d;d)Lp(d)\mathcal{D}^{\bm{\nu}}_{w}:W^{1,p}(\mathbb{R}^{d};\mathbb{R}^{d})\to L^{p}(\mathbb{R}^{d}) and 𝒞w𝛎:W1,p(3;3)Lp(3;3)\mathcal{C}^{\bm{\nu}}_{w}:W^{1,p}(\mathbb{R}^{3};\mathbb{R}^{3})\to L^{p}(\mathbb{R}^{3};\mathbb{R}^{3}) are bounded linear operators. Moreover, there exists a constant C>0C>0 depending on pp such that

(11) 𝒢w𝝂uLp(d;d)C(Mw1uLp(d;d)+Mw2uLp(d)),uW1,p(d),\|\mathcal{G}_{w}^{\bm{\nu}}u\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}\leq C\left(M_{w}^{1}\|\nabla u\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}+M_{w}^{2}\|u\|_{L^{p}(\mathbb{R}^{d})}\right),\quad u\in W^{1,p}(\mathbb{R}^{d}),
(12) 𝒟w𝝂𝒗Lp(d)C(Mw1𝒗Lp(d;d×d)+Mw2𝒗Lp(d;d)),𝒗W1,p(d;d),\|\mathcal{D}_{w}^{\bm{\nu}}\bm{v}\|_{L^{p}(\mathbb{R}^{d})}\leq C\left(M_{w}^{1}\|\nabla\bm{v}\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}+M_{w}^{2}\|\bm{v}\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}\right),\quad\bm{v}\in W^{1,p}(\mathbb{R}^{d};\mathbb{R}^{d}),

and if d=3d=3,

(13) 𝒞w𝝂𝒗Lp(3;3)C(Mw1𝒗Lp(3;3×3)+Mw2𝒗Lp(3;3)),𝒗W1,p(3;3).\|\mathcal{C}_{w}^{\bm{\nu}}\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3})}\leq C\left(M_{w}^{1}\|\nabla\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3\times 3})}+M_{w}^{2}\|\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3})}\right),\quad\bm{v}\in W^{1,p}(\mathbb{R}^{3};\mathbb{R}^{3}).

In addition, equalities (5)-(7) hold for a.e. xdx\in\mathbb{R}^{d}.

Analogous to the local operator, the integration by parts formula holds. Here we provide three types of integration by parts with proofs in the appendix. Note that the corresponding conditions in 2.1 (1)(2)(3) hold provided 𝒖Cc(d;N)\bm{u}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{N}), 𝒖Cc(d;d×N)\bm{u}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d\times N}) and 𝒖Cc(3;3)\bm{u}\in C^{\infty}_{c}(\mathbb{R}^{3};\mathbb{R}^{3}), respectively.

Proposition 2.1 (Nonlocal “half-ball” integration by parts).
  1. (1)

    Suppose 𝒖L1(d;N)\bm{u}\in L^{1}(\mathbb{R}^{d};\mathbb{R}^{N}), and w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(d×d)w(\bm{x}-\bm{y})\left|\bm{u}(\bm{x})-\bm{u}(\bm{y})\right|\in L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}). Then 𝒢w𝝂𝒖L1(d;d×N)\mathcal{G}^{\bm{\nu}}_{w}\bm{u}\in L^{1}(\mathbb{R}^{d};\mathbb{R}^{d\times N}) and for any 𝒗Cc1(d;d×N)\bm{v}\in C_{c}^{1}(\mathbb{R}^{d};\mathbb{R}^{d\times N}),

    (14) d𝒢w𝝂𝒖(𝒙):𝒗(𝒙)d𝒙=d𝒖(𝒙)𝒟w𝝂𝒗(𝒙)𝑑𝒙.\int_{\mathbb{R}^{d}}\mathcal{G}^{\bm{\nu}}_{w}\bm{u}(\bm{x}):\bm{v}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}\bm{u}(\bm{x})\cdot\mathcal{D}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}.
  2. (2)

    Suppose 𝒖L1(d;d×N)\bm{u}\in L^{1}(\mathbb{R}^{d};\mathbb{R}^{d\times N}), and w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(d×d)w(\bm{x}-\bm{y})\left|\bm{u}(\bm{x})-\bm{u}(\bm{y})\right|\in L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}). Then 𝒟w𝝂𝒖L1(d;N)\mathcal{D}^{\bm{\nu}}_{w}\bm{u}\in L^{1}(\mathbb{R}^{d};\mathbb{R}^{N}) and for any 𝒗Cc1(d;N)\bm{v}\in C_{c}^{1}(\mathbb{R}^{d};\mathbb{R}^{N}),

    (15) d𝒟w𝝂𝒖(𝒙)𝒗(𝒙)𝑑𝒙=d𝒖(𝒙):𝒢w𝝂𝒗(𝒙)d𝒙.\int_{\mathbb{R}^{d}}\mathcal{D}^{\bm{\nu}}_{w}\bm{u}(\bm{x})\cdot\bm{v}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}\bm{u}(\bm{x}):\mathcal{G}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}.
  3. (3)

    Let d=3d=3. Suppose 𝒖L1(3;3)\bm{u}\in L^{1}(\mathbb{R}^{3};\mathbb{R}^{3}), and w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(3×3)w(\bm{x}-\bm{y})\left|\bm{u}(\bm{x})-\bm{u}(\bm{y})\right|\in L^{1}(\mathbb{R}^{3}\times\mathbb{R}^{3}). Then 𝒞w𝝂𝒖L1(3;3)\mathcal{C}^{\bm{\nu}}_{w}\bm{u}\in L^{1}(\mathbb{R}^{3};\mathbb{R}^{3}) and for any 𝒗Cc1(3;3)\bm{v}\in C_{c}^{1}(\mathbb{R}^{3};\mathbb{R}^{3}),

    (16) 3𝒞w𝝂𝒖(𝒙)𝒗(𝒙)𝑑𝒙=3𝒖(𝒙)𝒞w𝝂𝒗(𝒙)𝑑𝒙.\int_{\mathbb{R}^{3}}\mathcal{C}^{\bm{\nu}}_{w}\bm{u}(\bm{x})\cdot\bm{v}(\bm{x})d\bm{x}=\int_{\mathbb{R}^{3}}\bm{u}(\bm{x})\cdot\mathcal{C}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}.
Remark 2.3.

As seen from the proof of 2.1 in the appendix, an equivalent definition of the divergence operator in eq. 3 is given as

𝒟w𝝂𝒗(𝒙)=limϵ0d\Bϵ(𝒙)[𝒚T𝒙T|𝒚𝒙|(χ𝝂(𝒚𝒙)𝒗(𝒚)+χ𝝂(𝒙𝒚)𝒗(𝒙))]Tw(𝒚𝒙)𝑑𝒚,\mathcal{D}_{w}^{\bm{\nu}}\bm{v}(\bm{x})=\lim_{\epsilon\to 0}\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}\left(\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{v}(\bm{x})\right)\right]^{T}w(\bm{y}-\bm{x})d\bm{y},

for 𝐱d\bm{x}\in\mathbb{R}^{d}.

We point out that nonlocal gradient, divergence and curl can be defined for complex-valued functions via eq. 2, eq. 3 and eq. 4, respectively, where the dot product in eq. 3 is understood as the inner product in d\mathbb{C}^{d}, i.e., 𝒛𝒘:=𝒛T𝒘¯\bm{z}\cdot\bm{w}:=\bm{z}^{T}\overline{\bm{w}} for 𝒛,𝒘d\bm{z},\bm{w}\in\mathbb{C}^{d}, and the cross product in eq. 4 is understood as the cross product in d\mathbb{C}^{d}. This extension will be useful in the proof of 4.2.

2.2. Distributional nonlocal operators

Previously, we defined nonlocal nonsymmetric operators in the principal value sense. It turns out that this notion is not enough to define nonlocal Sobolev spaces. Instead, we need the notion of distributional nonlocal gradient as the notion of weak derivative in the local setting. One way to define it is via its adjoint operator, i.e., nonlocal nonsymmetric divergence operator defined in the last subsection.

Following the idea in [43], we define the distributional nonlocal operators as follows.

Definition 2.2.

Let 1p1\leq p\leq\infty. Given 𝐮Lp(d;N)\bm{u}\in L^{p}(\mathbb{R}^{d};\mathbb{R}^{N}), we define the distributional nonlocal gradient 𝔊w𝛎𝐮(Cc(d;d×N))\mathfrak{G}_{w}^{\bm{\nu}}\bm{u}\in(C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d\times N}))^{\prime} as

(17) 𝔊w𝝂𝒖,ϕ:=d𝒖(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙,ϕCc(d;d×N).\langle\mathfrak{G}_{w}^{\bm{\nu}}\bm{u},\bm{\phi}\rangle:=-\int_{\mathbb{R}^{d}}\bm{u}(\bm{x})\cdot\mathcal{D}_{w}^{-\bm{\nu}}\bm{\phi}(\bm{x})d\bm{x},\quad\forall\bm{\phi}\in C_{c}^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d\times N}).

Given 𝐮Lp(d;d×N)\bm{u}\in L^{p}(\mathbb{R}^{d};\mathbb{R}^{d\times N}), we define the distributional nonlocal divergence 𝔇w𝛎𝐮(Cc(d;N))\mathfrak{D}_{w}^{\bm{\nu}}\bm{u}\in(C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{N}))^{\prime} as

(18) 𝔇w𝝂𝒖,ϕ:=d𝒖(𝒙):𝒢w𝝂ϕ(𝒙)d𝒙,ϕCc(d;N).\langle\mathfrak{D}_{w}^{\bm{\nu}}\bm{u},\bm{\phi}\rangle:=-\int_{\mathbb{R}^{d}}\bm{u}(\bm{x}):\mathcal{G}_{w}^{-\bm{\nu}}\bm{\phi}(\bm{x})d\bm{x},\quad\forall\bm{\phi}\in C_{c}^{\infty}(\mathbb{R}^{d};\mathbb{R}^{N}).

If d=3d=3 with 𝐮Lp(3;3)\bm{u}\in L^{p}(\mathbb{R}^{3};\mathbb{R}^{3}), we define the distributional nonlocal curl w𝛎𝐮(Cc(3;3))\mathfrak{C}_{w}^{\bm{\nu}}\bm{u}\in(C^{\infty}_{c}(\mathbb{R}^{3};\mathbb{R}^{3}))^{\prime} as

(19) w𝝂𝒖,ϕ:=3𝒖(𝒙)𝒞w𝝂ϕ(𝒙)𝑑𝒙,ϕCc(3;3).\langle\mathfrak{C}_{w}^{\bm{\nu}}\bm{u},\bm{\phi}\rangle:=\int_{\mathbb{R}^{3}}\bm{u}(\bm{x})\cdot\mathcal{C}_{w}^{-\bm{\nu}}\bm{\phi}(\bm{x})d\bm{x},\quad\forall\bm{\phi}\in C_{c}^{\infty}(\mathbb{R}^{3};\mathbb{R}^{3}).
Remark 2.4.

For uLp(d)u\in L^{p}(\mathbb{R}^{d}), 𝔊w𝛎u\mathfrak{G}_{w}^{\bm{\nu}}u is indeed a distribution as for any compact set KdK\subset\mathbb{R}^{d} and ϕCc(d;d)\bm{\phi}\in C_{c}^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d}) with support contained in KK,

|𝔊w𝝂u,ϕ|\displaystyle|\langle\mathfrak{G}_{w}^{\bm{\nu}}u,\bm{\phi}\rangle| uLp(d)𝒟w𝝂ϕLp(d)\displaystyle\leq\|u\|_{L^{p}(\mathbb{R}^{d})}\|\mathcal{D}_{w}^{-\bm{\nu}}\bm{\phi}\|_{L^{p^{\prime}}(\mathbb{R}^{d})}
C(Mw1ϕLp(d;d×d)+Mw2ϕLp(d;d))uLp(d)\displaystyle\leq C\left(M_{w}^{1}\|\nabla\bm{\phi}\|_{L^{p^{\prime}}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}+M_{w}^{2}\|\bm{\phi}\|_{L^{p^{\prime}}(\mathbb{R}^{d};\mathbb{R}^{d})}\right)\|u\|_{L^{p}(\mathbb{R}^{d})}
C|K|1p(Mw1ϕL(d;d×d)+Mw2ϕL(d;d))uLp(d),\displaystyle\leq C|K|^{\frac{1}{p^{\prime}}}\left(M_{w}^{1}\|\nabla\bm{\phi}\|_{L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}+M_{w}^{2}\|\bm{\phi}\|_{L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d})}\right)\|u\|_{L^{p}(\mathbb{R}^{d})},

where p=pp1p^{\prime}=\frac{p}{p-1} (p=p^{\prime}=\infty for p=1p=1 and p=1p^{\prime}=1 for p=p=\infty) and eq. 9 is used in the above inequalities. Similarly, it can be shown that 𝔇w𝛎𝐮\mathfrak{D}^{\bm{\nu}}_{w}\bm{u} and w𝛎𝐮\mathfrak{C}^{\bm{\nu}}_{w}\bm{u} are distributions using eq. 8 and eq. 10.

From the integration by parts formulas in 2.1, we immediately have the following results when the distributional operators 𝔊w𝝂𝒖\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}, 𝔇w𝝂𝒖\mathfrak{D}^{\bm{\nu}}_{w}\bm{u} and w𝝂𝒖\mathfrak{C}^{\bm{\nu}}_{w}\bm{u} coincide with 𝒢w𝝂𝒖\mathcal{G}^{\bm{\nu}}_{w}\bm{u}, 𝒟w𝝂𝒖\mathcal{D}^{\bm{\nu}}_{w}\bm{u} and 𝒞w𝝂𝒖\mathcal{C}^{\bm{\nu}}_{w}\bm{u}, respectively.

Corollary 2.1.
  1. (1)

    Suppose 𝒖L1(d;N)\bm{u}\in L^{1}(\mathbb{R}^{d};\mathbb{R}^{N}) and w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(d×d)w(\bm{x}-\bm{y})\left|\bm{u}(\bm{x})-\bm{u}(\bm{y})\right|\in L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}), then 𝒢w𝝂𝒖=𝔊w𝝂𝒖\mathcal{G}^{\bm{\nu}}_{w}\bm{u}=\mathfrak{G}^{\bm{\nu}}_{w}\bm{u} in L1(d;d×N)L^{1}(\mathbb{R}^{d};\mathbb{R}^{d\times N}).

  2. (2)

    Suppose 𝒖L1(d;d×N)\bm{u}\in L^{1}(\mathbb{R}^{d};\mathbb{R}^{d\times N}) and w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(d×d)w(\bm{x}-\bm{y})\left|\bm{u}(\bm{x})-\bm{u}(\bm{y})\right|\in L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}), then 𝒟w𝝂𝒖=𝔇w𝝂𝒖\mathcal{D}^{\bm{\nu}}_{w}\bm{u}=\mathfrak{D}^{\bm{\nu}}_{w}\bm{u} in L1(d;N)L^{1}(\mathbb{R}^{d};\mathbb{R}^{N}).

  3. (3)

    Suppose 𝒖L1(3;3)\bm{u}\in L^{1}(\mathbb{R}^{3};\mathbb{R}^{3}) and w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(3×3)w(\bm{x}-\bm{y})\left|\bm{u}(\bm{x})-\bm{u}(\bm{y})\right|\in L^{1}(\mathbb{R}^{3}\times\mathbb{R}^{3}), then 𝒞w𝝂𝒖=w𝝂𝒖\mathcal{C}^{\bm{\nu}}_{w}\bm{u}=\mathfrak{C}^{\bm{\nu}}_{w}\bm{u} in L1(3)L^{1}(\mathbb{R}^{3}).

2.3. Fourier symbols of nonlocal operators

In this subsection, we study the Fourier symbols of nonlocal operators defined in the previous subsection. These results will be used in the analysis in the subsequent sections.

Define

(20) 𝝀w𝝂(𝝃):=dχ𝝂(𝒛)𝒛|𝒛|w(𝒛)(e2πi𝝃𝒛1)𝑑𝒛,𝝃d.\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi}):=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}w(\bm{z})(e^{2\pi i\bm{\xi}\cdot\bm{z}}-1)d\bm{z},\quad\bm{\bm{\xi}}\in\mathbb{R}^{d}.

It is immediate that 𝝀w𝝂(𝝃)=𝝀w𝝂(𝝃)¯\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})=-\overline{\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})} for 𝝃d\bm{\bm{\xi}}\in\mathbb{R}^{d}. In fact, 𝝀w𝝂\bm{\lambda}_{w}^{\bm{\nu}} is the Fourier symbol of 𝒢w𝝂\mathcal{G}_{w}^{\bm{\nu}}, 𝒟w𝝂\mathcal{D}^{\bm{\nu}}_{w} and 𝒞w𝝂\mathcal{C}_{w}^{\bm{\nu}} in the sense described below. We now present this fact without proof since the proof is straightforward. Indeed, first prove the result for smooth functions with compact support and then use (11)-(13) for p=2p=2 and density of Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}) in H1(d)H^{1}(\mathbb{R}^{d}). Similar results can also be found in [38].

Lemma 2.3.

Let 𝐮H1(d;N)\bm{u}\in H^{1}(\mathbb{R}^{d};\mathbb{R}^{N}) and 𝐯H1(d;d×N)\bm{v}\in H^{1}(\mathbb{R}^{d};\mathbb{R}^{d\times N}). The Fourier transform of the nonlocal gradient operator 𝒢w𝛎\mathcal{G}_{w}^{\bm{\nu}} acting on 𝐮\bm{u} is given by

(21) (𝒢w𝝂𝒖)(𝝃)=𝝀w𝝂(𝝃)𝒖^(𝝃),𝝃d,\mathcal{F}(\mathcal{G}_{w}^{\bm{\nu}}\bm{u})(\bm{\xi})=\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\otimes\hat{\bm{u}}(\bm{\xi}),\quad\bm{\xi}\in\mathbb{R}^{d},

and the Fourier transform of the nonlocal divergence operator 𝒟w𝛎\mathcal{D}^{\bm{\nu}}_{w} acting on 𝐯\bm{v} is given by

(22) (𝒟w𝝂𝒗)(𝝃)=(𝝀w𝝂(𝝃)T𝒗^(𝝃))T,𝝃d.\mathcal{F}(\mathcal{D}^{\bm{\nu}}_{w}\bm{v})(\bm{\xi})=\left(\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})^{T}\hat{\bm{v}}(\bm{\xi})\right)^{T},\quad\bm{\xi}\in\mathbb{R}^{d}.

If, in particular, d=3d=3 and 𝐯H1(3;3)\bm{v}\in H^{1}(\mathbb{R}^{3};\mathbb{R}^{3}), then the Fourier transform of the nonlocal curl operator 𝒞w𝛎\mathcal{C}_{w}^{\bm{\nu}} acting on 𝐯\bm{v} is given by

(23) (𝒞w𝝂𝒗)(𝝃)=𝝀w𝝂(𝝃)×𝒗^(𝝃),𝝃3.\mathcal{F}(\mathcal{C}_{w}^{\bm{\nu}}\bm{v})(\bm{\xi})=\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\times\hat{\bm{v}}(\bm{\xi}),\quad\bm{\xi}\in\mathbb{R}^{3}.

Now we write out the real and imaginary part of 𝝀w𝝂(𝝃)\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi}) explicitly and show that the imaginary part is a scalar multiple of 𝝃\bm{\xi}. Moreover, the upper bound of 𝝀w𝝂(𝝃)\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi}) is linear in |𝝃||\bm{\xi}|. The proof of the following lemma is omitted since it follows from Lemma 2.3 and the last part of Theorem 2.4 in [38].

Lemma 2.4.

The Fourier symbol 𝛌w𝛎(𝛏)\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi}) can be expressed as

𝝀w𝝂(𝝃)=(𝝀w𝝂)(𝝃)+i(𝝀w𝝂)(𝝃),\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})=\Re(\bm{\lambda}_{w}^{\bm{\nu}})(\bm{\xi})+i\Im(\bm{\lambda}_{w}^{\bm{\nu}})(\bm{\xi}),

where

(24) (𝝀w𝝂)(𝝃)=dχ𝝂(𝒛)𝒛|𝒛|w(𝒛)(cos(2π𝝃𝒛)1)𝑑𝒛,\Re(\bm{\lambda}_{w}^{\bm{\nu}})(\bm{\xi})=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}w(\bm{z})(\cos(2\pi\bm{\xi}\cdot\bm{z})-1)d\bm{z},
(25) (𝝀w𝝂)(𝝃)=dχ𝝂(𝒛)𝒛|𝒛|w(𝒛)sin(2π𝝃𝒛)𝑑𝒛,\Im(\bm{\lambda}_{w}^{\bm{\nu}})(\bm{\xi})=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}w(\bm{z})\sin(2\pi\bm{\xi}\cdot\bm{z})d\bm{z},

and (𝛌w𝛎)(𝛏)=Λw(|𝛏|)𝛏|𝛏|\Im(\bm{\lambda}_{w}^{\bm{\nu}})(\bm{\xi})=\Lambda_{w}(|\bm{\xi}|)\frac{\bm{\xi}}{|\bm{\xi}|} with

(26) Λw(|𝝃|)=12dw(𝒛)|𝒛|z1sin(2π|𝝃|z1)𝑑𝒛.\Lambda_{w}(|\bm{\xi}|)=\frac{1}{2}\int_{\mathbb{R}^{d}}\frac{w(\bm{z})}{|\bm{z}|}z_{1}\sin(2\pi|\bm{\xi}|z_{1})d\bm{z}.

Moreover,

(27) |𝝀w𝝂(𝝃)|2(2πMw1|𝝃|+Mw2),𝝃d.\left|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\right|\leq\sqrt{2}\left(2\pi M_{w}^{1}|\bm{\xi}|+M_{w}^{2}\right),\quad\forall\bm{\bm{\xi}}\in\mathbb{R}^{d}.

In the following, we present two other observations of the Fourier symbol 𝝀w𝝂\bm{\lambda}_{w}^{\bm{\nu}} that are useful in Section 4. The first result concerns the positivity of |𝝀w𝝂||\bm{\lambda}_{w}^{\bm{\nu}}| away from the origin, and the second result asserts that 𝝀w𝝂\bm{\lambda}_{w}^{\bm{\nu}} is a smooth function.

Proposition 2.2.

For every d×dd\times d orthogonal matrix RR,

(28) 𝝀wR𝝂(𝝃)=R𝝀w𝝂(RT𝝃),𝝃𝟎.\bm{\lambda}_{w}^{R\bm{\nu}}(\bm{\xi})=R\bm{\lambda}_{w}^{\bm{\nu}}(R^{T}\bm{\xi}),\quad\forall\bm{\xi}\neq\bm{0}.

The same formula holds for both (𝛌w𝛎)(𝛏)\Re(\bm{\lambda}_{w}^{\bm{\nu}})(\bm{\xi}) and (𝛌w𝛎)(𝛏)\Im(\bm{\lambda}_{w}^{\bm{\nu}})(\bm{\xi}). Consequently,

(29) |𝝀w𝝂(𝝃)|>0,𝝃𝟎.\left|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\right|>0,\quad\forall\bm{\xi}\neq\bm{0}.
Proof.

Equation 28 can be easily seen from a change of variable. For a fixed unit vector 𝝂d\bm{\nu}\in\mathbb{R}^{d}, there exists an orthogonal matrix R𝝂R_{\bm{\nu}} such that 𝝂=R𝝂𝒆1\bm{\nu}=R_{\bm{\nu}}\bm{e}_{1}. By (28), for 𝝃𝟎\bm{\xi}\neq\bm{0},

|𝝀w𝝂(𝝃)|\displaystyle|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})| =|R𝝂𝝀w𝒆1(R𝝂T𝝃)|=|𝝀w𝒆1(R𝝂T𝝃)||𝝀w𝒆1(R𝝂T𝝃)𝒆1||(𝝀w𝒆1(R𝝂T𝝃)𝒆1)|\displaystyle=|R_{\bm{\nu}}{\bm{\lambda}}_{w}^{\bm{e}_{1}}(R_{\bm{\nu}}^{T}\bm{\xi})|=|{\bm{\lambda}}_{w}^{\bm{e}_{1}}(R_{\bm{\nu}}^{T}\bm{\xi})|\geq|{\bm{\lambda}}_{w}^{\bm{e}_{1}}(R_{\bm{\nu}}^{T}\bm{\xi})\cdot\bm{e}_{1}|\geq|\Re({\bm{\lambda}}_{w}^{\bm{e}_{1}}(R_{\bm{\nu}}^{T}\bm{\xi})\cdot\bm{e}_{1})|
={z1>0}z1|𝒛|w(𝒛)(1cos(2π(R𝝂T𝝃𝒛))d𝒛>0,\displaystyle=\int_{\{z_{1}>0\}}\frac{z_{1}}{|\bm{z}|}w(\bm{z})\left(1-\mathrm{cos}\left(2\pi(R_{\bm{\nu}}^{T}\bm{\xi}\cdot\bm{z}\right)\right)d\bm{z}>0,

where the last inequality holds because the integrand is nonnegative and the set

{𝒛d:z1>0,(R𝝂T𝝃)𝒛}\{\bm{z}\in\mathbb{R}^{d}:z_{1}>0,\ (R_{\bm{\nu}}^{T}\bm{\xi})\cdot\bm{z}\in\mathbb{Z}\}

is a set of measure zero in d\mathbb{R}^{d}. Thus, (29) holds. ∎

Proposition 2.3.

Suppose the kernel function ww satisfies eq. 1, and in addition, the support of ww is a compact set in d\mathbb{R}^{d}. Then the Fourier symbol 𝛌w𝛎C(d;d)\bm{\lambda}_{w}^{\bm{\nu}}\in C^{\infty}(\mathbb{R}^{d};\mathbb{C}^{d}).

Proof.

Notice that for any multi-index 𝜸\bm{\gamma} with |𝜸|>0|\bm{\gamma}|>0,

D𝜸(𝝀w𝝂)(𝝃)=dχ𝝂(𝒛)𝒛|𝒛|w(𝒛)(2πi𝒛)𝜸e2πi𝝃𝒛𝑑𝒛.D^{\bm{\gamma}}({\bm{\lambda}}_{w}^{\bm{\nu}})(\bm{\xi})=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}w(\bm{z})(2\pi i\bm{z})^{\bm{\gamma}}e^{2\pi i\bm{\xi}\cdot\bm{z}}d\bm{z}.

Since ww is a compactly supported kernel function, the integrand on the right-hand side of the above equation can be controlled by the integrable function |𝒛|w(𝒛)|\bm{z}|w(\bm{z}). Hence, 𝝀w𝝂C(d;d){\bm{\lambda}}_{w}^{\bm{\nu}}\in C^{\infty}(\mathbb{R}^{d};\mathbb{C}^{d}). ∎

2.4. Nonlocal vector identities for smooth functions

In this subsection, we present some nonlocal vector identities for smooth functions with compact support. These results will be generalized for a larger class of functions in Section 3 and become crucial for applications in Section 6.

The following lemma shows that 𝒞w𝝂𝒢w𝝂=0\mathcal{C}^{\bm{\nu}}_{w}\circ\mathcal{G}^{\bm{\nu}}_{w}=0 and 𝒟w𝝂𝒞w𝝂=0\mathcal{D}^{\bm{\nu}}_{w}\circ\mathcal{C}^{\bm{\nu}}_{w}=0, analogous to curlgrad=0\mathrm{curl}\circ\mathrm{grad}=0 and divcurl=0\mathrm{div}\circ\mathrm{curl}=0 in the local setting.

Lemma 2.5.

Let d=3d=3. Then for uCc(3)u\in C^{\infty}_{c}(\mathbb{R}^{3}) and 𝐯Cc(3;3)\bm{v}\in C^{\infty}_{c}(\mathbb{R}^{3};\mathbb{R}^{3}),

(30) 𝒞w𝝂𝒢w𝝂u(𝒙)=0,a.e.𝒙3,\mathcal{C}^{\bm{\nu}}_{w}\mathcal{G}^{\bm{\nu}}_{w}u(\bm{x})=0,\quad\text{a.e.}\ \bm{x}\in\mathbb{R}^{3},

and

(31) 𝒟w𝝂𝒞w𝝂𝒗(𝒙)=0,a.e.𝒙3.\mathcal{D}^{\bm{\nu}}_{w}\mathcal{C}^{\bm{\nu}}_{w}\bm{v}(\bm{x})=0,\quad\text{a.e.}\ \bm{x}\in\mathbb{R}^{3}.
Proof.

First note that by 2.1, 𝒢w𝝂uH1(3;3)\mathcal{G}^{\bm{\nu}}_{w}u\in H^{1}(\mathbb{R}^{3};\mathbb{R}^{3}) and 𝒞w𝝂𝒗H1(3;3)\mathcal{C}^{\bm{\nu}}_{w}\bm{v}\in H^{1}(\mathbb{R}^{3};\mathbb{R}^{3}). Then the conditions for 2.3 hold and one can apply the Fourier transform to L2L^{2} functions 𝒞w𝝂𝒢w𝝂u\mathcal{C}^{\bm{\nu}}_{w}\mathcal{G}^{\bm{\nu}}_{w}u and 𝒟w𝝂𝒞w𝝂𝒗\mathcal{D}^{\bm{\nu}}_{w}\mathcal{C}^{\bm{\nu}}_{w}\bm{v}. By 2.3, eq. 30 and eq. 31 follows from

𝝀w𝝂×(𝝀w𝝂u^)=0\bm{\lambda}_{w}^{\bm{\nu}}\times(\bm{\lambda}_{w}^{\bm{\nu}}\hat{u})=0

and

(𝝀w𝝂)T(𝝀w𝝂×𝒗^)=0(\bm{\lambda}_{w}^{\bm{\nu}})^{T}(\bm{\lambda}_{w}^{\bm{\nu}}\times\hat{\bm{v}})=0

respectively. ∎

Next, we show two nonlocal vector identities analogous to the following vector calculus identities in local setting111In eq. 32, the two types of curls in 2D are defined as Curl𝒗:=v2x1v1x2 and Curlϕ:=(ϕx2,ϕx1)T,\mathrm{Curl}\ \bm{v}:=\frac{\partial v_{2}}{\partial x_{1}}-\frac{\partial v_{1}}{\partial x_{2}}\text{ and }\textbf{Curl}\ \phi:=\left(\frac{\partial\phi}{\partial x_{2}},-\frac{\partial\phi}{\partial x_{1}}\right)^{T}, for a vector field 𝒗\bm{v} and a scalar field ϕ\phi. In eq. 33, the curl of a vector field 𝒗\bm{v} in 3D is defined as Curl𝒗:=(v3x2v2x3,v1x3v1x3,v2x1v1x2)T.\textbf{Curl}\ \bm{v}:=\left(\frac{\partial v_{3}}{\partial x_{2}}-\frac{\partial v_{2}}{\partial x_{3}},\frac{\partial v_{1}}{\partial x_{3}}-\frac{\partial v_{1}}{\partial x_{3}},\frac{\partial v_{2}}{\partial x_{1}}-\frac{\partial v_{1}}{\partial x_{2}}\right)^{T}. :

(32) (𝒗)=(𝒗)CurlCurl𝒗,d=2;\nabla\cdot(\nabla\bm{v})=\nabla(\nabla\cdot\bm{v})-\textbf{Curl}\ \mathrm{Curl}\ \bm{v},\quad d=2;
(33) (𝒗)=(𝒗)CurlCurl𝒗,d=3.\nabla\cdot(\nabla\bm{v})=\nabla(\nabla\cdot\bm{v})-\textbf{Curl}\ \textbf{Curl}\ \bm{v},\quad d=3.
Lemma 2.6.

For 𝐮Cc(2;2)\bm{u}\in C^{\infty}_{c}(\mathbb{R}^{2};\mathbb{R}^{2}),

(34) 𝒟w𝝂𝒢w𝝂𝒖=𝒢w𝝂𝒟w𝝂𝒖(0110)𝒢w𝝂𝒟w𝝂[(0110)𝒖].\mathcal{D}_{w}^{-\bm{\nu}}\mathcal{G}_{w}^{\bm{\nu}}\bm{u}=\mathcal{G}_{w}^{\bm{\nu}}\mathcal{D}_{w}^{-\bm{\nu}}\bm{u}-\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\mathcal{G}_{w}^{-\bm{\nu}}\mathcal{D}_{w}^{\bm{\nu}}\left[\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\bm{u}\right].
Proof.

As remarked at the beginning of the proof of 2.5, it is valid to apply the Fourier transform. Applying the Fourier transform and 2.3, the left hand side of eq. 34 becomes |𝝀w𝝂(𝝃)|2𝒖^(𝝃)-|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})|^{2}\hat{\bm{u}}(\bm{\xi}) and the right hand side becomes

𝝀w𝝂(𝝃)𝝀w𝝂(𝝃)T𝒖^(𝝃)(0110)𝝀w𝝂(𝝃)𝝀w𝝂(𝝃)T(0110)𝒖^(𝝃)\displaystyle\quad\ \bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})^{T}\hat{\bm{u}}(\bm{\xi})-\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})^{T}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\hat{\bm{u}}(\bm{\xi})
=[(λ1(𝝃)λ1(𝝃)¯λ1(𝝃)λ2(𝝃)¯λ2(𝝃)λ1(𝝃)¯λ2(𝝃)λ2(𝝃)¯)\displaystyle=\bigg{[}-\begin{pmatrix}\lambda_{1}(\bm{\xi})\overline{\lambda_{1}(\bm{\xi})}&\lambda_{1}(\bm{\xi})\overline{\lambda_{2}(\bm{\xi})}\\ \lambda_{2}(\bm{\xi})\overline{\lambda_{1}(\bm{\xi})}&\lambda_{2}(\bm{\xi})\overline{\lambda_{2}(\bm{\xi})}\end{pmatrix}
+(0110)(λ1(𝝃)¯λ1(𝝃)λ1(𝝃)¯λ2(𝝃)λ2(𝝃)¯λ1(𝝃)λ2(𝝃)¯λ2(𝝃))(0110)]𝒖^(𝝃)\displaystyle\qquad\qquad\qquad\quad+\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\begin{pmatrix}\overline{\lambda_{1}(\bm{\xi})}\lambda_{1}(\bm{\xi})&\overline{\lambda_{1}(\bm{\xi})}\lambda_{2}(\bm{\xi})\\ \overline{\lambda_{2}(\bm{\xi})}\lambda_{1}(\bm{\xi})&\overline{\lambda_{2}(\bm{\xi})}\lambda_{2}(\bm{\xi})\end{pmatrix}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\bigg{]}\hat{\bm{u}}(\bm{\xi})
=[(λ1(𝝃)λ1(𝝃)¯λ1(𝝃)λ2(𝝃)¯λ2(𝝃)λ1(𝝃)¯λ2(𝝃)λ2(𝝃)¯)+(λ2(𝝃)¯λ2(𝝃)λ2(𝝃)¯λ1(𝝃)λ1(𝝃)¯λ2(𝝃)λ1(𝝃)¯λ1(𝝃))]𝒖^(𝝃)\displaystyle=\left[-\begin{pmatrix}\lambda_{1}(\bm{\xi})\overline{\lambda_{1}(\bm{\xi})}&\lambda_{1}(\bm{\xi})\overline{\lambda_{2}(\bm{\xi})}\\ \lambda_{2}(\bm{\xi})\overline{\lambda_{1}(\bm{\xi})}&\lambda_{2}(\bm{\xi})\overline{\lambda_{2}(\bm{\xi})}\end{pmatrix}+\begin{pmatrix}-\overline{\lambda_{2}(\bm{\xi})}\lambda_{2}(\bm{\xi})&\overline{\lambda_{2}(\bm{\xi})}\lambda_{1}(\bm{\xi})\\ \overline{\lambda_{1}(\bm{\xi})}\lambda_{2}(\bm{\xi})&-\overline{\lambda_{1}(\bm{\xi})}\lambda_{1}(\bm{\xi})\end{pmatrix}\right]\hat{\bm{u}}(\bm{\xi})
=|𝝀w𝝂(𝝃)|2𝒖^(𝝃).\displaystyle=-|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})|^{2}\hat{\bm{u}}(\bm{\xi}).

Therefore, eq. 34 holds for 𝒖Cc(2;2)\bm{u}\in C_{c}^{\infty}(\mathbb{R}^{2};\mathbb{R}^{2}). ∎

Lemma 2.7.

For 𝐮Cc(3;3)\bm{u}\in C^{\infty}_{c}(\mathbb{R}^{3};\mathbb{R}^{3}),

(35) 𝒟w𝝂𝒢w𝝂𝒖=𝒢w𝝂𝒟w𝝂𝒖𝒞w𝝂𝒞w𝝂𝒖.\mathcal{D}_{w}^{-\bm{\nu}}\mathcal{G}_{w}^{\bm{\nu}}\bm{u}=\mathcal{G}_{w}^{\bm{\nu}}\mathcal{D}_{w}^{-\bm{\nu}}\bm{u}-\mathcal{C}_{w}^{-\bm{\nu}}\mathcal{C}_{w}^{\bm{\nu}}\bm{u}.
Proof.

Applying the Fourier transform to eq. 35 and using 2.3 yield

(𝒢w𝝂𝒟w𝝂𝒖𝒞w𝝂𝒞w𝝂𝒖)(𝝃)\displaystyle\quad\ \mathcal{F}(\mathcal{G}_{w}^{\bm{\nu}}\mathcal{D}_{w}^{-\bm{\nu}}\bm{u}-\mathcal{C}_{w}^{-\bm{\nu}}\mathcal{C}_{w}^{\bm{\nu}}\bm{u})(\bm{\xi})
=𝝀w𝝂(𝝃)𝝀w𝝂(𝝃)T𝒖^(𝝃)𝝀w𝝂(𝝃)×(𝝀w𝝂(𝝃)×𝒖^(𝝃))\displaystyle=\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})^{T}\hat{\bm{u}}(\bm{\xi})-\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})\times(\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\times\hat{\bm{u}}(\bm{\xi}))
=𝝀w𝝂(𝝃)𝝀w𝝂(𝝃)T𝒖^(𝝃)(𝝀w𝝂(𝝃)T𝒖^(𝝃))𝝀w𝝂(𝝃)+𝝀w𝝂(𝝃)T𝝀w𝝂(𝝃)𝒖^(𝝃)\displaystyle=\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})^{T}\hat{\bm{u}}(\bm{\xi})-(\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})^{T}\hat{\bm{u}}(\bm{\xi}))\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})+\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})^{T}\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\hat{\bm{u}}(\bm{\xi})
=|𝝀w𝝂(𝝃)|2𝒖^(𝝃)=(𝒟w𝝂(𝒢w𝝂𝒖))(𝝃),\displaystyle=-|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})|^{2}\hat{\bm{u}}(\bm{\xi})=\mathcal{F}(\mathcal{D}_{w}^{-\bm{\nu}}(\mathcal{G}_{w}^{\bm{\nu}}\bm{u}))(\bm{\xi}),

where we used 𝝀w𝝂(𝝃)=𝝀w𝝂(𝝃)¯\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})=-\overline{\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})}. ∎

3. Nonlocal Sobolev-type spaces

In this section, we define the nonlocal Sobolev-type spaces in which we prove the Poincaré inequality. The notion is defined via the distributional nonlocal gradient introduced in the previous section, motivated by the definition of classical Sobolev spaces. A similar notion was introduced in [12] for fractional gradient. For simplicity, we only consider the case p=2p=2, while the definitions and results in Section 3.1 below can be extended to a general p[1,)p\in[1,\infty).

3.1. Definitions and properties of nonlocal Sobolev-type spaces

For the rest of the paper, we adopt the convention that a domain is an open connected set (not necessarily bounded). Let Ωd\Omega\subset\mathbb{R}^{d} be a domain and N+N\in\mathbb{Z}^{+} a positive integer. Given a kernel function ww satisfying eq. 1 and a unit vector 𝝂d\bm{\nu}\in\mathbb{R}^{d}, define the associated energy space 𝒮w𝝂(Ω;N)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}) by

(36) 𝒮w𝝂(Ω;N):={𝒖L2(d;N):𝒖=𝟎 a.e. on d\Ω,𝔊w𝝂𝒖L2(d;d×N)},\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}):=\{\bm{u}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{N}):\bm{u}=\bm{0}\text{ a.e. on }\mathbb{R}^{d}\backslash\Omega,\;\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times N})\},

equipped with norm

𝒖𝒮w𝝂(Ω;N):=(𝒖L2(d;N)2+𝔊w𝝂𝒖L2(d;d×N)2)1/2,\|\bm{u}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N})}:=\left(\|\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{N})}^{2}+\|\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times N})}^{2}\right)^{1/2},

as well as the corresponding inner product. For any Ωd\Omega\subset\mathbb{R}^{d}, it is not hard to see that 𝒮w𝝂(Ω;N)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}) is a closed subspace of 𝒮w𝝂(d;N)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d};\mathbb{R}^{N}). When N=1N=1, we simply denote 𝒮w𝝂(Ω):=𝒮w𝝂(Ω;)\mathcal{S}_{w}^{\bm{\nu}}(\Omega):=\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}). Notice that any function in 𝒮w𝝂(Ω;N)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}) is a vector field where each component of it is a function in 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega). For the rest of this section, we will show 𝒮w𝝂(Ω;N)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}) is a separable Hilbert space for certain domain Ω\Omega. Since functions in 𝒮w𝝂(Ω;N)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}) can be understood componentwise as functions in 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega), we will work with 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) for the rest of this subsection and the following results also hold for 𝒮w𝝂(Ω;N)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}) where N+N\in\mathbb{Z}^{+}. The results of this subsection can also be easily extended to a general p[1,)p\in[1,\infty).

Remark 3.1.

𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) is a nonlocal analogue of the Sobolev space H01(Ω)H_{0}^{1}(\Omega). If the kernel function ww has compact support, e.g., suppwBδ(𝟎)\text{supp}\ w\subset B_{\delta}(\bm{0}) for δ>0\delta>0, then 𝔊w𝛎u\mathfrak{G}_{w}^{\bm{\nu}}u vanishes outside Ωδ:={𝐱d:dist(𝐱,Ω)<δ}\Omega_{\delta}:=\{\bm{x}\in\mathbb{R}^{d}:\text{dist}(\bm{x},\Omega)<\delta\}. In this case, we may equivalently define 𝒮w𝛎(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) as functions in L2(Ω2δ)L^{2}(\Omega_{2\delta}) that vanish on Ω2δ\Ω\Omega_{2\delta}\backslash\Omega with 𝔊w𝛎uL2(Ωδ;d)\mathfrak{G}_{w}^{\bm{\nu}}u\in L^{2}(\Omega_{\delta};\mathbb{R}^{d}).

Theorem 3.1.

Let Ωd\Omega\subset\mathbb{R}^{d} be a domain. The function space 𝒮w𝛎(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) is a Hilbert space.

Proof.

It suffices to prove that 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) is complete. Let {uk}k\{u_{k}\}_{k\in\mathbb{N}} be a Cauchy sequence in 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}). Since {uk}k\{u_{k}\}_{k\in\mathbb{N}} is a Cauchy sequence in L2(d)L^{2}(\mathbb{R}^{d}), there exists uL2(d)u\in L^{2}(\mathbb{R}^{d}) such that ukuu_{k}\to u in L2(d)L^{2}(\mathbb{R}^{d}) and 𝒗L2(d;d)\bm{v}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}) such that 𝔊w𝝂uk𝒗\mathfrak{G}^{\bm{\nu}}_{w}u_{k}\to\bm{v} in L2(d;d)L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). Now we show 𝔊w𝝂u=𝒗\mathfrak{G}^{\bm{\nu}}_{w}u=\bm{v} in the sense of distributions. By definition, for any ϕCc(d;d)\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}), it suffices to show

(37) du(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙=d𝒗(𝒙)ϕ(𝒙)𝑑𝒙.-\int_{\mathbb{R}^{d}}u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x}=\int_{\mathbb{R}^{d}}\bm{v}(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x}.

For kk\in\mathbb{N}, we have

(38) duk(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙=d𝔊w𝝂uk(𝒙)ϕ(𝒙)𝑑𝒙.-\int_{\mathbb{R}^{d}}u_{k}(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x}=\int_{\mathbb{R}^{d}}\mathfrak{G}^{\bm{\nu}}_{w}u_{k}(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x}.

Since ϕCc(d;d)\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}), by 2.1, we know 𝒟w𝝂ϕL2(d)\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}\in L^{2}(\mathbb{R}^{d}). Then taking kk to infinity in (38) yields (37). Thus, 𝔊w𝝂u=𝒗L2(d;d)\mathfrak{G}^{\bm{\nu}}_{w}u=\bm{v}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}) and ukuu_{k}\to u in 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}). Hence, 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) is a Hilbert space. Since 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) is a closed subspace of 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}), the normed space 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) is also complete. ∎

We next present a density result on 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) which is crucial in many applications. If Ωd\Omega\neq\mathbb{R}^{d}, the density result holds for domains that are bounded with continuous boundaries or epigraphs. We say Ω\Omega is an epigraph if there exists a continuous function ζ:d1\zeta:\mathbb{R}^{d-1}\to\mathbb{R} such that (up to a rigid motion),

Ω={𝒙=(𝒙,xd)d|xd>ζ(𝒙)}.\Omega=\{\bm{x}=(\bm{x}^{\prime},x_{d})\in\mathbb{R}^{d}\,|\,x_{d}>\zeta(\bm{x}^{\prime})\}.

If Ω\Omega is a bounded domain with a continuous boundary, then its boundary can be covered by finitely many balls where each patch is characterized by an epigraph.

Theorem 3.2.

Let Ω\Omega be a bounded domain with a continuous boundary, an epigraph, or d\mathbb{R}^{d}. Let Cc(Ω)C_{c}^{\infty}(\Omega) denote the space of smooth functions defined on d\mathbb{R}^{d} with compact support contained in Ω\Omega. Then Cc(Ω)C_{c}^{\infty}(\Omega) is dense in 𝒮w𝛎(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega).

The main ingredients of the proof of 3.2 are several lemmas stated below about cut-off, translation and mollification in nonlocal Sobolev spaces which we present in the following. First of all, a generalized ‘product rule’ for the nonlocal operators is useful.

Proposition 3.1.

For φCc(d)\varphi\in C^{\infty}_{c}(\mathbb{R}^{d}) and ϕCc(d;d)\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}),

(39) 𝒟w𝝂(φϕ)=φ𝒟w𝝂ϕ+𝒢w𝝂φϕ+S(φ,ϕ),\mathcal{D}^{-\bm{\nu}}_{w}(\varphi\bm{\phi})=\varphi\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}+\mathcal{G}^{\bm{\nu}}_{w}\varphi\cdot\bm{\phi}+S(\varphi,\bm{\phi}),

where S(φ,ϕ):dS(\varphi,\bm{\phi}):\mathbb{R}^{d}\to\mathbb{R} is a function given by

(40) S(φ,ϕ)(𝒙):=d𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)ϕ(𝒚)χ𝝂(𝒚𝒙)ϕ(𝒙))(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚.S(\varphi,\bm{\phi})(\bm{x}):=\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})-\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x}))(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}.

Similarly,

𝒢w𝝂(φψ)=φ𝒢w𝝂ψ+ψ𝒢w𝝂φ+S𝒢(φ,ψ),φ,ψCc(d),\mathcal{G}^{\bm{\nu}}_{w}(\varphi\psi)=\varphi\mathcal{G}^{\bm{\nu}}_{w}\psi+\psi\mathcal{G}^{\bm{\nu}}_{w}\varphi+S_{\mathcal{G}}(\varphi,\psi),\quad\forall\varphi,\psi\in C^{\infty}_{c}(\mathbb{R}^{d}),

where

S𝒢(φ,ψ)(𝒙):=dχ𝝂(𝒛)𝒛|𝒛|(φ(𝒙+𝒛)φ(𝒙))(ψ(𝒙+𝒛)ψ(𝒙))w(𝒛)𝑑𝒛,𝒙d.S_{\mathcal{G}}(\varphi,\psi)(\bm{x}):=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}(\varphi(\bm{x}+\bm{z})-\varphi(\bm{x}))(\psi(\bm{x}+\bm{z})-\psi(\bm{x}))w(\bm{z})d\bm{z},\quad\bm{x}\in\mathbb{R}^{d}.
Proof.

We only prove the produce rule for 𝒟w𝝂\mathcal{D}^{-\bm{\nu}}_{w} as the product rule for 𝒢w𝝂\mathcal{G}^{\bm{\nu}}_{w} is similar and simpler. First note that the function S(φ,ϕ)S(\varphi,\bm{\phi}) is well-defined with the pointwise estimate

(41) |S(φ,ϕ)(𝒙)|d2ϕL(d;d)2φW1,(d)min(1,|𝒚𝒙|)w(𝒚𝒙)𝑑𝒚<.|S(\varphi,\bm{\phi})(\bm{x})|\leq\int_{\mathbb{R}^{d}}2\|\bm{\phi}\|_{L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d})}\cdot 2\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}\min(1,|\bm{y}-\bm{x}|)w(\bm{y}-\bm{x})d\bm{y}<\infty.

Observe that for 𝒙,𝒚d\bm{x},\bm{y}\in\mathbb{R}^{d} and ϵ>0\epsilon>0,

χ[ϵ,)(|𝒚𝒙|){χ𝝂(𝒙𝒚)φ(𝒚)ϕ(𝒚)+χ𝝂(𝒚𝒙)φ(𝒙)ϕ(𝒙)}\displaystyle\ \quad\chi_{[\epsilon,\infty)}(|\bm{y}-\bm{x}|)\{\chi_{\bm{\nu}}(\bm{x}-\bm{y})\varphi(\bm{y})\bm{\phi}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\varphi(\bm{x})\bm{\phi}(\bm{x})\}
=χ[ϵ,)(|𝒚𝒙|){[χ𝝂(𝒙𝒚)ϕ(𝒚)+χ𝝂(𝒚𝒙)ϕ(𝒙)]φ(𝒙)+χ𝝂(𝒙𝒚)ϕ(𝒚)(φ(𝒚)φ(𝒙))}\displaystyle=\leavevmode\resizebox{469.75499pt}{}{$\chi_{[\epsilon,\infty)}(|\bm{y}-\bm{x}|)\{[\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x})]\varphi(\bm{x})+\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})(\varphi(\bm{y})-\varphi(\bm{x}))\}$}
=χ[ϵ,)(|𝒚𝒙|){[χ𝝂(𝒙𝒚)ϕ(𝒚)+χ𝝂(𝒚𝒙)ϕ(𝒙)]φ(𝒙)+χ𝝂(𝒚𝒙)ϕ(𝒙)(φ(𝒚)φ(𝒙))\displaystyle=\leavevmode\resizebox{469.75499pt}{}{$\chi_{[\epsilon,\infty)}(|\bm{y}-\bm{x}|)\{[\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x})]\varphi(\bm{x})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x})(\varphi(\bm{y})-\varphi(\bm{x}))$}
+[χ𝝂(𝒙𝒚)ϕ(𝒚)χ𝝂(𝒚𝒙)ϕ(𝒙)](φ(𝒚)φ(𝒙))}.\displaystyle\quad\quad+[\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})-\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x})](\varphi(\bm{y})-\varphi(\bm{x}))\}.

Therefore

d\Bϵ(𝒙)𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)φ(𝒚)𝒗(𝒚)+χ𝝂(𝒚𝒙)φ(𝒙)𝒗(𝒙))w(𝒚𝒙)𝑑𝒚\displaystyle\quad\ \int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\varphi(\bm{y})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\varphi(\bm{x})\bm{v}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}
=φ(𝒙)d\Bϵ(𝒙)𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)𝒗(𝒚)+χ𝝂(𝒚𝒙)𝒗(𝒙))w(𝒚𝒙)𝑑𝒚\displaystyle=\varphi(\bm{x})\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{v}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}
+ϕ(𝒙)d\Bϵ(𝒙)χ𝝂(𝒚𝒙)(φ(𝒚)φ(𝒙))𝒚𝒙|𝒚𝒙|w(𝒚𝒙)𝑑𝒚\displaystyle\quad+\bm{\phi}(\bm{x})\cdot\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\chi_{\bm{\nu}}(\bm{y}-\bm{x})(\varphi(\bm{y})-\varphi(\bm{x}))\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})d\bm{y}
+d\Bϵ(𝒙)𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)ϕ(𝒚)χ𝝂(𝒚𝒙)ϕ(𝒙))(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚.\displaystyle\quad+\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})-\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x}))(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}.

Thus, taking the limit as ϵ0\epsilon\to 0, by definition of the nonlocal gradient operator in eq. 2 and the equivalent form of the divergence operator in 2.3, we have

(42) 𝒟w𝝂(φϕ)(𝒙)=φ(𝒙)𝒟w𝝂ϕ(𝒙)+𝒢w𝝂φ(𝒙)ϕ(𝒙)+S(φ,ϕ)(𝒙),\mathcal{D}^{-\bm{\nu}}_{w}(\varphi\bm{\phi})(\bm{x})=\varphi(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})+\mathcal{G}^{\bm{\nu}}_{w}\varphi(\bm{x})\cdot\bm{\phi}(\bm{x})+S(\varphi,\bm{\phi})(\bm{x}),

where we used

limϵ0d\Bϵ(𝒙)𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)ϕ(𝒚)χ𝝂(𝒚𝒙)ϕ(𝒙))(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚\displaystyle\lim_{\epsilon\to 0}\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})-\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x}))(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}
=d𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)ϕ(𝒚)χ𝝂(𝒚𝒙)ϕ(𝒙))(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚.\displaystyle=\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})-\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x}))(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}.

Indeed, similar to (41), for every 𝒙d\bm{x}\in\mathbb{R}^{d},

|χ[ϵ,)(|𝒚𝒙|)𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)ϕ(𝒚)χ𝝂(𝒚𝒙)ϕ(𝒙))(φ(𝒚)φ(𝒙))w(𝒚𝒙)|4ϕL(d;d)φW1,(d)min(1,|𝒚𝒙|)w(𝒚𝒙),\begin{split}&\left|\chi_{[\epsilon,\infty)}(|\bm{y}-\bm{x}|)\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})-\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x}))(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})\right|\\ \leq&4\|\bm{\phi}\|_{L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d})}\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}\min(1,|\bm{y}-\bm{x}|)w(\bm{y}-\bm{x}),\end{split}

so the above limit is justified by the dominated convergence theorem. ∎

The generalized produce rule presented above is helpful in showing the following result, which says that 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) is closed under the multiplication with Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}).

Lemma 3.1 (Closedness under multiplication with bump functions).

For u𝒮w𝛎(d)u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) and φCc(d)\varphi\in C^{\infty}_{c}(\mathbb{R}^{d}), φu𝒮w𝛎(d)\varphi u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) and

(43) φu𝒮w𝝂(d)CφW1,(d)u𝒮w𝝂(d),φCc(d),u𝒮w𝝂(d),\|\varphi u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}\leq C\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}\|u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})},\quad\forall\varphi\in C^{\infty}_{c}(\mathbb{R}^{d}),\ u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}),

where CC depends on dd, Mw1M_{w}^{1} and Mw2M_{w}^{2}. As a result, for u𝒮w𝛎(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega) and φCc(d)\varphi\in C^{\infty}_{c}(\mathbb{R}^{d}) with suppφΩ\text{supp}\,\varphi\subset\Omega, φu𝒮w𝛎(Ω)\varphi u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega) for any domain Ωd\Omega\subset\mathbb{R}^{d}.

Proof.

First, notice that

φuL2(d)φL(d)uL2(d).\|\varphi u\|_{L^{2}(\mathbb{R}^{d})}\leq\|\varphi\|_{L^{\infty}(\mathbb{R}^{d})}\|u\|_{L^{2}(\mathbb{R}^{d})}.

Therefore we only need to show 𝔊w𝝂(φu)L2(d;d)\mathfrak{G}^{\bm{\nu}}_{w}(\varphi u)\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). The rest of proof in fact shows a ‘product rule’ for nonlocal distributional gradient 𝔊w𝝂\mathfrak{G}^{\bm{\nu}}_{w} using the ‘product rule’ of nonlocal divergence 𝒟w𝝂\mathcal{D}^{-\bm{\nu}}_{w} derived in 3.1. Since u𝒮w𝝂(d)u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}), there exists 𝒘=𝔊w𝝂uL2(d;d)\bm{w}=\mathfrak{G}^{\bm{\nu}}_{w}u\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}) such that

(44) d𝒘(𝒙)ϕ(𝒙)𝑑𝒙=du(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙,ϕCc(d;d).\int_{\mathbb{R}^{d}}\bm{w}(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x},\quad\forall\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}).

To show 𝔊w𝝂(φu)L2(d;d)\mathfrak{G}^{\bm{\nu}}_{w}(\varphi u)\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}), it suffices to find 𝒗L2(d;d)\bm{v}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}) such that

(45) d𝒗(𝒙)ϕ(𝒙)𝑑𝒙=dφ(𝒙)u(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙,ϕCc(d;d).\int_{\mathbb{R}^{d}}\bm{v}(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}\varphi(\bm{x})u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x},\quad\forall\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}).

By 3.1, we have

φ𝒟w𝝂ϕ=𝒟w𝝂(φϕ)𝒢w𝝂φϕS(φ,ϕ),\varphi\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}=\mathcal{D}^{-\bm{\nu}}_{w}(\varphi\bm{\phi})-\mathcal{G}^{\bm{\nu}}_{w}\varphi\cdot\bm{\phi}-S(\varphi,\bm{\phi}),

thus, for any ϕCc(d;d)\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}),

dφ(𝒙)u(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙\displaystyle\quad\,-\int_{\mathbb{R}^{d}}\varphi(\bm{x})u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x}
=du(𝒙)𝒟w𝝂(φϕ)(𝒙)𝑑𝒙+du(𝒙)𝒢w𝝂φ(𝒙)ϕ(𝒙)𝑑𝒙+du(𝒙)S(φ,ϕ)(𝒙)𝑑𝒙\displaystyle=-\int_{\mathbb{R}^{d}}u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}(\varphi\bm{\phi})(\bm{x})d\bm{x}+\int_{\mathbb{R}^{d}}u(\bm{x})\mathcal{G}^{\bm{\nu}}_{w}\varphi(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x}+\int_{\mathbb{R}^{d}}u(\bm{x})S(\varphi,\bm{\phi})(\bm{x})d\bm{x}
=d𝒘(𝒙)φ(𝒙)ϕ(𝒙)𝑑𝒙+du(𝒙)𝒢w𝝂φ(𝒙)ϕ(𝒙)𝑑𝒙+dH(u,φ)(𝒙)ϕ(𝒙)𝑑𝒙,\displaystyle=\int_{\mathbb{R}^{d}}\bm{w}(\bm{x})\varphi(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x}+\int_{\mathbb{R}^{d}}u(\bm{x})\mathcal{G}^{\bm{\nu}}_{w}\varphi(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x}+\int_{\mathbb{R}^{d}}H(u,\varphi)(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x},

where we use φϕCc(d;d)\varphi\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}) and H(u,φ):ddH(u,\varphi):\mathbb{R}^{d}\to\mathbb{R}^{d} is a vector-valued function whose expression will be given in a moment. Comparing this with (45), we notice that the vector-valued function 𝒗\bm{v} should be

(46) 𝒗(𝒙)=𝒘(𝒙)φ(𝒙)+u(𝒙)𝒢w𝝂φ(𝒙)+H(u,φ)(𝒙).\bm{v}(\bm{x})=\bm{w}(\bm{x})\varphi(\bm{x})+u(\bm{x})\mathcal{G}^{\bm{\nu}}_{w}\varphi(\bm{x})+H(u,\varphi)(\bm{x}).

It remains to show this function is in L2(d;d)L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). By the definition of S(φ,ϕ)S(\varphi,\bm{\phi}),

S(φ,ϕ)(𝒙)\displaystyle\quad\;S(\varphi,\bm{\phi})(\bm{x})
=d𝒚𝒙|𝒚𝒙|(χ𝝂(𝒙𝒚)ϕ(𝒚)χ𝝂(𝒚𝒙)ϕ(𝒙))(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚\displaystyle=\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{\phi}(\bm{y})-\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{\phi}(\bm{x}))(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}
=d𝒚𝒙|𝒚𝒙|χ𝝂(𝒙𝒚)ϕ(𝒚)(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚\displaystyle=\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\chi_{\bm{\nu}}(\bm{x}-\bm{y})\cdot\bm{\phi}(\bm{y})(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}
d𝒚𝒙|𝒚𝒙|χ𝝂(𝒚𝒙)(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚ϕ(𝒙)\displaystyle\quad-\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\chi_{\bm{\nu}}(\bm{y}-\bm{x})(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}\cdot\bm{\phi}(\bm{x})
=:H1(φ,ϕ)(𝒙)+H2(φ)(𝒙)ϕ(𝒙).\displaystyle=:H_{1}(\varphi,\bm{\phi})(\bm{x})+H_{2}(\varphi)(\bm{x})\cdot\bm{\phi}(\bm{x}).

Note that both H1(φ,ϕ)H_{1}(\varphi,\bm{\phi}) and H2(φ)H_{2}(\varphi) are well-defined maps on d\mathbb{R}^{d} due to the similar reason for the the pointwise estimate (41) for S(φ,ϕ)S(\varphi,\bm{\phi}). For example,

(47) |H2(φ)(𝒙)|=|d𝒚𝒙|𝒚𝒙|χ𝝂(𝒚𝒙)(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚|d(2φL(d)χ|𝒚𝒙|>1(𝒚)+φL(d)|𝒚𝒙|χB1(𝒙)(𝒚))w(𝒚𝒙)𝑑𝒚2φW1,(d)(Mw1+Mw2),𝒙d.\begin{split}&|H_{2}(\varphi)(\bm{x})|=\left|-\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\chi_{\bm{\nu}}(\bm{y}-\bm{x})(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}\right|\\ \leq&\int_{\mathbb{R}^{d}}(2\|\varphi\|_{L^{\infty}(\mathbb{R}^{d})}\chi_{|\bm{y}-\bm{x}|>1}(\bm{y})+\|\nabla\varphi\|_{L^{\infty}(\mathbb{R}^{d})}|\bm{y}-\bm{x}|\chi_{B_{1}(\bm{x})}(\bm{y}))w(\bm{y}-\bm{x})d\bm{y}\\ \leq&2\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}(M_{w}^{1}+M_{w}^{2}),\ \bm{x}\in\mathbb{R}^{d}.\end{split}

Observe that by Fubini’s theorem

du(𝒙)H1(φ,ϕ)(𝒙)𝑑𝒙\displaystyle\quad\ \int_{\mathbb{R}^{d}}u(\bm{x})H_{1}(\varphi,\bm{\phi})(\bm{x})d\bm{x}
=du(𝒙)d𝒚𝒙|𝒚𝒙|χ𝝂(𝒙𝒚)ϕ(𝒚)(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒚𝑑𝒙\displaystyle=\int_{\mathbb{R}^{d}}u(\bm{x})\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\chi_{\bm{\nu}}(\bm{x}-\bm{y})\cdot\bm{\phi}(\bm{y})(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}d\bm{x}
=dϕ(𝒚)F(u,φ)(𝒚)𝑑𝒚,\displaystyle=\int_{\mathbb{R}^{d}}\bm{\phi}(\bm{y})\cdot F(u,\varphi)(\bm{y})d\bm{y},

where F(u,φ):ddF(u,\varphi):\mathbb{R}^{d}\to\mathbb{R}^{d} is given by

F(u,φ)(𝒚):=du(𝒙)𝒚𝒙|𝒚𝒙|χ𝝂(𝒙𝒚)(φ(𝒚)φ(𝒙))w(𝒚𝒙)𝑑𝒙.F(u,\varphi)(\bm{y}):=\int_{\mathbb{R}^{d}}u(\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\chi_{\bm{\nu}}(\bm{x}-\bm{y})(\varphi(\bm{y})-\varphi(\bm{x}))w(\bm{y}-\bm{x})d\bm{x}.

Using Holder’s inequality, one can show F(u,φ)L2(d;d)F(u,\varphi)\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). Indeed,

(d|F(u,φ)(𝒚)|2𝑑𝒚)12\displaystyle\quad\left(\int_{\mathbb{R}^{d}}|F(u,\varphi)(\bm{y})|^{2}d\bm{y}\right)^{\frac{1}{2}}
2φW1,(d)(d|d|u(𝒙)|min(1,|𝒚𝒙|)w(𝒚𝒙)𝑑𝒙|2𝑑𝒚)12\displaystyle\leq 2\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}\left(\int_{\mathbb{R}^{d}}\left|\int_{\mathbb{R}^{d}}|u(\bm{x})|\min(1,|\bm{y}-\bm{x}|)w(\bm{y}-\bm{x})d\bm{x}\right|^{2}d\bm{y}\right)^{\frac{1}{2}}
2φW1,(d)(Mw1+Mw2)12(dd|u(𝒙)|2min(1,|𝒚𝒙|)w(𝒚𝒙)𝑑𝒙𝑑𝒚)12\displaystyle\leq 2\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}(M_{w}^{1}+M_{w}^{2})^{\frac{1}{2}}\left(\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}|u(\bm{x})|^{2}\min(1,|\bm{y}-\bm{x}|)w(\bm{y}-\bm{x})d\bm{x}d\bm{y}\right)^{\frac{1}{2}}
2φW1,(d)(Mw1+Mw2)uL2(d)<.\displaystyle\leq 2\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}(M_{w}^{1}+M_{w}^{2})\|u\|_{L^{2}(\mathbb{R}^{d})}<\infty.

Combining the above discussions, we obtain

(48) H(u,φ)(𝒙)=F(u,φ)(𝒙)+u(𝒙)H2(φ)(𝒙),H(u,\varphi)(\bm{x})=F(u,\varphi)(\bm{x})+u(\bm{x})H_{2}(\varphi)(\bm{x}),

with

H(u,φ)L2(d;d)F(u,φ)L2(d;d)+H2(φ)L(d;d)uL2(d)4φW1,(d)(Mw1+Mw2)uL2(d).\begin{split}\|H(u,\varphi)\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}&\leq\|F(u,\varphi)\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}+\|H_{2}(\varphi)\|_{L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d})}\|u\|_{L^{2}(\mathbb{R}^{d})}\\ &\leq 4\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}(M_{w}^{1}+M_{w}^{2})\|u\|_{L^{2}(\mathbb{R}^{d})}.\end{split}

Therefore, by (46), 𝒗=𝒘φ+u𝒢w𝝂φ+H(u,φ)L2(d;d)\bm{v}=\bm{w}\varphi+u\mathcal{G}^{\bm{\nu}}_{w}\varphi+H(u,\varphi)\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}) and

𝒗L2(d;d)\displaystyle\quad\;\|\bm{v}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}
𝒘L2(d;d)φL(d)+uL2(d)𝒢w𝝂φL(d;d)+H(u,φ)L2(d;d)\displaystyle\leq\|\bm{w}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\|\varphi\|_{L^{\infty}(\mathbb{R}^{d})}+\|u\|_{L^{2}(\mathbb{R}^{d})}\|\mathcal{G}^{\bm{\nu}}_{w}\varphi\|_{L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d})}+\|H(u,\varphi)\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}
𝒘L2(d;d)φL(d)+uL2(d)C(Mw1+Mw2)φW1,(d)\displaystyle\leq\|\bm{w}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\|\varphi\|_{L^{\infty}(\mathbb{R}^{d})}+\|u\|_{L^{2}(\mathbb{R}^{d})}C(M_{w}^{1}+M_{w}^{2})\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}
+4φW1,(d)(Mw1+Mw2)uL2(d)\displaystyle\quad+4\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}(M_{w}^{1}+M_{w}^{2})\|u\|_{L^{2}(\mathbb{R}^{d})}
CφW1,(d)u𝒮w𝝂(d),\displaystyle\leq C\|\varphi\|_{W^{1,\infty}(\mathbb{R}^{d})}\|u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})},

where we have used 2.1. This combined with the L2L^{2} estimate on φu\varphi u leads to eq. 43. ∎

Next, we present two results regarding the translation and mollification of functions in 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}), which are standard techniques useful for proving density of smooth functions. For f:dkf:\mathbb{R}^{d}\to\mathbb{R}^{k} and a given vector 𝒂d\bm{a}\in\mathbb{R}^{d}, denote the translation operator τ𝒂f(𝒙):=f(𝒙+𝒂)\tau_{\bm{a}}f(\bm{x}):=f(\bm{x}+\bm{a}). In addition, we let ηϵ\eta_{\epsilon} be the standard mollifiers for ϵ>0\epsilon>0, i.e. ηϵ(𝒙)=1ϵdη(xϵ)\eta_{\epsilon}(\bm{x})=\frac{1}{\epsilon^{d}}\eta(\frac{x}{\epsilon}) where ηCc(d)\eta\in C^{\infty}_{c}(\mathbb{R}^{d}) and dη(𝒙)𝑑𝒙=1\int_{\mathbb{R}^{d}}\eta(\bm{x})d\bm{x}=1. The statement of the following two lemmas are new but the proofs follow the standard arguments of similar results in the classical Sobolev spaces. We therefore leave their proofs in the appendix.

Lemma 3.2 (Continuity of translation).

For u𝒮w𝛎(d)u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) and 𝐚d\bm{a}\in\mathbb{R}^{d}, τ𝐚u𝒮w𝛎(d)\tau_{\bm{a}}u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) and

lim|𝒂|0τ𝒂uu𝒮w𝝂(d)=0.\lim_{|\bm{a}|\to 0}\|\tau_{\bm{a}}u-u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}=0.
Lemma 3.3 (Mollification in 𝒮w𝝂(d)\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})).

For u𝒮w𝛎(d)u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) and ϵ>0\epsilon>0, ηϵu𝒮w𝛎(d)\eta_{\epsilon}*u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) and

(49) limϵ0ηϵuu𝒮w𝝂(d)=0.\lim_{\epsilon\to 0}\|\eta_{\epsilon}*u-u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}=0.

With the necessary components presented in Lemmas 3.1, 3.2 and 3.3, the proof of 3.2 uses the standard mollification and partition of unity techniques (see [1, 24] for instance). Here we present its proof for completeness. Similar arguments can be found in [25] or [27] (Theorem 3.76(i)).

Proof of 3.2.

We prove the result for Ω\Omega being a bounded domain with continuous boundary. The other two cases are more straightforward. Since Ω\partial\Omega is compact, there exist 𝒙iΩ\bm{x}_{i}\in\partial\Omega, i=1,,Ni=1,\cdots,N and r>0r>0 such that

Ωi=1NBr/2(𝒙i),\partial\Omega\subset\bigcup_{i=1}^{N}B_{r/2}(\bm{x}_{i}),

and

ΩBr(𝒙i)={𝒙=(𝒙,xd)Br(𝒙i)|xd>ζi(𝒙)}\Omega\cap B_{r}(\bm{x}_{i})=\{\bm{x}=(\bm{x}^{\prime},x_{d})\in B_{r}(\bm{x}_{i})\,|\,x_{d}>\zeta_{i}(\bm{x}^{\prime})\}
ΩcBr(𝒙i)={𝒙=(𝒙,xd)Br(𝒙i)|xdζi(𝒙)}\Omega^{c}\cap B_{r}(\bm{x}_{i})=\{\bm{x}=(\bm{x}^{\prime},x_{d})\in B_{r}(\bm{x}_{i})\,|\,x_{d}\leq\zeta_{i}(\bm{x}^{\prime})\}

for some continuous functions ζi:d1\zeta_{i}:\mathbb{R}^{d-1}\to\mathbb{R} up to relabelling the coordinates. Let Ωr/2:={xΩ:dist(x,Ω)>r/2}\Omega^{\circ}_{r/2}:=\{x\in\Omega:\mathrm{dist}(x,\partial\Omega)>r/2\}. Then,

Ωi=1NBr(𝒙i)Ωr/2.\Omega\subset\bigcup_{i=1}^{N}B_{r}(\bm{x}_{i})\cup\Omega^{\circ}_{r/2}.

Let {φi}i=0N\{\varphi_{i}\}_{i=0}^{N} be a smooth partition of unity subordinate to the above constructed sets. That is we have φi0\varphi_{i}\geq 0, i=0Nφi=1\sum_{i=0}^{N}\varphi_{i}=1 and φ0Cc(Ωr/2)\varphi_{0}\in C^{\infty}_{c}(\Omega^{\circ}_{r/2}) and φiCc(Br(𝒙i))\varphi_{i}\in C^{\infty}_{c}(B_{r}(\bm{x}_{i})), 1iN1\leq i\leq N. Let u𝒮w𝝂(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega) and define

ui:=φiu for all i{0,,N}.u^{i}:=\varphi_{i}u\quad\text{ for all }i\in\{0,\cdots,N\}.

By 3.1, ui𝒮w𝝂(Ω)u^{i}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega). For μ>0\mu>0, we define

uμi(𝒙)=ui(𝒙,xdμ)for i{1,,N},𝒙=(𝒙,xd)d.u^{i}_{\mu}(\bm{x})=u^{i}(\bm{x}^{\prime},x_{d}-\mu)\quad\text{for }\ i\in\{1,\cdots,N\},\ \bm{x}=(\bm{x}^{\prime},x_{d})\in\mathbb{R}^{d}.

Fix σ>0\sigma>0, by Lemma 3.2, there exists μ(0,12min1iNdist(suppφi,Br(𝒙i)))\mu\in(0,\frac{1}{2}\min_{1\leq i\leq N}\mathrm{dist}(\text{supp}\ \varphi_{i},\partial B_{r}(\bm{x}_{i}))) such that

uμiui𝒮w𝝂(d)<σ2(N+1),1iN.\|u^{i}_{\mu}-u^{i}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}<\frac{\sigma}{2(N+1)},\quad\forall 1\leq i\leq N.

Fix this μ\mu, it follows that ηϵuμiCc(Ω)\eta_{\epsilon}*u^{i}_{\mu}\in C^{\infty}_{c}(\Omega) for a positive number ϵ\epsilon less than min1iNdist(suppuμi,Ω)/2\min_{1\leq i\leq N}\mathrm{dist}(\text{supp}\ u^{i}_{\mu},\partial\Omega)/2. Indeed, since suppuμiWμi¯ΩBr(𝒙i)\text{supp}\,u^{i}_{\mu}\subset\overline{W^{i}_{\mu}}\subset\Omega\cap B_{r}(\bm{x}_{i}), where Wμi:={𝒛=(𝒛,zd)Br(𝒙i):zdμ>ζi(𝒛)}W^{i}_{\mu}:=\{\bm{z}=(\bm{z}^{\prime},z_{d})\in B_{r}(\bm{x}_{i}):z_{d}-\mu>\zeta_{i}(\bm{z}^{\prime})\}, 1iN1\leq i\leq N,

supp(ηϵuμi)Bϵ(𝟎)+suppuμi¯Ω.\text{supp}(\eta_{\epsilon}*u^{i}_{\mu})\subset\overline{B_{\epsilon}(\bm{0})+\text{supp}\ u^{i}_{\mu}}\subset\Omega.

Since uμi𝒮w𝝂(d)u^{i}_{\mu}\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}), by Lemma 3.3, ηϵuμi𝒮w𝝂(d)\eta_{\epsilon}*u^{i}_{\mu}\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}) and there exists ϵ>0\epsilon>0 such that ηϵu0Cc(Ω)\eta_{\epsilon}*u^{0}\in C^{\infty}_{c}(\Omega),

ηϵuμiuμi𝒮w𝝂(d)<σ2(N+1)\|\eta_{\epsilon}*u^{i}_{\mu}-u^{i}_{\mu}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}<\frac{\sigma}{2(N+1)}

and

ηϵu0u0𝒮w𝝂(d)<σ2(N+1).\|\eta_{\epsilon}*u^{0}-u^{0}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}<\frac{\sigma}{2(N+1)}.

Let vϵ:=ηϵu0+i=1Nηϵuμiv_{\epsilon}:=\eta_{\epsilon}*u^{0}+\sum_{i=1}^{N}\eta_{\epsilon}*u^{i}_{\mu}, we have vϵCc(Ω)v_{\epsilon}\in C^{\infty}_{c}(\Omega) and vϵu𝒮w𝝂(d)<σ\|v_{\epsilon}-u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}<\sigma. Therefore, the lemma is proved. ∎

3.2. Nonlocal vector inequalities and identities

In this subsection, we derive a number of results for Sobolev-type functions using 3.2. The first two results are the analogs of H1H(div)H^{1}\subset H(\mathrm{div}) and H1H(curl)H^{1}\subset H(\mathrm{curl}) in the local setting. We assume Ω\Omega is a bounded domain with a continuous boundary, an epigraph, or d\mathbb{R}^{d} so that the density result holds.

Proposition 3.2.

For 𝐮𝒮w𝛎(Ω;d)\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}), 𝔇w±𝛎𝐮L2(d)\mathfrak{D}_{w}^{\pm\bm{\nu}}\bm{u}\in L^{2}(\mathbb{R}^{d}) and

𝔇w±𝝂𝒖L2(d)𝔊w𝝂𝒖L2(d;d×d).\|\mathfrak{D}_{w}^{\pm\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{d})}\leq\|\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}.

Thus, 𝔇w±𝛎:𝒮w𝛎(Ω;d)L2(d)\mathfrak{D}_{w}^{\pm\bm{\nu}}:\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d})\to L^{2}(\mathbb{R}^{d}) is a bounded linear operator with operator norm no more than 11. In addition, there exists {𝐮(n)}n=1Cc(Ω;d)\{\bm{u}^{(n)}\}_{n=1}^{\infty}\subset C^{\infty}_{c}(\Omega;\mathbb{R}^{d}) such that 𝐮(n)𝐮\bm{u}^{(n)}\to\bm{u} in 𝒮w𝛎(Ω;d)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}) and 𝒟w±𝛎𝐮(n)𝔇w±𝛎𝐮\mathcal{D}_{w}^{\pm\bm{\nu}}\bm{u}^{(n)}\to\mathfrak{D}_{w}^{\pm\bm{\nu}}\bm{u} in L2(d)L^{2}(\mathbb{R}^{d}) as nn\to\infty.

Proof.

We first show the inequality for smooth functions with compact support, that is, assuming 𝒖Cc(Ω;d)\bm{u}\in C^{\infty}_{c}(\Omega;\mathbb{R}^{d}),

(50) 𝒟w±𝝂𝒖L2(d)𝒢w𝝂𝒖L2(d;d×d).\|\mathcal{D}_{w}^{\pm\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{d})}\leq\|\mathcal{G}_{w}^{\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}.

We only prove eq. 50 for 𝒟w𝝂\mathcal{D}^{-\bm{\nu}}_{w} since the result also holds for 𝒟w𝝂\mathcal{D}^{\bm{\nu}}_{w} by noticing that 𝒢w𝝂𝒖L2(d;d×d)=𝒢w𝝂𝒖L2(d;d×d)\|\mathcal{G}_{w}^{\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}=\|\mathcal{G}_{w}^{-\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})} using Plancherel’s theorem. For 1id1\leq i\leq d and a scalar function p:dp:\mathbb{R}^{d}\to\mathbb{R}, we introduce the notation 𝒢i\mathcal{G}_{i} given by

𝒢w𝝂p:=(𝒢1p,,𝒢dp)T.\mathcal{G}^{\bm{\nu}}_{w}p:=(\mathcal{G}_{1}p,\cdots,\mathcal{G}_{d}p)^{T}.

We use Fourier transform to show that

(51) 𝒟w𝝂𝒖L2(d)2=i,j=1d(𝒢iuj,𝒢jui)L2(d),\left\|\mathcal{D}_{w}^{-\bm{\nu}}\bm{u}\right\|_{L^{2}(\mathbb{R}^{d})}^{2}=\sum_{i,j=1}^{d}\left(\mathcal{G}_{i}u_{j},\mathcal{G}_{j}u_{i}\right)_{L^{2}(\mathbb{R}^{d})},

which implies the desired result by Young’s inequality for products. By 2.3 and 𝝀w𝝂(𝝃)=𝝀w𝝂(𝝃)¯\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})=-\overline{\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})}, for 𝝃d\bm{\xi}\in\mathbb{R}^{d},

(𝒟w𝝂𝒖)(𝝃)=i=1dλi(𝝃)¯u^i(𝝃)and(𝒢iuj)(𝝃)=λi(𝝃)u^j(𝝃), 1i,jd,\mathcal{F}(\mathcal{D}_{w}^{-\bm{\nu}}\bm{u})(\bm{\xi})=-\sum_{i=1}^{d}\overline{\lambda_{i}(\bm{\xi})}\hat{u}_{i}(\bm{\xi})\quad\text{and}\quad\mathcal{F}(\mathcal{G}_{i}u_{j})(\bm{\xi})=\lambda_{i}(\bm{\xi})\hat{u}_{j}(\bm{\xi}),\ 1\leq i,j\leq d,

where the Fourier symbol

𝝀w𝝂(𝝃):=(λ1(𝝃),,λd(𝝃))T.\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi}):=(\lambda_{1}(\bm{\xi}),\cdots,\lambda_{d}(\bm{\xi}))^{T}.

Therefore, by Plancherel’s theorem, we conclude the proof of eq. 50 as

i,j=1d(𝒢iuj,𝒢jui)L2(d)=i,j=1d((𝒢iuj),(𝒢jui))L2(d;d)\displaystyle\sum_{i,j=1}^{d}\left(\mathcal{G}_{i}u_{j},\mathcal{G}_{j}u_{i}\right)_{L^{2}(\mathbb{R}^{d})}=\sum_{i,j=1}^{d}\left(\mathcal{F}(\mathcal{G}_{i}u_{j}),\mathcal{F}(\mathcal{G}_{j}u_{i})\right)_{L^{2}(\mathbb{R}^{d};\mathbb{C}^{d})}
=\displaystyle= i,j=1ddλi(𝝃)u^j(𝝃)λj(𝝃)¯u^i(𝝃)¯𝑑𝝃=dj=1dλj(𝝃)¯u^j(𝝃)i=1dλi(𝝃)u^i(𝝃)¯d𝝃\displaystyle\sum_{i,j=1}^{d}\int_{\mathbb{R}^{d}}\lambda_{i}(\bm{\xi})\hat{u}_{j}(\bm{\xi})\overline{\lambda_{j}(\bm{\xi})}\overline{\hat{u}_{i}(\bm{\xi})}d\bm{\xi}=\int_{\mathbb{R}^{d}}\sum_{j=1}^{d}\overline{\lambda_{j}(\bm{\xi})}\hat{u}_{j}(\bm{\xi})\sum_{i=1}^{d}\lambda_{i}(\bm{\xi})\overline{\hat{u}_{i}(\bm{\xi})}d\bm{\xi}
=\displaystyle= (𝒟w𝝂𝒖)L2(d;d)2=𝒟w𝝂𝒖L2(d)2.\displaystyle\left\|\mathcal{F}(\mathcal{D}_{w}^{-\bm{\nu}}\bm{u})\right\|_{L^{2}(\mathbb{R}^{d};\mathbb{C}^{d})}^{2}=\left\|\mathcal{D}_{w}^{-\bm{\nu}}\bm{u}\right\|_{L^{2}(\mathbb{R}^{d})}^{2}.

By 3.2, there exists {𝒖(n)}n=1Cc(Ω;d)\{\bm{u}^{(n)}\}_{n=1}^{\infty}\subset C^{\infty}_{c}(\Omega;\mathbb{R}^{d}) such that 𝒖(n)𝒖\bm{u}^{(n)}\to\bm{u} in 𝒮w𝝂(Ω;d)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}), and in particular, 𝒢w𝝂𝒖(n)𝔊w𝝂𝒖\mathcal{G}^{\bm{\nu}}_{w}\bm{u}^{(n)}\to\mathfrak{G}^{\bm{\nu}}_{w}\bm{u} in L2(d;d×d)L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d}) as nn\to\infty. Applying eq. 50 fo 𝒖(n)\bm{u}^{(n)}, {𝒟w±𝝂𝒖(n)}n\{\mathcal{D}^{\pm\bm{\nu}}_{w}\bm{u}^{(n)}\}_{n\in\mathbb{N}} is a Cauchy sequence in L2(d)L^{2}(\mathbb{R}^{d}), and thus has a limit in L2(d)L^{2}(\mathbb{R}^{d}) by completeness. Since 𝒖(n)𝒖\bm{u}^{(n)}\to\bm{u} in L2(d;d)L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}), by applying 2.1(2) to 𝒖(n)\bm{u}^{(n)} one derives that the limit is 𝔇w±𝝂𝒖L2(d)\mathfrak{D}_{w}^{\pm\bm{\nu}}\bm{u}\in L^{2}(\mathbb{R}^{d}) with the desired estimate. ∎

Proposition 3.3.

Let d=3d=3. For 𝐮𝒮w𝛎(Ω;3)\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}), w±𝛎𝐮L2(3;3)\mathfrak{C}_{w}^{\pm\bm{\nu}}\bm{u}\in L^{2}(\mathbb{R}^{3};\mathbb{R}^{3}) and

w±𝝂𝒖L2(3;3)𝔊w𝝂𝒖L2(3;3×3).\|\mathfrak{C}_{w}^{\pm\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3})}\leq\|\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3\times 3})}.

Thus, w±𝛎:𝒮w𝛎(Ω;3)L2(3;3)\mathfrak{C}_{w}^{\pm\bm{\nu}}:\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3})\to L^{2}(\mathbb{R}^{3};\mathbb{R}^{3}) is a bounded linear operator with operator norm no more than 11. In addition, there exists {𝐮(n)}n=1Cc(Ω;3)\{\bm{u}^{(n)}\}_{n=1}^{\infty}\subset C^{\infty}_{c}(\Omega;\mathbb{R}^{3}) such that 𝐮(n)𝐮\bm{u}^{(n)}\to\bm{u} in 𝒮w𝛎(Ω;3)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}) and 𝒞w±𝛎𝐮(n)w±𝛎𝐮\mathcal{C}_{w}^{\pm\bm{\nu}}\bm{u}^{(n)}\to\mathfrak{C}_{w}^{\pm\bm{\nu}}\bm{u} in L2(3;3)L^{2}(\mathbb{R}^{3};\mathbb{R}^{3}) as nn\to\infty.

Proof.

Using the density result 3.2 and integration by parts formula 2.1(3) as in the last paragraph of the proof of 3.2, it suffices to show

(52) 𝒞w±𝝂𝒖L2(3;3)𝒢w𝝂𝒖L2(3;3×3)\|\mathcal{C}_{w}^{\pm\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3})}\leq\|\mathcal{G}_{w}^{\bm{\nu}}\bm{u}\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3\times 3})}

for 𝒖Cc(Ω;3)\bm{u}\in C^{\infty}_{c}(\Omega;\mathbb{R}^{3}). To show eq. 52, by Fourier transform, it suffices to show that

(𝒞w𝝂𝒖)L2(3;3)(𝒢w𝝂𝒖)L2(3;3×3).\|\mathcal{F}(\mathcal{C}^{\bm{\nu}}_{w}\bm{u})\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3})}\leq\|\mathcal{F}(\mathcal{G}^{\bm{\nu}}_{w}\bm{u})\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3\times 3})}.

Since |𝒂×𝒃||𝒂||𝒃||\bm{a}\times\bm{b}|\leq|\bm{a}||\bm{b}| for 𝒂,𝒃3\bm{a},\bm{b}\in\mathbb{C}^{3}, by 2.3 eq. 52 holds as

(𝒞w𝝂𝒖)L2(3;3)2\displaystyle\|\mathcal{F}(\mathcal{C}^{\bm{\nu}}_{w}\bm{u})\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3})}^{2} =3|𝝀w𝝂(𝝃)×𝒖^(𝝃)|2𝑑𝝃3|𝝀w𝝂(𝝃)|2|𝒖^(𝝃)|2𝑑𝝃\displaystyle=\int_{\mathbb{R}^{3}}|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\times\hat{\bm{u}}(\bm{\xi})|^{2}d\bm{\xi}\leq\int_{\mathbb{R}^{3}}|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})|^{2}|\hat{\bm{u}}(\bm{\xi})|^{2}d\bm{\xi}
=j=133|𝝀w𝝂(𝝃)u^j(𝝃)|2𝑑𝝃=(𝒢w𝝂𝒖)L2(3;3×3)2.\displaystyle=\sum_{j=1}^{3}\int_{\mathbb{R}^{3}}|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\hat{u}_{j}(\bm{\xi})|^{2}d\bm{\xi}=\|\mathcal{F}(\mathcal{G}^{\bm{\nu}}_{w}\bm{u})\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3\times 3})}^{2}.

Recall that 𝔇w𝝂𝒗(Cc(d))(Cc(Ω))\mathfrak{D}_{w}^{-\bm{\nu}}\bm{v}\in(C^{\infty}_{c}(\mathbb{R}^{d}))^{\prime}\subset(C^{\infty}_{c}(\Omega))^{\prime} for 𝒗L2(d;d)\bm{v}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). By 3.2 one can define 𝔇w𝝂𝒗(𝒮w𝝂(Ω))\mathfrak{D}_{w}^{-\bm{\nu}}\bm{v}\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega))^{\ast} for 𝒗L2(d;d)\bm{v}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). More precisely, given u𝒮w𝝂(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega), define

(53) 𝔇w𝝂𝒗,u:=d𝒗(𝒙)𝔊w𝝂u(𝒙)𝑑𝒙.\langle\mathfrak{D}_{w}^{-\bm{\nu}}\bm{v},u\rangle:=-\int_{\mathbb{R}^{d}}\bm{v}(\bm{x})\cdot\mathfrak{G}^{\bm{\nu}}_{w}u(\bm{x})d\bm{x}.

Then

|𝔇w𝝂𝒗,u|𝒗L2(d;d)𝔊w𝝂uL2(d;d)𝒗L2(d;d)u𝒮w𝝂(Ω)|\langle\mathfrak{D}_{w}^{-\bm{\nu}}\bm{v},u\rangle|\leq\|\bm{v}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\|\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\leq\|\bm{v}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\|u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)}

implies that 𝔇w𝝂𝒗(𝒮w𝝂(Ω))\mathfrak{D}_{w}^{-\bm{\nu}}\bm{v}\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega))^{\ast}. Therefore, 𝔇w𝝂:L2(d;d)(𝒮w𝝂(Ω))\mathfrak{D}_{w}^{-\bm{\nu}}:L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})\to(\mathcal{S}_{w}^{\bm{\nu}}(\Omega))^{\ast} is a bounded linear operator with operator norm no more than 11. The same property holds for 𝔊w𝝂\mathfrak{G}^{\bm{\nu}}_{w} and w𝝂\mathfrak{C}^{\bm{\nu}}_{w} using 3.2 and 3.3, respectively. We summarize this observation in the following proposition.

Proposition 3.4.

𝔇w𝝂:L2(d;d)(𝒮w±𝝂(Ω))\mathfrak{D}_{w}^{\mp\bm{\nu}}:L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})\to(\mathcal{S}_{w}^{\pm\bm{\nu}}(\Omega))^{\ast} defined by eq. 53 is a bounded linear operator with operator norm no more than 11. We can similarly define 𝔊w±𝛎:L2(d)(𝒮w𝛎(Ω;d))\mathfrak{G}_{w}^{\pm\bm{\nu}}:L^{2}(\mathbb{R}^{d})\to(\mathcal{S}^{\bm{\nu}}_{w}(\Omega;\mathbb{R}^{d}))^{\ast} and w±𝛎:L2(3;3)(𝒮w𝛎(Ω;3))\mathfrak{C}_{w}^{\pm\bm{\nu}}:L^{2}(\mathbb{R}^{3};\mathbb{R}^{3})\to(\mathcal{S}^{\bm{\nu}}_{w}(\Omega;\mathbb{R}^{3}))^{\ast} and they are bounded linear operators with operator norms no more than 11.

Based on the above results, the nonlocal vector identities in Section 2.4 hold for functions in the space 𝒮w𝝂(Ω;N)\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{N}). The vector identities shown below are crucial for establishing the nonlocal Helmholtz decomposition in Section 6.3.

Lemma 3.4.

Let d=3d=3. Then for u𝒮w𝛎(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega) and 𝐯𝒮w𝛎(Ω;3)\bm{v}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}), in the sense of distributions,

(54) w𝝂𝔊w𝝂u=0,\mathfrak{C}^{\bm{\nu}}_{w}\mathfrak{G}^{\bm{\nu}}_{w}u=0,

and

(55) 𝔇w𝝂w𝝂𝒗=0.\mathfrak{D}^{\bm{\nu}}_{w}\mathfrak{C}^{\bm{\nu}}_{w}\bm{v}=0.
Proof.

Since u𝒮w𝝂(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega), 𝔊w𝝂uL2(3;3)\mathfrak{G}^{\bm{\nu}}_{w}u\in L^{2}(\mathbb{R}^{3};\mathbb{R}^{3}). By definition, w𝝂𝔊w𝝂u(Cc(3;3))\mathfrak{C}^{\bm{\nu}}_{w}\mathfrak{G}^{\bm{\nu}}_{w}u\in(C_{c}^{\infty}(\mathbb{R}^{3};\mathbb{R}^{3}))^{\prime} and for ϕCc(3;3)\bm{\phi}\in C_{c}^{\infty}(\mathbb{R}^{3};\mathbb{R}^{3}),

(56) w𝝂𝔊w𝝂u,ϕ=3𝔊w𝝂u(𝒙)𝒞w𝝂ϕ(𝒙)𝑑𝒙=limn3𝒢w𝝂u(n)(𝒙)𝒞w𝝂ϕ(𝒙)𝑑𝒙,\langle\mathfrak{C}_{w}^{\bm{\nu}}\mathfrak{G}^{\bm{\nu}}_{w}u,\bm{\phi}\rangle=\int_{\mathbb{R}^{3}}\mathfrak{G}^{\bm{\nu}}_{w}u(\bm{x})\cdot\mathcal{C}_{w}^{-\bm{\nu}}\bm{\phi}(\bm{x})d\bm{x}=\lim_{n\to\infty}\int_{\mathbb{R}^{3}}\mathcal{G}^{\bm{\nu}}_{w}u^{(n)}(\bm{x})\cdot\mathcal{C}_{w}^{-\bm{\nu}}\bm{\phi}(\bm{x})d\bm{x},

where the sequence {u(n)}n=1Cc(3)\left\{u^{(n)}\right\}_{n=1}^{\infty}\subset C^{\infty}_{c}(\mathbb{R}^{3}) is chosen according to 3.2 such that 𝒢w𝝂u(n)𝔊w𝝂u\mathcal{G}^{\bm{\nu}}_{w}u^{(n)}\to\mathfrak{G}^{\bm{\nu}}_{w}u in L2(3;3)L^{2}(\mathbb{R}^{3};\mathbb{R}^{3}). By integration by parts formula,

3𝒢w𝝂u(n)(𝒙)𝒞w𝝂ϕ(𝒙)𝑑𝒙=3u(n)(𝒙)𝒟w𝝂𝒞w𝝂ϕ(𝒙)𝑑𝒙.\int_{\mathbb{R}^{3}}\mathcal{G}^{\bm{\nu}}_{w}u^{(n)}(\bm{x})\cdot\mathcal{C}_{w}^{-\bm{\nu}}\bm{\phi}(\bm{x})d\bm{x}=\int_{\mathbb{R}^{3}}u^{(n)}(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\mathcal{C}_{w}^{-\bm{\nu}}\bm{\phi}(\bm{x})d\bm{x}.

Since 𝒟w𝝂𝒞w𝝂ϕ=0\mathcal{D}^{-\bm{\nu}}_{w}\mathcal{C}_{w}^{-\bm{\nu}}\bm{\phi}=0 by 2.5, we have w𝝂𝔊w𝝂u=0(Cc(3;3))\mathfrak{C}^{\bm{\nu}}_{w}\mathfrak{G}^{\bm{\nu}}_{w}u=0\in(C_{c}^{\infty}(\mathbb{R}^{3};\mathbb{R}^{3}))^{\prime}. Then by 3.4, w𝝂𝔊w𝝂u=0(𝒮w𝝂(Ω;3))\mathfrak{C}^{\bm{\nu}}_{w}\mathfrak{G}^{\bm{\nu}}_{w}u=0\in(\mathcal{S}^{\bm{\nu}}_{w}(\Omega;\mathbb{R}^{3}))^{\ast}. Equation 55 can be shown similarly. ∎

Remark 3.2.

For d=2d=2, one can show by similar arguments as those in 2.5 and 3.4 that

𝔇w𝝂(0110)𝔊w𝝂u=0\mathfrak{D}^{\bm{\nu}}_{w}\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\mathfrak{G}^{\bm{\nu}}_{w}u=0

for u𝒮w𝛎(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega), which can be seen as a 2d version of eq. 54 or eq. 55.

Lemma 3.5.

For 𝐮𝒮w𝛎(Ω;2)\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}),

(57) 𝔇w𝝂𝔊w𝝂𝒖=𝔊w𝝂𝔇w𝝂𝒖(0110)𝔊w𝝂𝔇w𝝂[(0110)𝒖].\mathfrak{D}_{w}^{-\bm{\nu}}\mathfrak{G}_{w}^{\bm{\nu}}\bm{u}=\mathfrak{G}_{w}^{\bm{\nu}}\mathfrak{D}_{w}^{-\bm{\nu}}\bm{u}-\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\mathfrak{G}_{w}^{-\bm{\nu}}\mathfrak{D}_{w}^{\bm{\nu}}\left[\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\bm{u}\right].
Proof.

Notice that the left hand side and the right hand side of (57) are understood as elements in (𝒮w𝝂(Ω;2))(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}))^{\ast} by 3.4, i.e., for any 𝒗𝒮w𝝂(Ω;2)\bm{v}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}), we need

(58) (𝔊w𝝂𝒖,𝔊w𝝂𝒗)L2(2;2×2)=(𝔇w𝝂𝒖,𝔇w𝝂𝒗)L2(2)+(𝔇w𝝂J𝒖,𝔇w𝝂J𝒗)L2(2;2),(\mathfrak{G}_{w}^{\bm{\nu}}\bm{u},\mathfrak{G}_{w}^{\bm{\nu}}\bm{v})_{L^{2}(\mathbb{R}^{2};\mathbb{R}^{2\times 2})}=(\mathfrak{D}_{w}^{-\bm{\nu}}\bm{u},\mathfrak{D}_{w}^{-\bm{\nu}}\bm{v})_{L^{2}(\mathbb{R}^{2})}+(\mathfrak{D}_{w}^{\bm{\nu}}J\bm{u},\mathfrak{D}_{w}^{\bm{\nu}}J\bm{v})_{L^{2}(\mathbb{R}^{2};\mathbb{R}^{2})},

where J=(0110)J=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}. First notice that eq. 58 holds for all functions 𝒗Cc(Ω;2)\bm{v}\in C_{c}^{\infty}(\Omega;\mathbb{R}^{2}) by 2.6 and the definitions of 𝔊w𝝂\mathfrak{G}_{w}^{\bm{\nu}} and 𝔇w𝝂\mathfrak{D}_{w}^{\bm{\nu}}. Now for any 𝒗𝒮w𝝂(Ω;2)\bm{v}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}), using 3.2 and 3.2, there exists a sequence 𝒗(n)Cc(Ω;2)\bm{v}^{(n)}\in C_{c}^{\infty}(\Omega;\mathbb{R}^{2}) such that 𝒢w𝝂𝒗(n)𝔊w𝝂𝒗\mathcal{G}_{w}^{\bm{\nu}}\bm{v}^{(n)}\to\mathfrak{G}_{w}^{\bm{\nu}}\bm{v} in L2(2,2×2)L^{2}(\mathbb{R}^{2},\mathbb{R}^{2\times 2}) and 𝒟w±𝝂𝒗(n)𝔇w±𝝂𝒗\mathcal{D}_{w}^{\pm\bm{\nu}}\bm{v}^{(n)}\to\mathfrak{D}_{w}^{\pm\bm{\nu}}\bm{v} in L2(2)L^{2}(\mathbb{R}^{2}). Then eq. 58 holds for 𝒗𝒮w𝝂(Ω;2)\bm{v}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}) by taking limits. ∎

Lemma 3.6.

For 𝐮𝒮w𝛎(Ω;3)\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}),

(59) 𝔇w𝝂𝔊w𝝂𝒖=𝔊w𝝂𝔇w𝝂𝒖w𝝂w𝝂𝒖.\mathfrak{D}_{w}^{-\bm{\nu}}\mathfrak{G}_{w}^{\bm{\nu}}\bm{u}=\mathfrak{G}_{w}^{\bm{\nu}}\mathfrak{D}_{w}^{-\bm{\nu}}\bm{u}-\mathfrak{C}_{w}^{-\bm{\nu}}\mathfrak{C}_{w}^{\bm{\nu}}\bm{u}.
Proof.

Notice that the left hand side and the right hand side of (59) are understood as elements in (𝒮w𝝂(Ω;3))(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}))^{\ast}. The proof is similar to the proof of 3.5 by using 2.7. ∎

4. Nonlocal Poincaré inequality for integrable kernels with compact support

In this section, we prove the Poincaré inequality for integrable kernels with compact support. Throughout this section, we assume that wL1(d)w\in L^{1}(\mathbb{R}^{d}) and

(60) suppwK for some compact set Kd.\text{supp}\ w\subset K\text{ for some compact set }K\subset\mathbb{R}^{d}.

We also assume for the rest of this paper that Ωd\Omega\subset\mathbb{R}^{d} is a bounded domain. Our major result in this section is the Poincaré inequality stated below.

Theorem 4.1 (Poincaré inequality for integrable kernels with compact support).

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded domain. Assume that wL1(d)w\in L^{1}(\mathbb{R}^{d}) and satisfies (60), then the Poincaré inequality holds for u𝒮w𝛎(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega). That is, there exists a constant Π=Π(w,Ω,𝛎)>0\Pi=\Pi(w,\Omega,\bm{\nu})>0 such that

(61) uL2(Ω)Π𝔊w𝝂uL2(d;d),u𝒮w𝝂(Ω),\|u\|_{L^{2}(\Omega)}\leq\Pi\|\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})},\quad\forall u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega),

where 𝔊w𝛎u=𝒢w𝛎uL2(d;d)\mathfrak{G}^{\bm{\nu}}_{w}u=\mathcal{G}_{w}^{\bm{\nu}}u\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}).

In the following, we establish necessary ingredients for the proof of 4.1. We first give a list of new notations that will be used in this section.

  • For the kernel ww, let c^\hat{c} be a constant depending only on ww defined as

    (62) c^:={z1>0}z1|𝒛|w(𝒛)𝑑𝒛>0.\hat{c}:=\int_{\{z_{1}>0\}}\frac{z_{1}}{|\bm{z}|}w(\bm{z})d\bm{z}>0.
  • For a fixed unit vector 𝝂d\bm{\nu}\in\mathbb{R}^{d}, define a vector-valued function 𝜷𝝂:dd{\bm{\beta}}^{\bm{\nu}}:\mathbb{R}^{d}\to\mathbb{R}^{d} by

    (63) 𝜷𝝂(𝒛):=χ𝝂(𝒛)𝒛|𝒛|w(𝒛).{\bm{\beta}}^{\bm{\nu}}(\bm{z}):=\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}w(\bm{z}).
  • Let 𝒄𝝂\bm{c}_{\bm{\nu}} be a constant vector depending only on ww and 𝝂\bm{\nu} defined as

    (64) 𝒄𝝂:=d𝜷𝝂(𝒛)𝑑𝒛=dχ𝝂(𝒛)𝒛|𝒛|w(𝒛)𝑑𝒛.\bm{c}_{\bm{\nu}}:=\int_{\mathbb{R}^{d}}{\bm{\beta}}^{\bm{\nu}}(\bm{z})d\bm{z}=\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}}{|\bm{z}|}w(\bm{z})d\bm{z}.
  • Let 𝑭,𝑮:dd\bm{F},\bm{G}:\mathbb{R}^{d}\to\mathbb{R}^{d} be vector-valued functions. Define their convolution as the following scalar-valued function

    (𝑭𝑮)(𝒙):=d𝑭(𝒙𝒚)𝑮(𝒚)𝑑𝒚.(\bm{F}*\bm{G})(\bm{x}):=\int_{\mathbb{R}^{d}}\bm{F}(\bm{x}-\bm{y})\cdot\bm{G}(\bm{y})d\bm{y}.

There are a few properties related to the above defined quantities. We list these properties here without proof since they are not hard to see.

  • For any d×dd\times d orthogonal matrix RR,

    (65) 𝒄R𝝂=R𝒄𝝂.\bm{c}_{R\bm{\nu}}=R\bm{c}_{\bm{\nu}}.

    Consequently,

    (66) 𝒄𝝂=c^𝝂.\bm{c}_{\bm{\nu}}=\hat{c}\bm{\nu}.
  • From Young’s convolution inequality, for 1p1\leq p\leq\infty we have

    𝑭𝑮Lp(d)𝑭L1(d;d)𝑮Lp(d;d).\|\bm{F}*\bm{G}\|_{L^{p}(\mathbb{R}^{d})}\leq\|\bm{F}\|_{L^{1}(\mathbb{R}^{d};\mathbb{R}^{d})}\|\bm{G}\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}.

With the integrability assumption of ww, we notice that 𝒢w𝝂\mathcal{G}_{w}^{\bm{\nu}} is well-defined on Lp(d)L^{p}(\mathbb{R}^{d}), 𝒟w𝝂\mathcal{D}^{\bm{\nu}}_{w} is well-defined on Lp(d;d)L^{p}(\mathbb{R}^{d};\mathbb{R}^{d}), and the limiting process in the definition of 𝒢w𝝂\mathcal{G}_{w}^{\bm{\nu}} and 𝒟w𝝂\mathcal{D}^{\bm{\nu}}_{w} can be dropped. In addition, each of 𝒢w𝝂\mathcal{G}_{w}^{\bm{\nu}} and 𝒟w𝝂\mathcal{D}^{\bm{\nu}}_{w} can be rewritten as a convolution operator plus a multiplication operator using the notations 𝜷𝝂\bm{\beta}^{\bm{\nu}} and 𝒄𝝂\bm{c}_{\bm{\nu}} above. These lead to a stronger version of integration by parts formula and an equivalent characterization of 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega).

Proposition 4.1.

The following statements are true.
(1) For 1p1\leq p\leq\infty, 𝒢w𝛎:Lp(d)Lp(d;d)\mathcal{G}_{w}^{\bm{\nu}}:L^{p}(\mathbb{R}^{d})\to L^{p}(\mathbb{R}^{d};\mathbb{R}^{d}), 𝒟w𝛎:Lp(d;d)Lp(d)\mathcal{D}^{\bm{\nu}}_{w}:L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})\to L^{p}(\mathbb{R}^{d}), and 𝒞w𝛎:Lp(3;3)Lp(3;3)\mathcal{C}^{\bm{\nu}}_{w}:L^{p}(\mathbb{R}^{3};\mathbb{R}^{3})\to L^{p}(\mathbb{R}^{3};\mathbb{R}^{3}) are bounded operators with estimates

𝒢w𝝂uLp(d;d)2wL1(d)uLp(d),\displaystyle\|\mathcal{G}_{w}^{\bm{\nu}}u\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}\leq 2\|w\|_{L^{1}(\mathbb{R}^{d})}\|u\|_{L^{p}(\mathbb{R}^{d})},
𝒟w𝝂𝒗Lp(d)CwL1(d)𝒗Lp(d;d),\displaystyle\|\mathcal{D}^{\bm{\nu}}_{w}\bm{v}\|_{L^{p}(\mathbb{R}^{d})}\leq C\|w\|_{L^{1}(\mathbb{R}^{d})}\|\bm{v}\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})},
𝒞w𝝂𝒗Lp(3;3)CwL1(3)𝒗Lp(3;3).\displaystyle\|\mathcal{C}^{\bm{\nu}}_{w}\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3})}\leq C\|w\|_{L^{1}(\mathbb{R}^{3})}\|\bm{v}\|_{L^{p}(\mathbb{R}^{3};\mathbb{R}^{3})}.

for some C=C(d)>0C=C(d)>0. Moreover, for uLp(d)u\in L^{p}(\mathbb{R}^{d}) and 𝐯Lp(d;d)\bm{v}\in L^{p}(\mathbb{R}^{d};\mathbb{R}^{d}),

(67) 𝒢w𝝂u(𝒙)=d𝜷𝝂(𝒚𝒙)u(𝒚)𝑑𝒚𝒄𝝂u(𝒙),a.e.𝒙d,\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})=\int_{\mathbb{R}^{d}}\bm{\beta}^{\bm{\nu}}(\bm{y}-\bm{x})u(\bm{y})d\bm{y}-\bm{c}_{\bm{\nu}}u(\bm{x}),\quad\text{a.e.}\ \bm{x}\in\mathbb{R}^{d},
(68) 𝒟w𝝂𝒗(𝒙)=𝜷𝝂𝒗(𝒙)+𝒄𝝂𝒗(𝒙),a.e.𝒙d.\mathcal{D}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})=-\bm{\beta}^{\bm{\nu}}*\bm{v}(\bm{x})+\bm{c}_{\bm{\nu}}\cdot\bm{v}(\bm{x}),\quad\text{a.e.}\ \bm{x}\in\mathbb{R}^{d}.

(2) Suppose uLp(d)u\in L^{p}(\mathbb{R}^{d}) and 𝐯Lp(d;d)\bm{v}\in L^{p^{\prime}}(\mathbb{R}^{d};\mathbb{R}^{d}), where p=pp1p^{\prime}=\frac{p}{p-1} and 1p1\leq p\leq\infty (p=p^{\prime}=\infty for p=1p=1 and p=1p^{\prime}=1 for p=p=\infty). Then

d𝒢w𝝂u(𝒙)𝒗(𝒙)𝑑𝒙=du(𝒙)𝒟w𝝂𝒗(𝒙)𝑑𝒙.\int_{\mathbb{R}^{d}}\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})\cdot\bm{v}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}.

Similarly, for 𝐮Lp(3;3)\bm{u}\in L^{p}(\mathbb{R}^{3};\mathbb{R}^{3}) and 𝐯Lp(3;3)\bm{v}\in L^{p^{\prime}}(\mathbb{R}^{3};\mathbb{R}^{3}),

3𝒞w𝝂𝒖(𝒙)𝒗(𝒙)𝑑𝒙=3𝒖(𝒙)𝒞w𝝂𝒗(𝒙)𝑑𝒙.\int_{\mathbb{R}^{3}}\mathcal{C}^{\bm{\nu}}_{w}\bm{u}(\bm{x})\cdot\bm{v}(\bm{x})d\bm{x}=\int_{\mathbb{R}^{3}}\bm{u}(\bm{x})\cdot\mathcal{C}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}.
Proof.

Notice that since wL1(d)w\in L^{1}(\mathbb{R}^{d}), the integrand in eq. 2 is Lebesgue integrable on d\mathbb{R}^{d} for uLp(d)u\in L^{p}(\mathbb{R}^{d}), and therefore the limiting process can be dropped. The characterizations (67) and (68) follow directly from 2.1, (63) and (64). For instance, (68) holds as

𝒟w𝝂𝒗(𝒙)=d𝒚𝒙|𝒚𝒙|χ𝝂(𝒙𝒚)(𝒗(𝒚)𝒗(𝒙))w(𝒚𝒙)𝑑𝒚=dχ𝝂(𝒙𝒚)𝒙𝒚|𝒙𝒚|w(𝒚𝒙)𝒗(𝒚)𝑑𝒚+dχ𝝂(𝒛)𝒛𝒗(𝒙)|𝒛|w(𝒛)𝑑𝒛=𝜷𝝂𝒗(𝒙)+𝒄𝝂𝒗(𝒙)inLp(d),\begin{split}\mathcal{D}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})&=\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot\chi_{\bm{\nu}}(\bm{x}-\bm{y})(\bm{v}(\bm{y})-\bm{v}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y}\\ &=-\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{x}-\bm{y})\frac{\bm{x}-\bm{y}}{|\bm{x}-\bm{y}|}w(\bm{y}-\bm{x})\cdot\bm{v}(\bm{y})d\bm{y}+\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{z})\frac{\bm{z}\cdot\bm{v}(\bm{x})}{|\bm{z}|}w(\bm{z})d\bm{z}\\ &=-{\bm{\beta}}^{\bm{\nu}}\ast\bm{v}(\bm{x})+\bm{c}_{\bm{\nu}}\cdot\bm{v}(\bm{x})\quad\text{in}\ L^{p}(\mathbb{R}^{d}),\end{split}

where the convolution term is well-defined thanks to the Young’s convolution inequality and the fact that wL1(d)w\in L^{1}(\mathbb{R}^{d}). Suppose 1<p<1<p<\infty. For p=1p=1 and p=p=\infty we can show the estimate similarly. Using Holder’s inequality, we obtain

d|𝒢w𝝂u(𝒙)|p𝑑𝒙=d|d(u(𝒚)u(𝒙))𝒚𝒙|𝒚𝒙|w(𝒚𝒙)𝑑𝒚|p𝑑𝒙\displaystyle\quad\,\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|^{p}d\bm{x}=\int_{\mathbb{R}^{d}}\left|\int_{\mathbb{R}^{d}}(u(\bm{y})-u(\bm{x}))\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})d\bm{y}\right|^{p}d\bm{x}
d(d|u(𝒚)u(𝒙)|w(𝒚𝒙)𝑑𝒚)p𝑑𝒙\displaystyle\leq\int_{\mathbb{R}^{d}}\left(\int_{\mathbb{R}^{d}}|u(\bm{y})-u(\bm{x})|w(\bm{y}-\bm{x})d\bm{y}\right)^{p}d\bm{x}
d(dw(𝒚𝒙)𝑑𝒚)pp(dw(𝒚𝒙)|u(𝒚)u(𝒙)|p𝑑𝒚)𝑑𝒙\displaystyle\leq\int_{\mathbb{R}^{d}}\left(\int_{\mathbb{R}^{d}}w(\bm{y}-\bm{x})d\bm{y}\right)^{\frac{p}{p^{\prime}}}\left(\int_{\mathbb{R}^{d}}w(\bm{y}-\bm{x})|u(\bm{y})-u(\bm{x})|^{p}d\bm{y}\right)d\bm{x}
2p1wL1(d)ppddw(𝒚𝒙)(|u|p(𝒚)+|u|p(𝒙))𝑑𝒚𝑑𝒙\displaystyle\leq 2^{p-1}\|w\|_{L^{1}(\mathbb{R}^{d})}^{\frac{p}{p^{\prime}}}\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}w(\bm{y}-\bm{x})(|u|^{p}(\bm{y})+|u|^{p}(\bm{x}))d\bm{y}d\bm{x}
=2p1wL1(d)pp(d|u|p(𝒚)𝑑𝒚dw(𝒚𝒙)𝑑𝒙+d|u|p(𝒙)𝑑𝒙dw(𝒚𝒙)𝑑𝒚)\displaystyle=2^{p-1}\|w\|_{L^{1}(\mathbb{R}^{d})}^{\frac{p}{p^{\prime}}}\left(\int_{\mathbb{R}^{d}}|u|^{p}(\bm{y})d\bm{y}\int_{\mathbb{R}^{d}}w(\bm{y}-\bm{x})d\bm{x}+\int_{\mathbb{R}^{d}}|u|^{p}(\bm{x})d\bm{x}\int_{\mathbb{R}^{d}}w(\bm{y}-\bm{x})d\bm{y}\right)
2pwL1(d)puLp(d)p,\displaystyle\leq 2^{p}\|w\|_{L^{1}(\mathbb{R}^{d})}^{p}\|u\|_{L^{p}(\mathbb{R}^{d})}^{p},

where p=p/(p1)p^{\prime}=p/(p-1). This shows (1). The estimates for 𝒟w𝝂𝒗\mathcal{D}^{\bm{\nu}}_{w}\bm{v} and 𝒞w𝝂𝒗\mathcal{C}^{\bm{\nu}}_{w}\bm{v} can be shown similarly.

The integration by parts formulas in (2) can be shown by a change of integration order via Fubini’s theorem, for example,

d𝒢w𝝂u(𝒙)𝒗(𝒙)𝑑𝒙\displaystyle\int_{\mathbb{R}^{d}}\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})\cdot\bm{v}(\bm{x})d\bm{x} =ddχ𝝂(𝒚𝒙)(u(𝒚)u(𝒙))𝒚𝒙|𝒚𝒙|w(𝒚𝒙)𝒗(𝒙)𝑑𝒚𝑑𝒙\displaystyle=\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})(u(\bm{y})-u(\bm{x}))\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})\cdot\bm{v}(\bm{x})d\bm{y}d\bm{x}
=ddχ𝝂(𝒙𝒚)u(𝒙)𝒚𝒙|𝒚𝒙|w(𝒚𝒙)𝒗(𝒚)d𝒚d𝒙\displaystyle=\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}-\chi_{\bm{\nu}}(\bm{x}-\bm{y})u(\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})\cdot\bm{v}(\bm{y})d\bm{y}d\bm{x}
ddχ𝝂(𝒚𝒙)u(𝒙)𝒚𝒙|𝒚𝒙|w(𝒚𝒙)𝒗(𝒙)𝑑𝒚𝑑𝒙\displaystyle\quad\quad-\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})u(\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})\cdot\bm{v}(\bm{x})d\bm{y}d\bm{x}
=du(𝒙)𝒟w𝝂𝒗(𝒙)𝑑𝒙.\displaystyle=-\int_{\mathbb{R}^{d}}u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}.

Here Fubini’s theorem is justified since |u|(𝒚)|𝒗|(𝒙)w(𝒚𝒙)L1(d×d)|u|(\bm{y})|\bm{v}|(\bm{x})w(\bm{y}-\bm{x})\in L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}) for uLp(d)u\in L^{p}(\mathbb{R}^{d}) and 𝒗Lp(d;d)\bm{v}\in L^{p^{\prime}}(\mathbb{R}^{d};\mathbb{R}^{d}). Indeed,

dd|u|(𝒚)|𝒗|(𝒙)w(𝒚𝒙)𝑑𝒚𝑑𝒙\displaystyle\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}|u|(\bm{y})|\bm{v}|(\bm{x})w(\bm{y}-\bm{x})d\bm{y}d\bm{x} =d|u|(𝒚)(w|𝒗|)(𝒚)𝑑𝒚\displaystyle=\int_{\mathbb{R}^{d}}|u|(\bm{y})(w*|\bm{v}|)(\bm{y})d\bm{y}
uLp(d)w|𝒗|Lp(d)\displaystyle\leq\|u\|_{L^{p}(\mathbb{R}^{d})}\|w*|\bm{v}|\|_{L^{p^{\prime}}(\mathbb{R}^{d})}
CuLp(d)wL1(d)𝒗Lp(d;d)<,\displaystyle\leq C\|u\|_{L^{p}(\mathbb{R}^{d})}\|w\|_{L^{1}(\mathbb{R}^{d})}\|\bm{v}\|_{L^{p^{\prime}}(\mathbb{R}^{d};\mathbb{R}^{d})}<\infty,

where we used Young’s convolution inequality and CC is a constant only depending on the dimension dd. The second integration by parts formula in (2) can be shown similarly. ∎

An immediate result from 4.1 is the following equivalent characterization of 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega).

Corollary 4.1 (An equivalent characterization of 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega)).

With the integrability assumption of ww, 𝒢w𝛎u=𝔊w𝛎u\mathcal{G}^{\bm{\nu}}_{w}u=\mathfrak{G}^{\bm{\nu}}_{w}u for uL2(d)u\in L^{2}(\mathbb{R}^{d}) with u=0u=0 a.e. in Ωc\Omega^{c}, and the function space 𝒮w𝛎(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) defined by eq. 36 satisfies

𝒮w𝝂(Ω)={uL2(d):u=0 a.e. in Ωc}.\mathcal{S}_{w}^{\bm{\nu}}(\Omega)=\{u\in L^{2}(\mathbb{R}^{d}):u=0\text{ a.e. in }\Omega^{c}\}.

We next show a crucial result for proving the Poincaré inequality. It claims that the operator 𝒢w𝝂\mathcal{G}_{w}^{\bm{\nu}} restricted to 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) is injective.

Proposition 4.2.

Assume that wL1(d)w\in L^{1}(\mathbb{R}^{d}) and satisfies (60). If u𝒮w𝛎(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega) satisfies 𝒢w𝛎u(𝐱)=0\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})=0 for a.e. 𝐱d\bm{x}\in\mathbb{R}^{d}, then u0u\equiv 0.

Proof.

Note that the nonlocal integration by parts formula in 4.1 also holds for complex-valued functions. That is, for uLp(d;)u\in L^{p}(\mathbb{R}^{d};\mathbb{C}) and 𝒗Lp(d;d)\bm{v}\in L^{p^{\prime}}(\mathbb{R}^{d};\mathbb{C}^{d}) with pp and pp^{\prime} given by 4.1,

(69) (𝒢w𝝂u,𝒗)L2(d;d)=(u,𝒟w𝝂𝒗)L2(d;),(\mathcal{G}^{\bm{\nu}}_{w}u,\bm{v})_{L^{2}(\mathbb{R}^{d};\mathbb{C}^{d})}=-(u,\mathcal{D}^{-\bm{\nu}}_{w}\bm{v})_{L^{2}(\mathbb{R}^{d};\mathbb{C})},

where the L2L^{2}-norm is given by

(𝑭,𝑮)L2(d;n)=d𝑭(𝒙)T𝑮(𝒙)¯𝑑𝒙,𝑭,𝑮L2(d;n),n=1,d.(\bm{F},\bm{G})_{L^{2}(\mathbb{R}^{d};\mathbb{C}^{n})}=\int_{\mathbb{R}^{d}}\bm{F}(\bm{x})^{T}\overline{\bm{G}(\bm{x})}d\bm{x},\quad\forall\bm{F},\bm{G}\in L^{2}(\mathbb{R}^{d};\mathbb{C}^{n}),\ n=1,d.

Thus, for any 𝝋\bm{\varphi} in the Schwartz space 𝒮(d;d)L2(d;d)\mathscr{S}(\mathbb{R}^{d};\mathbb{C}^{d})\subset L^{2}(\mathbb{R}^{d};\mathbb{C}^{d}), we have

(70) 0\displaystyle 0 =(𝒢w𝝂u,𝝋)L2(d;d)=(u,𝒟w𝝂𝝋)L2(d;)\displaystyle=(\mathcal{G}_{w}^{\bm{\nu}}u,\bm{\varphi})_{L^{2}(\mathbb{R}^{d};\mathbb{C}^{d})}=-(u,\mathcal{D}^{-\bm{\nu}}_{w}\bm{\varphi})_{L^{2}(\mathbb{R}^{d};\mathbb{C})}
=(u,(𝒟w𝝂𝝋))L2(d;)=(u^,(𝝀w𝝂)T𝝋^)L2(d;),\displaystyle=-(\mathcal{F}u,\mathcal{F}(\mathcal{D}^{-\bm{\nu}}_{w}\bm{\varphi}))_{L^{2}(\mathbb{R}^{d};\mathbb{C})}=-(\hat{u},({\bm{\lambda}}_{w}^{-\bm{\nu}})^{T}\hat{\bm{\varphi}})_{L^{2}(\mathbb{R}^{d};\mathbb{C})},

where u^=uL2(d;)\hat{u}=\mathcal{F}u\in L^{2}(\mathbb{R}^{d};\mathbb{C}) since uL2(d)u\in L^{2}(\mathbb{R}^{d}) and 𝒟w𝝂𝝋L2(d;)\mathcal{D}^{-\bm{\nu}}_{w}\bm{\varphi}\in L^{2}(\mathbb{R}^{d};\mathbb{C}) by eq. 69.

Since L2(d;)𝒮(d;)L^{2}(\mathbb{R}^{d};\mathbb{C})\subset\mathscr{S}^{\prime}(\mathbb{R}^{d};\mathbb{C}), we view u^=u𝒮(d;)\hat{u}=\mathcal{F}u\in\mathscr{S}^{\prime}(\mathbb{R}^{d};\mathbb{C}) as a tempered distribution. Now we prove the following claim:

(71) u^,ϕ=0,ϕCc(d\{𝟎};).\langle\hat{u},\phi\rangle=0,\quad\forall\phi\in C_{c}^{\infty}(\mathbb{R}^{d}\backslash\{\bm{0}\};\mathbb{C}).

Let 𝝋:dd\bm{\varphi}:\mathbb{R}^{d}\to\mathbb{C}^{d} be defined as

𝝋:=1(𝝀w𝝂¯|𝝀w𝝂|2ϕ).\bm{\varphi}:=\mathcal{F}^{-1}\left(\frac{\overline{{\bm{\lambda}}_{w}^{-\bm{\nu}}}}{|{\bm{\lambda}}_{w}^{-\bm{\nu}}|^{2}}\phi\right).

Since |𝝀w𝝂(𝝃)|>0|{\bm{\lambda}}_{w}^{-\bm{\nu}}(\bm{\xi})|>0 for 𝝃𝟎\bm{\xi}\neq\bm{0} by 2.2 and ϕ(𝝃)=0\phi(\bm{\xi})=0 in a neighborhood of 𝝃=𝟎\bm{\xi}=\bm{0}, 𝝀w𝝂(𝝃)¯ϕ(𝝃)/|𝝀w𝝂(𝝃)|2\overline{{\bm{\lambda}}_{w}^{-\bm{\nu}}(\bm{\xi})}\phi(\bm{\xi})/|{\bm{\lambda}}_{w}^{-\bm{\nu}}(\bm{\xi})|^{2} is a well-defined vector-valued function on d\mathbb{R}^{d}. Moreover, 𝝀w𝝂(𝝃)¯ϕ(𝝃)/|𝝀w𝝂(𝝃)|2Cc(d\{𝟎},d)𝒮(d;d)\overline{{\bm{\lambda}}_{w}^{-\bm{\nu}}(\bm{\xi})}\phi(\bm{\xi})/|{\bm{\lambda}}_{w}^{-\bm{\nu}}(\bm{\xi})|^{2}\in C_{c}^{\infty}(\mathbb{R}^{d}\backslash\{\bm{0}\},\mathbb{C}^{d})\subset\mathscr{S}(\mathbb{R}^{d};\mathbb{C}^{d}) since 𝝀w𝝂C(d;d){\bm{\lambda}}_{w}^{-\bm{\nu}}\in C^{\infty}(\mathbb{R}^{d};\mathbb{C}^{d}) by 2.3 and ϕCc(d\{𝟎},)\phi\in C_{c}^{\infty}(\mathbb{R}^{d}\backslash\{\bm{0}\},\mathbb{C}). Hence 𝝋𝒮(d;d)\bm{\varphi}\in\mathscr{S}(\mathbb{R}^{d};\mathbb{C}^{d}) since \mathcal{F} is an isomorphism on 𝒮(d;d)\mathscr{S}(\mathbb{R}^{d};\mathbb{C}^{d}). Observing that 𝝀w𝝂(𝝃)T𝝋^(𝝃)=ϕ(𝝃){\bm{\lambda}}_{w}^{-\bm{\nu}}(\bm{\xi})^{T}\hat{\bm{\varphi}}(\bm{\xi})=\phi(\bm{\xi}), the claim follows from eq. 70.

Now from the claim, we have suppu^{𝟎}\text{supp}\ \hat{u}\subset\{\bm{0}\}. Then by Corollary 2.4.2 in [28], uu is a polynomial, i.e.,

u(𝒙)=|α|kaαxαu(\bm{x})=\sum_{|\alpha|\leq k}a_{\alpha}x^{\alpha}

for some nonnegative integer kk and real numbers aαa_{\alpha} for |α|k|\alpha|\leq k. Since u=0u=0 in Ωc\Omega^{c}, it follows u0u\equiv 0. ∎

The last ingredient of the proof of the Poincaré inequality is the weak lower semicontinuity of the Dirichlet integral d|𝒢w𝝂u(𝒙)|2𝑑𝒙\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|^{2}d\bm{x}. To establish this result, we need Proposition A.3 in [6] which is stated as a lemma below.

Lemma 4.1 (Proposition A.3 in [6]).

Let Ωm\Omega\subset\mathbb{R}^{m} be bounded open, and let H:s{±}H:\mathbb{R}^{s}\to\mathbb{R}\cup\{\pm\infty\} be convex, lower semicontinuous and bounded below. Let θj,θL1(Ω;s)\theta_{j},\,\theta\in L^{1}(\Omega;\mathbb{R}^{s}) with θjθ\theta_{j}\stackrel{{\scriptstyle\ast}}{{\rightharpoonup}}\theta (i.e., Ωθjϕ𝑑𝐱Ωθϕ𝑑𝐱\int_{\Omega}\theta_{j}\phi d\bm{x}\to\int_{\Omega}\theta\phi d\bm{x} for all ϕCc(Ω)\phi\in C_{c}(\Omega)). Then

ΩH(θ(𝒙))𝑑𝒙lim infjΩH(θj(𝒙))𝑑𝒙.\int_{\Omega}H(\theta(\bm{x}))d\bm{x}\leq\liminf_{j\to\infty}\int_{\Omega}H(\theta_{j}(\bm{x}))d\bm{x}.
Proposition 4.3.

Suppose that {un}\{u_{n}\} converges weakly to uu in 𝒮w𝛎(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega). Then

(72) d|𝒢w𝝂u(𝒙)|2𝑑𝒙lim infnd|𝒢w𝝂un(𝒙)|2𝑑𝒙.\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|^{2}d\bm{x}\leq\liminf_{n\to\infty}\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x})|^{2}d\bm{x}.
Proof.

Let H(𝒙):=|𝒙|2H(\bm{x}):=|\bm{x}|^{2}. Then HH is convex, continuous and bounded below. Let θn(𝒙):=𝒢w𝝂un(𝒙)\theta_{n}(\bm{x}):=\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x}) and θ(𝒙):=𝒢w𝝂u(𝒙)\theta(\bm{x}):=\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x}). Then for any open and precompact set DdD\subset\mathbb{R}^{d}, θn,θL1(D;d)\theta_{n},\theta\in L^{1}(D;\mathbb{R}^{d}) because

D|𝒢w𝝂u(𝒙)|𝑑𝒙(D|𝒢w𝝂u(𝒙)|2𝑑𝒙)12(μ(D))12<.\int_{D}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|d\bm{x}\leq\left(\int_{D}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|^{2}d\bm{x}\right)^{\frac{1}{2}}(\mu(D))^{\frac{1}{2}}<\infty.

For any ϕCc(D)\phi\in C_{c}(D), 1kd1\leq k\leq d, define a linear functional Fϕk:𝒮w𝝂(Ω)F^{k}_{\phi}:\mathcal{S}_{w}^{\bm{\nu}}(\Omega)\to\mathbb{R} by

Fϕk(u):=D[𝒢w𝝂u(𝒙)]kϕ𝑑𝒙,F^{k}_{\phi}(u):=\int_{D}[\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})]_{k}\phi d\bm{x},

where [𝒢w𝝂u(𝒙)]k[\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})]_{k} denotes the kk-th component of 𝒢w𝝂\mathcal{G}_{w}^{\bm{\nu}}. Then FϕkF^{k}_{\phi} is a bounded linear functional since

|Fϕk(u)|(D|[𝒢w𝝂u(𝒙)]k|2𝑑𝒙)12(D|ϕ(𝒙)|2𝑑𝒙)12Cu𝒮w𝝂(Ω).\displaystyle\left|F^{k}_{\phi}(u)\right|\leq\left(\int_{D}|[\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})]_{k}|^{2}d\bm{x}\right)^{\frac{1}{2}}\cdot\left(\int_{D}|\phi(\bm{x})|^{2}d\bm{x}\right)^{\frac{1}{2}}\leq C\|u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)}.

Now since unuu_{n}\rightharpoonup u in 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega), we have Fϕk(un)Fϕk(u)F^{k}_{\phi}(u_{n})\to F^{k}_{\phi}(u) as nn\to\infty. Therefore θjθL1(D;d)\theta_{j}\stackrel{{\scriptstyle\ast}}{{\rightharpoonup}}\theta\in L^{1}(D;\mathbb{R}^{d}) and this yields

D|𝒢w𝝂u(𝒙)|2𝑑𝒙lim infnD|𝒢w𝝂un(𝒙)|2𝑑𝒙lim infnd|𝒢w𝝂un(𝒙)|2𝑑𝒙.\int_{D}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|^{2}d\bm{x}\leq\liminf_{n\to\infty}\int_{D}|\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x})|^{2}d\bm{x}\leq\liminf_{n\to\infty}\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x})|^{2}d\bm{x}.

by 4.1. Since DdD\subset\mathbb{R}^{d} is arbitrary, eq. 72 is true. ∎

Finally, we are ready to prove 4.1.

Proof of 4.1.

We argue by contradiction. Suppose there exists {un}𝒮w𝝂(Ω)\{u_{n}\}\subset\mathcal{S}_{w}^{\bm{\nu}}(\Omega) with unL2(Ω)=1\|u_{n}\|_{L^{2}(\Omega)}=1 such that 𝒢w𝝂unL2(d;d)0\|\mathcal{G}_{w}^{\bm{\nu}}u_{n}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\to 0. Then un𝒮w𝝂(Ω)\|u_{n}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)} is bounded. Since 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) is a Hilbert space by 3.1, there exists a subsequence of {un}\{u_{n}\}, still denoted by {un}\{u_{n}\} for convenience, that convergences weakly to some u𝒮w𝝂(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega).

In the first step, we show u=0u=0, i.e., un0u_{n}\rightharpoonup 0 in 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega). By the weakly lower semi-continuous result in 4.3, we have

d|𝒢w𝝂u(𝒙)|2𝑑𝒙lim infnd|𝒢w𝝂un(𝒙)|2𝑑𝒙.\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|^{2}d\bm{x}\leq\liminf_{n\to\infty}\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x})|^{2}d\bm{x}.

Now that 𝒢w𝝂unL2(d;d)0\|\mathcal{G}_{w}^{\bm{\nu}}u_{n}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\to 0,

d|𝒢w𝝂u(𝒙)|2𝑑𝒙=0,\int_{\mathbb{R}^{d}}|\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})|^{2}d\bm{x}=0,

and thus 𝒢w𝝂u(𝒙)=0\mathcal{G}_{w}^{\bm{\nu}}u(\bm{x})=0 for a.e. 𝒙d\bm{x}\in\mathbb{R}^{d}. By 4.2, u0u\equiv 0 and the first step is done.

In the second step, we show unu_{n} converges to 0 strongly in L2L^{2}, which contradicts the assumption unL2(Ω)=1\|u_{n}\|_{L^{2}(\Omega)}=1.

Using the integration by parts formula and the characterizations of 𝒟w𝝂\mathcal{D}^{-\bm{\nu}}_{w} and 𝒢w𝝂\mathcal{G}_{w}^{\bm{\nu}} consecutively in 4.1, it follows that (recall that un=0u_{n}=0 in Ωc\Omega^{c}),

(73) 𝒢w𝝂unL2(d;d)2\displaystyle\|\mathcal{G}_{w}^{\bm{\nu}}u_{n}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}^{2}
=\displaystyle= dun(𝒙)𝒟w𝝂𝒢w𝝂un(𝒙)𝑑𝒙\displaystyle-\int_{\mathbb{R}^{d}}u_{n}(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\circ\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x})d\bm{x}
=\displaystyle= dun(𝒙)(𝜷𝝂𝒢w𝝂un(𝒙)𝒄𝝂𝒢w𝝂un(𝒙))𝑑𝒙\displaystyle\int_{\mathbb{R}^{d}}u_{n}(\bm{x})\left({\bm{\beta}}^{\bm{\nu}}\ast\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x})-\bm{c}_{\bm{\nu}}\cdot\mathcal{G}_{w}^{\bm{\nu}}u_{n}(\bm{x})\right)d\bm{x}
=\displaystyle= dun(𝒙)(𝜷𝝂𝒢w𝝂un)(𝒙)𝑑𝒙\displaystyle\int_{\mathbb{R}^{d}}u_{n}(\bm{x})\left({\bm{\beta}}^{\bm{\nu}}\ast\mathcal{G}_{w}^{\bm{\nu}}u_{n}\right)(\bm{x})d\bm{x}
dun(𝒙)𝒄𝝂(d𝜷𝝂(𝒚𝒙)un(𝒚)𝑑𝒚𝒄𝝂un(𝒙))𝑑𝒙\displaystyle\qquad-\int_{\mathbb{R}^{d}}u_{n}(\bm{x})\bm{c}_{\bm{\nu}}\cdot\left(\int_{\mathbb{R}^{d}}{\bm{\beta}}^{\bm{\nu}}(\bm{y}-\bm{x})u_{n}(\bm{y})d\bm{y}-\bm{c}_{\bm{\nu}}u_{n}(\bm{x})\right)d\bm{x}
=\displaystyle= (un,𝜷𝝂𝒢w𝝂un)Ω+|𝒄𝝂|2unL2(Ω)2(un,Kun)Ω\displaystyle(u_{n},{\bm{\beta}}^{\bm{\nu}}*\mathcal{G}_{w}^{\bm{\nu}}u_{n})_{\Omega}+|\bm{c}_{\bm{\nu}}|^{2}\|u_{n}\|^{2}_{L^{2}(\Omega)}-(u_{n},Ku_{n})_{\Omega}

where (,)Ω(\cdot,\cdot)_{\Omega} denote the L2(Ω)L^{2}(\Omega) inner product and K:L2(Ω)L2(Ω)K:L^{2}(\Omega)\to L^{2}(\Omega) is defined as

Ku(𝒙):=Ω𝒄𝝂𝜷𝝂(𝒚𝒙)u(𝒚)𝑑𝒚.Ku(\bm{x}):=\int_{\Omega}\bm{c}_{\bm{\nu}}\cdot{\bm{\beta}}^{\bm{\nu}}(\bm{y}-\bm{x})u(\bm{y})d\bm{y}.

Note that KK is well-defined as |Ku||𝒄𝝂|w|u|L2(d)|Ku|\leq|\bm{c}_{\bm{\nu}}|w*|u|\in L^{2}(\mathbb{R}^{d}). Now notice that by Young’s convolution inequality

(un,𝜷𝝂𝒢w𝝂un)ΩunL2(Ω)𝜷𝝂𝒢w𝝂unL2(Ω)dwL1(d)𝒢w𝝂unL2(d;d)0.(u_{n},{\bm{\beta}}^{\bm{\nu}}*\mathcal{G}_{w}^{\bm{\nu}}u_{n})_{\Omega}\leq\|u_{n}\|_{L^{2}(\Omega)}\|{\bm{\beta}}^{\bm{\nu}}*\mathcal{G}_{w}^{\bm{\nu}}u_{n}\|_{L^{2}(\Omega)}\leq\sqrt{d}\|w\|_{L^{1}(\mathbb{R}^{d})}\|\mathcal{G}_{w}^{\bm{\nu}}u_{n}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\to 0.

as nn\to\infty. In addition, |𝒄𝝂|2=|c^|2>0|\bm{c}_{\bm{\nu}}|^{2}=|\hat{c}|^{2}>0 by eqs. 62 and 66. Therefore, if (un,Kun)Ω0(u_{n},Ku_{n})_{\Omega}\to 0, then we reach a contradiction since eq. 73 implies unL2(Ω)0\|u_{n}\|_{L^{2}(\Omega)}\to 0. In the following, we proceed to show (un,Kun)Ω0(u_{n},Ku_{n})_{\Omega}\to 0 as nn\to\infty.

Notice that by definition

Ku(𝒙)=Ω𝒄𝝂χ𝝂(𝒚𝒙)𝒚𝒙|𝒚𝒙|w(𝒚𝒙)u(𝒚)𝑑𝒚=Ωc^χ𝝂(𝒚𝒙)𝝂(𝒚𝒙)|𝒚𝒙|w(𝒚𝒙)u(𝒚)𝑑𝒚=:Ωk(𝒙𝒚)u(𝒚)d𝒚,\begin{split}Ku(\bm{x})&=\int_{\Omega}\bm{c}_{\bm{\nu}}\cdot\chi_{\bm{\nu}}(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})u(\bm{y})d\bm{y}\\ &=\int_{\Omega}\hat{c}\cdot\chi_{\bm{\nu}}(\bm{y}-\bm{x})\frac{\bm{\nu}\cdot(\bm{y}-\bm{x})}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})u(\bm{y})d\bm{y}\\ &=:\int_{\Omega}k(\bm{x}-\bm{y})u(\bm{y})d\bm{y},\end{split}

where we have used eq. 66 and k(𝒛):=c^χ𝝂(𝒛)𝝂(𝒛)|𝒛|w(𝒛)0k(\bm{z}):=\hat{c}\cdot\chi_{\bm{\nu}}(-\bm{z})\frac{\bm{\nu}\cdot(-\bm{z})}{|\bm{z}|}w(\bm{z})\geq 0. Notice that kL1(d)k\in L^{1}(\mathbb{R}^{d}) as wL1(d)w\in L^{1}(\mathbb{R}^{d}), it follows from Corollary 4.28 of [11] that K:L2(Ω)L2(Ω)K:L^{2}(\Omega)\to L^{2}(\Omega) is compact. From the first step we have un0u_{n}\rightharpoonup 0 in 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega) and thus un0u_{n}\rightharpoonup 0 in L2(d)L^{2}(\mathbb{R}^{d}). Thus Kun0Ku_{n}\to 0 in L2(Ω)L^{2}(\Omega) as K:L2(Ω)L2(Ω)K:L^{2}(\Omega)\to L^{2}(\Omega) is compact. Therefore,

|(un,Kun)Ω|unL2(Ω)KunL2(Ω)0,n.|(u_{n},Ku_{n})_{\Omega}|\leq\|u_{n}\|_{L^{2}(\Omega)}\|Ku_{n}\|_{L^{2}(\Omega)}\to 0,\ n\to\infty.

Hence the proof is completed. ∎

5. Nonlocal Poincaré inequality for general kernel functions

Our main goal in this section is to prove the Poincaré inequality for general kernel functions beyond the integrable and compactly supported ones used in Section 4. Throughout this section, we assume that the kernel function satisfies eq. 1 and the assumptions given as follows.

Assumption 5.1.

Assume that ww satisfies eq. 1, and either one of the following conditions holds true:

  1. (1)

    d|𝒙|w(𝒙)𝑑𝒙<\int_{\mathbb{R}^{d}}|\bm{x}|w(\bm{x})d\bm{x}<\infty;

  2. (2)

    there exists R>0R>0 such that w(𝒙)=c0|𝒙|d+αw(\bm{x})=\frac{c_{0}}{|\bm{x}|^{d+\alpha}} for some c0>0c_{0}>0 and α(0,1]\alpha\in(0,1] when |𝒙|>R|\bm{x}|>R.

We use w¯\overline{w} to denote the radial representation of ww, i.e., w¯:[0,)[0,)\overline{w}:[0,\infty)\to[0,\infty) satisfies w¯(|𝐱|)=w(𝐱)\overline{w}(|\bm{x}|)=w(\bm{x}) for 𝐱d\bm{x}\in\mathbb{R}^{d}.

Remark 5.1.

Notice that 5.1 covers many cases of kernel functions seen in the literature. For example, compactly supported kernels, the fractional kernel w(𝐱)=C|𝐱|dαw(\bm{x})=C|\bm{x}|^{-d-\alpha} used to study the Riesz fractional derivatives in [46, 47], as well as the tempered fractional kernel w(𝐱)=Ceλ|𝐱||𝐱|dαw(\bm{x})=Ce^{-\lambda|\bm{x}|}|\bm{x}|^{-d-\alpha} in [44].

Under 5.1, we can show the following result.

Theorem 5.1 (Poincaré inequality for general kernel functions).

Let Ω\Omega be a bounded domain with a continuous boundary. Under Assumption 5.1, the Poincaré inequality holds, i.e., there exists a constant Π=Π(w,Ω,𝛎)>0\Pi=\Pi(w,\Omega,\bm{\nu})>0 such that

(74) uL2(Ω)Π𝔊w𝝂uL2(d;d),u𝒮w𝝂(Ω).\|u\|_{L^{2}(\Omega)}\leq\Pi\|\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})},\quad\forall u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega).

For general kernel functions, we do not have a direct analogue of eq. 73 since the single integral defining 𝒟w𝝂\mathcal{D}^{\bm{\nu}}_{w} cannot be separated into two parts. Motivated by the fact that singular kernels usually correspond to stronger norms than integrable kernels, e.g., the Riesz fractional gradients lead to Bessel potential spaces [46], it is a natural idea to choose an integrable and compactly supported kernel by which ww is bounded below, i.e., a kernel ϕ\phi satisfying eq. 1 and

(75) 0ϕ(𝒙)w(𝒙),suppϕB1(𝟎)¯ and  0<dϕ(𝒙)𝑑𝒙<,0\leq\phi(\bm{x})\leq w(\bm{x}),\;\text{supp}\ \phi\subset\overline{B_{1}(\bm{0})}\;\text{ and }\;0<\int_{\mathbb{R}^{d}}\phi(\bm{x})d\bm{x}<\infty,

and utilize the Poincaré inequality for integrable kernels with compact support. This further requires a comparison of the norms 𝔊w𝝂uL2(d;d)\|\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})} and 𝔊ϕ𝝂uL2(d;d)\|\mathfrak{G}_{\phi}^{\bm{\nu}}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})} which is not a trivial task. Here, we resort to the Fourier analysis. Let be 𝝀w𝝂{\bm{\lambda}}_{w}^{\bm{\nu}} and 𝝀ϕ𝝂{\bm{\lambda}}_{\phi}^{\bm{\nu}} be the Fourier symbols are defined by eq. 20. Notice that if there exists a constant C=C(w,d)>0C=C(w,d)>0 independent of 𝝂\bm{\nu} such that

(76) |𝝀w𝝂(𝝃)|C|𝝀ϕ𝝂(𝝃)|,𝝃d,|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})|\geq C|{\bm{\lambda}}_{\phi}^{\bm{\nu}}(\bm{\xi})|,\quad\forall\bm{\xi}\in\mathbb{R}^{d},

then we have for any uCc(Ω)u\in C^{\infty}_{c}(\Omega),

𝒢w𝝂uL2(d;d)2\displaystyle\|\mathcal{G}_{w}^{\bm{\nu}}u\|^{2}_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})} =d|𝝀w𝝂(𝝃)|2|u^(𝝃)|2𝑑𝝃\displaystyle=\int_{\mathbb{R}^{d}}\left|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})\right|^{2}|\hat{u}(\bm{\xi})|^{2}d\bm{\xi}
C2d|𝝀ϕ𝝂(𝝃)|2|u^(𝝃)|2𝑑𝝃=C2𝒢ϕ𝝂uL2(d;d)2,\displaystyle\geq C^{2}\int_{\mathbb{R}^{d}}\left|{\bm{\lambda}}_{\phi}^{\bm{\nu}}(\bm{\xi})\right|^{2}|\hat{u}(\bm{\xi})|^{2}d\bm{\xi}=C^{2}\|\mathcal{G}_{\phi}^{\bm{\nu}}u\|^{2}_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})},

and the Poincaré inequality for general kernels can be further inferred.

Lemma 5.1.

Assume that ww satisfies 5.1. Then there exists a kernel function ϕ\phi satisfying eq. 1 and eq. 75 such that eq. 76 holds, where 𝛌w𝛎{\bm{\lambda}}_{w}^{\bm{\nu}} and 𝛌ϕ𝛎{\bm{\lambda}}_{\phi}^{\bm{\nu}} are the Fourier symbols defined by eq. 20.

Proof.

We divide the proof of eq. 76 into two steps. Along the proof, the desired kernel function ϕ\phi will be constructed, and more precisely, is defined by eq. 85. Without loss of generality in the following steps we assume d2d\geq 2. The case for d=1d=1 is similar.

Step I. We prove that there exists N1=N1(w,d)(0,1)N_{1}=N_{1}(w,d)\in(0,1) and C1=C1(w,d)>0C_{1}=C_{1}(w,d)>0 such that

(77) |𝝀w𝝂(𝝃)|C1|𝝀ϕ𝝂(𝝃)|,|𝝃|<N1.|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})|\geq C_{1}|{\bm{\lambda}}_{\phi}^{\bm{\nu}}(\bm{\xi})|,\quad\forall|\bm{\xi}|<N_{1}.

Since ϕ\phi is integrable and satisfies eq. 75, there exists C>0C>0 depending on ww (as ϕ\phi itself depends on ww) such that for |𝝃|1|\bm{\xi}|\leq 1,

(78) |𝝀ϕ𝝂(𝝃)|22π|𝝃||𝒛|1ϕ(𝒛)𝑑𝒛=C|𝝃|.|{\bm{\lambda}}_{\phi}^{\bm{\nu}}(\bm{\xi})|\leq 2\sqrt{2}\pi|\bm{\xi}|\int_{|\bm{z}|\leq 1}\phi(\bm{z})d\bm{z}=C|\bm{\xi}|.

Observe that (𝝀w𝝂)(𝝃)\Im({\bm{\lambda}}_{w}^{\bm{\nu}})(\bm{\xi}) is a scalar multiple of 𝝃|𝝃|\frac{\bm{\xi}}{|\bm{\xi}|} as a result of 2.4, i.e., (𝝀w𝝂)(𝝃)=Λw(|𝝃|)𝝃|𝝃|\Im({\bm{\lambda}}_{w}^{\bm{\nu}})(\bm{\xi})=\Lambda_{w}(|\bm{\xi}|)\frac{\bm{\xi}}{|\bm{\xi}|} where Λw\Lambda_{w} is given by eq. 26. Using polar coordinates, we obtain

Λw(|𝝃|)\displaystyle\Lambda_{w}(|\bm{\xi}|) =120π0cos(θ)w¯(r)sin(2π|𝝃|rcos(θ))rd1sind2(θ)𝑑r𝑑θ\displaystyle=\frac{1}{2}\int_{0}^{\pi}\int_{0}^{\infty}\cos(\theta)\overline{w}(r)\sin(2\pi|\bm{\xi}|r\cos(\theta))r^{d-1}\sin^{d-2}(\theta)drd\theta
0πsind3(φ1)𝑑φ10πsind4(φ2)𝑑φ20πsin(φd3)𝑑φd302π𝑑φd2\displaystyle\quad\quad\int_{0}^{\pi}\sin^{d-3}(\varphi_{1})d\varphi_{1}\int_{0}^{\pi}\sin^{d-4}(\varphi_{2})d\varphi_{2}\cdots\int_{0}^{\pi}\sin(\varphi_{d-3})d\varphi_{d-3}\int_{0}^{2\pi}d\varphi_{d-2}
=12ωd20π0cos(θ)w¯(r)sin(2π|𝝃|rcos(θ))rd1sind2(θ)𝑑r𝑑θ\displaystyle=\frac{1}{2}\omega_{d-2}\int_{0}^{\pi}\int_{0}^{\infty}\cos(\theta)\overline{w}(r)\sin(2\pi|\bm{\xi}|r\cos(\theta))r^{d-1}\sin^{d-2}(\theta)drd\theta
=ωd20π20cos(θ)sind2(θ)rd1w¯(r)sin(2π|𝝃|rcos(θ))𝑑r𝑑θ,\displaystyle=\omega_{d-2}\int_{0}^{\frac{\pi}{2}}\int_{0}^{\infty}\cos(\theta)\sin^{d-2}(\theta)r^{d-1}\overline{w}(r)\sin(2\pi|\bm{\xi}|r\cos(\theta))drd\theta,

where ωd1=2πd/2Γ(d/2)\omega_{d-1}=\frac{2\pi^{d/2}}{\Gamma(d/2)} is the surface area of (d1)(d-1)-sphere 𝕊d1\mathbb{S}^{d-1}. Now we claim that for ww in 5.1,

Λw(|𝝃|)C(w,d)|𝝃|,|𝝃|<N1.\Lambda_{w}(|\bm{\xi}|)\geq C(w,d)|\bm{\xi}|,\quad\forall|\bm{\xi}|<N_{1}.

Then eq. 77 holds by eq. 78 and |𝝀w𝝂(𝝃)||(𝝀w𝝂)(𝝃)|Λw(|𝝃|)|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})|\geq|\Im({\bm{\lambda}}_{w}^{\bm{\nu}})(\bm{\xi})|\geq\Lambda_{w}(|\bm{\xi}|). We prove the claim by two cases to conclude Step I.

Case (i). Suppose ww satisfies 5.1 (1). Then 0rdw¯(r)𝑑r<\int_{0}^{\infty}r^{d}\overline{w}(r)dr<\infty. Since g(r,θ):=cos2(θ)sind2(θ)rdw¯(r)L1((0,)×(0,π2))g(r,\theta):=\cos^{2}(\theta)\sin^{d-2}(\theta)r^{d}\overline{w}(r)\in L^{1}\left((0,\infty)\times(0,\frac{\pi}{2})\right), by dominated convergence theorem,

Λw(|𝝃|)|𝝃|\displaystyle\frac{\Lambda_{w}(|\bm{\xi}|)}{|\bm{\xi}|} =2πωd20π20cos2(θ)sind2(θ)rdw¯(r)sin(2π|𝝃|rcos(θ))2π|𝝃|rcos(θ)𝑑r𝑑θ\displaystyle=2\pi\omega_{d-2}\int_{0}^{\frac{\pi}{2}}\int_{0}^{\infty}\cos^{2}(\theta)\sin^{d-2}(\theta)r^{d}\overline{w}(r)\frac{\sin(2\pi|\bm{\xi}|r\cos(\theta))}{2\pi|\bm{\xi}|r\cos(\theta)}drd\theta
2πωd20π2cos2(θ)sind2(θ)𝑑θ0rdw¯(r)𝑑r>0,as |𝝃|0.\displaystyle\to 2\pi\omega_{d-2}\int_{0}^{\frac{\pi}{2}}\cos^{2}(\theta)\sin^{d-2}(\theta)d\theta\int_{0}^{\infty}r^{d}\overline{w}(r)dr>0,\quad\text{as }|\bm{\xi}|\to 0.

Therefore, there exists N1=N1(w,d)(0,1)N_{1}=N_{1}(w,d)\in(0,1) such that the claim holds.

Case (ii). Suppose ww satisfies 5.1 (2). Assume without loss of generality that R=1R=1. Then 01rdw¯(r)𝑑r<\int_{0}^{1}r^{d}\overline{w}(r)dr<\infty and w¯(r)=c0rd+α\overline{w}(r)=\frac{c_{0}}{r^{d+\alpha}} with α(0,1]\alpha\in(0,1] for r>1r>1. We estimate Λw(|𝝃|)\Lambda_{w}(|\bm{\xi}|) by discussing r1r\leq 1 and r>1r>1. On the one hand, since sinx12x\sin x\geq\frac{1}{2}x for |x||x| sufficiently small, there exists N1(0,1)N_{1}\in(0,1) such that 01rd1w¯(r)sin(2π|𝝃|rcos(θ))𝑑rπ|𝝃|cos(θ)01rdw¯(r)𝑑r\int_{0}^{1}r^{d-1}\overline{w}(r)\sin(2\pi|\bm{\xi}|r\cos(\theta))dr\geq\pi|\bm{\xi}|\cos(\theta)\int_{0}^{1}r^{d}\overline{w}(r)dr for |𝝃|<N1|\bm{\xi}|<N_{1}. On the other hand, there exist C=C(α,c0)C=C(\alpha,c_{0}) and N1=N1(α)(0,1)N_{1}=N_{1}(\alpha)\in(0,1) such that

1rd1c0rd+αsin(2π|𝝃|rcos(θ))𝑑r=c0|𝝃|α(cosθ)α|𝝃|cosθ1r1+αsin(2πr)𝑑r\displaystyle\quad\int_{1}^{\infty}r^{d-1}\frac{c_{0}}{r^{d+\alpha}}\sin(2\pi|\bm{\xi}|r\cos(\theta))dr=c_{0}|\bm{\xi}|^{\alpha}(\cos\theta)^{\alpha}\int_{|\bm{\xi}|\cos\theta}^{\infty}\frac{1}{r^{1+\alpha}}\sin(2\pi r)dr
C|𝝃|α(cosθ)αC|𝝃|cos(θ),|𝝃|<N1,θ(0,π2),\displaystyle\geq C|\bm{\xi}|^{\alpha}(\cos\theta)^{\alpha}\geq C|\bm{\xi}|\cos(\theta),\quad\forall|\bm{\xi}|<N_{1},\ \theta\in\left(0,\frac{\pi}{2}\right),

where we used 01r1+αsin(2πr)𝑑r(0,)\int_{0}^{\infty}\frac{1}{r^{1+\alpha}}\sin(2\pi r)dr\in(0,\infty) and dominated convergence theorem for the second last inequality. Combining both cases for r1r\leq 1 and r>1r>1 yields the claim.

Step II. We prove that there exist N2=N2(w,d)>1N_{2}=N_{2}(w,d)>1 and C2=C2(w,d)>0C_{2}=C_{2}(w,d)>0 such that

(79) |𝝀w𝝂(𝝃)|C2|𝝀ϕ𝝂(𝝃)|,|𝝃|>N2,|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})|\geq C_{2}|{\bm{\lambda}}_{\phi}^{\bm{\nu}}(\bm{\xi})|,\quad\forall|\bm{\xi}|>N_{2},

and C3=C3(w,d)>0C_{3}=C_{3}(w,d)>0 such that

(80) |𝝀w𝝂(𝝃)|C3|𝝀ϕ𝝂(𝝃)|,|𝝃|[N1,N2].|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})|\geq C_{3}|{\bm{\lambda}}_{\phi}^{\bm{\nu}}(\bm{\xi})|,\quad\forall|\bm{\xi}|\in[N_{1},N_{2}].

Since ϕ\phi is integrable,

(81) |𝝀ϕ𝝂(𝝃)|2|𝒛|1ϕ(𝒛)𝑑𝒛=2Iϕ,𝝃d,|{\bm{\lambda}}_{\phi}^{\bm{\nu}}(\bm{\xi})|\leq 2\int_{|\bm{z}|\leq 1}\phi(\bm{z})d\bm{z}=2I_{\phi},\quad\bm{\xi}\in\mathbb{R}^{d},

where Iϕ:=dϕ(𝒛)𝑑𝒛(0,)I_{\phi}:=\int_{\mathbb{R}^{d}}\phi(\bm{z})d\bm{z}\in(0,\infty) depends on ww. Denote 𝝃^:=𝝃|𝝃|\hat{\bm{\xi}}:=\frac{\bm{\xi}}{|\bm{\xi}|}. Recall that 𝝂=R𝝂𝒆1\bm{\nu}=R_{\bm{\nu}}\bm{e}_{1}. Then

(82) |𝝀w𝝂(𝝃)|\displaystyle|{\bm{\lambda}}_{w}^{\bm{\nu}}(\bm{\xi})| |(𝝀w𝝂)(𝝃)||𝝂(𝝀w𝝂)(𝝃)|\displaystyle\geq|\Re({\bm{\lambda}}_{w}^{\bm{\nu}})(\bm{\xi})|\geq|\bm{\nu}\cdot\Re({\bm{\lambda}}_{w}^{\bm{\nu}})(\bm{\xi})|
={𝒛𝝂>0}𝒛𝝂|𝒛|w(𝒛)(1cos(2π𝝃𝒛))𝑑𝒛\displaystyle=\int_{\{\bm{z}\cdot\bm{\nu}>0\}}\frac{\bm{z}\cdot\bm{\nu}}{|\bm{z}|}w(\bm{z})(1-\cos(2\pi\bm{\xi}\cdot\bm{z}))d\bm{z}
={z1>0}z1|𝒛|w(𝒛)(1cos(2π|𝝃|(R𝝂T𝝃^)𝒛))𝑑𝒛.\displaystyle=\int_{\{z_{1}>0\}}\frac{z_{1}}{|\bm{z}|}w(\bm{z})(1-\cos(2\pi|\bm{\xi}|(R_{\bm{\nu}}^{T}\hat{\bm{\xi}})\cdot\bm{z}))d\bm{z}.

For any fL1(d)f\in L^{1}(\mathbb{R}^{d}), we define a function I:+×𝕊d1I:\mathbb{R}_{+}\times\mathbb{S}^{d-1}\to\mathbb{R} by

I(ρ,𝜼):=df(z)(1cos(2πρ𝜼𝒛))𝑑𝒛.I(\rho,\bm{\eta}):=\int_{\mathbb{R}^{d}}f(z)(1-\cos(2\pi\rho\bm{\eta}\cdot\bm{z}))d\bm{z}.

Claim. For ww in 5.1, there exists fL1(d)f\in L^{1}(\mathbb{R}^{d}) depending only on ww such that

(83) 0f(𝒛)χ𝒆1(𝒛)z1|𝒛|w(𝒛)0\leq f(\bm{z})\leq\chi_{\bm{e}_{1}}(\bm{z})\frac{z_{1}}{|\bm{z}|}w(\bm{z})

and

(84) I(|𝝃|,𝜼)C(d)Iϕ,|𝝃|>N2,𝜼𝕊d1.I(|\bm{\xi}|,\bm{\eta})\geq C(d)I_{\phi},\quad\forall|\bm{\xi}|>N_{2},\ \bm{\eta}\in\mathbb{S}^{d-1}.

Once the claim is proved, eq. 79 and eq. 80 follows. Indeed, eq. 79 holds by eq. 81, eq. 82, eq. 83 and eq. 84. Notice that II is a continuous function and I(|𝝃|,𝜼)>0I(|\bm{\xi}|,\bm{\eta})>0 for any |𝝃|[N1,N2]|\bm{\xi}|\in[N_{1},N_{2}] and 𝜼𝕊d1\bm{\eta}\in\mathbb{S}^{d-1}, eq. 80 holds as min[N1,N2]×𝕊d1I(ρ,𝜼)>0\min_{[N_{1},N_{2}]\times\mathbb{S}^{d-1}}I(\rho,\bm{\eta})>0 and Iϕ>0I_{\phi}>0. We prove the claim to conclude Step II and thus finish the whole proof.

Proof of the claim. We choose ϕ\phi as

(85) ϕ(𝒙)=min(1,w(𝒙)χB1(𝟎)(𝒙)).\phi(\bm{x})=\min(1,w(\bm{x})\chi_{B_{1}(\bm{0})}(\bm{x})).

Then ϕ\phi satisfies eq. 1 and eq. 75. Let f(𝒛):=χ𝒆1(𝒛)z1|𝒛|ϕ(𝒛)f(\bm{z}):=\chi_{\bm{e}_{1}}(\bm{z})\frac{z_{1}}{|\bm{z}|}\phi(\bm{z}), then fL1(d)f\in L^{1}(\mathbb{R}^{d}) satisfies eq. 83 and for V:=(0,)×(0,π/2)×(0,π)d3×(0,2π)V:=(0,\infty)\times(0,\pi/2)\times(0,\pi)^{d-3}\times(0,2\pi),

df(𝒛)𝑑𝒛={z1>0}z1|𝒛|ϕ(𝒛)𝑑𝒛\displaystyle\quad\ \int_{\mathbb{R}^{d}}f(\bm{z})d\bm{z}=\int_{\{z_{1}>0\}}\frac{z_{1}}{|\bm{z}|}\phi(\bm{z})d\bm{z}
=∫⋯∫Vcos(φ1)ϕ¯(r)rd1sind2(φ1)sind3(φ2)sin(φd2)𝑑r𝑑φ1𝑑φ2𝑑φd2𝑑θ\displaystyle=\idotsint_{V}\cos(\varphi_{1})\bar{\phi}(r)r^{d-1}\sin^{d-2}(\varphi_{1})\sin^{d-3}(\varphi_{2})\cdots\sin(\varphi_{d-2})drd\varphi_{1}d\varphi_{2}\cdots d\varphi_{d-2}d\theta
=ωd20π2cos(φ1)sind2(φ1)𝑑φ10rd1ϕ¯(r)𝑑r\displaystyle=\omega_{d-2}\int_{0}^{\frac{\pi}{2}}\cos(\varphi_{1})\sin^{d-2}(\varphi_{1})d\varphi_{1}\int_{0}^{\infty}r^{d-1}\bar{\phi}(r)dr
=ωd2(d1)ωd1dϕ(𝒛)𝑑𝒛,\displaystyle=\frac{\omega_{d-2}}{(d-1)\omega_{d-1}}\int_{\mathbb{R}^{d}}\phi(\bm{z})d\bm{z},

where ϕ¯\bar{\phi} is the radial representation of ϕ\phi. Note that the above computation holds for d3d\geq 3 and can be easily done for d=1d=1 or 22. Then by Riemann-Lebesgue lemma, there exists N2=N2(w,d)>1N_{2}=N_{2}(w,d)>1 such that |𝝃|>N2\forall|\bm{\xi}|>N_{2},

df(𝒛)cos(2π|𝝃|𝜼𝒛)𝑑𝒛<12df(𝒛)𝑑𝒛,𝜼𝕊d1.\int_{\mathbb{R}^{d}}f(\bm{z})\cos(2\pi|\bm{\xi}|\bm{\eta}\cdot\bm{z})d\bm{z}<\frac{1}{2}\int_{\mathbb{R}^{d}}f(\bm{z})d\bm{z},\quad\forall\bm{\eta}\in\mathbb{S}^{d-1}.

Then for |𝝃|>N2|\bm{\xi}|>N_{2} and 𝜼𝕊d1\bm{\eta}\in\mathbb{S}^{d-1},

I(|𝝃|,𝜼)=df(𝒛)(1cos(2π|𝝃|𝜼𝒛))𝑑𝒛12df(𝒛)𝑑𝒛C(d)Iϕ,I(|\bm{\xi}|,\bm{\eta})=\int_{\mathbb{R}^{d}}f(\bm{z})(1-\cos(2\pi|\bm{\xi}|\bm{\eta}\cdot\bm{z}))d\bm{z}\geq\frac{1}{2}\int_{\mathbb{R}^{d}}f(\bm{z})d\bm{z}\geq C(d)I_{\phi},

where C(d)=ωd22(d1)ωd1C(d)=\frac{\omega_{d-2}}{2(d-1)\omega_{d-1}}. Then the claim is proved. ∎

Now we are ready to prove the the Poincaré inequality for general kernels.

Proof of 5.1.

For any uCc(Ω)u\in C_{c}^{\infty}(\Omega), we have 𝔊w𝝂u=𝒢w𝝂u\mathfrak{G}^{\bm{\nu}}_{w}u=\mathcal{G}_{w}^{\bm{\nu}}u. By 5.1 and the comment below eq. 76, there exists a kernel function ϕ\phi satisfying eq. 1 and eq. 75 such that

𝔊w𝝂uL2(d;d)=𝒢w𝝂uL2(d;d)C𝒢ϕ𝝂uL2(d;d)\|\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}=\|\mathcal{G}_{w}^{\bm{\nu}}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\geq C\|\mathcal{G}_{\phi}^{\bm{\nu}}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}

for some C=C(w,d)>0C=C(w,d)>0. Therefore, using 4.1 for the integrable and compactly supported kernel ϕ\phi, we obtain

uL2(Ω)Π(ϕ,𝝂,Ω)𝒢ϕ𝝂uL2(d;d)C1Π(ϕ,𝝂,Ω)𝔊w𝝂uL2(d;d),uCc(Ω).\|u\|_{L^{2}(\Omega)}\leq\Pi(\phi,\bm{\nu},\Omega)\|\mathcal{G}_{\phi}^{\bm{\nu}}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\leq C^{-1}\Pi(\phi,\bm{\nu},\Omega)\|\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})},\;\forall u\in C_{c}^{\infty}(\Omega).

Denote Π(w,𝝂,Ω):=C1Π(ϕ,𝝂,Ω)\Pi(w,\bm{\nu},\Omega):=C^{-1}\Pi(\phi,\bm{\nu},\Omega). By the density result in 3.2, for every u𝒮w𝝂(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega), there exists {uj}j=1Cc(Ω)\{u_{j}\}_{j=1}^{\infty}\subset C_{c}^{\infty}(\Omega) such that ujuu_{j}\to u in 𝒮w𝝂(Ω)\mathcal{S}_{w}^{\bm{\nu}}(\Omega). Hence,

ujuL2(Ω)0, and 𝔊w𝝂uj𝔊w𝝂uLp(d;d)0,j.\|u_{j}-u\|_{L^{2}(\Omega)}\to 0,\ \text{ and }\ \|\mathfrak{G}^{\bm{\nu}}_{w}u_{j}-\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{p}(\mathbb{R}^{d};\mathbb{R}^{d})}\to 0,\quad j\to\infty.

Since

ujL2(Ω)Π(w,𝝂,Ω)𝔊w𝝂ujL2(d;d),\|u_{j}\|_{L^{2}(\Omega)}\leq\Pi(w,\bm{\nu},\Omega)\|\mathfrak{G}^{\bm{\nu}}_{w}u_{j}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})},

letting jj\to\infty yields (74). ∎

6. Applications

In this section, we provide some applications of the nonlocal Poincaré inequality. Assume that ww is a kernel function satisfying 5.1, 𝝂d\bm{\nu}\in\mathbb{R}^{d} is a fixed unit vector, Ωd\Omega\subset\mathbb{R}^{d} is an open bounded domain with a continuous boundary. Note that by the nonlocal Poincaré inequality 4.1 and 5.1, the full norm u𝒮w𝝂(Ω)\|u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)} is equivalent to the seminorm 𝔊w𝝂uL2(d;d)\|\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})} for u𝒮w𝝂(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega). Thus in this section we abuse the notation and use 𝒮w𝝂(Ω)\|\cdot\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)} to denote the seminorm.

6.1. Nonlocal convection-diffusion equation

In Section 2 we have defined nonlocal gradient and divergence operator for the fixed unit vector 𝝂\bm{\nu}. It turns out that one can define these notions corresponding to a unit vector field 𝒏=𝒏(𝒙)\bm{n}=\bm{n}(\bm{x}) as well. Specifically, for a measurable function u:du:\mathbb{R}^{d}\to\mathbb{R} and a measurable vector field 𝒗:dd\bm{v}:\mathbb{R}^{d}\to\mathbb{R}^{d}, 𝒢w𝒏u:Ω\mathcal{G}^{\bm{n}}_{w}u:\Omega\to\mathbb{R} and 𝒟w𝒏𝒗:Ωd\mathcal{D}^{\bm{n}}_{w}\bm{v}:\Omega\to\mathbb{R}^{d} are defined by

𝒢w𝒏u(𝒙):=limϵ0d\Bϵ(𝒙)χ𝒏(𝒙)(𝒚𝒙)(u(𝒚)u(𝒙))𝒚𝒙|𝒚𝒙|w(𝒚𝒙)𝑑𝒚\mathcal{G}^{\bm{n}}_{w}u(\bm{x}):=\lim_{\epsilon\to 0}\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\chi_{\bm{n}(\bm{x})}(\bm{y}-\bm{x})(u(\bm{y})-u(\bm{x}))\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}w(\bm{y}-\bm{x})d\bm{y}

and

𝒟w𝒏𝒗(𝒙):=limϵ0d\Bϵ(𝒙)𝒚𝒙|𝒚𝒙|(χ𝒏(𝒚)(𝒚𝒙)𝒗(𝒚)+χ𝒏(𝒙)(𝒙𝒚)𝒗(𝒙))w(𝒚𝒙)𝑑𝒚,\mathcal{D}^{\bm{n}}_{w}\bm{v}(\bm{x}):=\lim_{\epsilon\to 0}\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{n}(\bm{y})}(\bm{y}-\bm{x})\bm{v}(\bm{y})+\chi_{\bm{n}(\bm{x})}(\bm{x}-\bm{y})\bm{v}(\bm{x}))w(\bm{y}-\bm{x})d\bm{y},

respectively. Let ϕ\phi be an integrable kernel with compact support satisfying eq. 1. Then the integration by parts formula in 4.1 (2) holds for the vector field 𝒏\bm{n} and kernel ϕ\phi. The proof is similar and thus omitted.

For a diffusivity function ϵ=ϵ(𝒙)L(d)\epsilon=\epsilon(\bm{x})\in L^{\infty}(\mathbb{R}^{d}) with a positive lower bound ϵ1>0\epsilon_{1}>0, a vector field 𝒃L(d;d)\bm{b}\in L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d}) and a function f(𝒮w𝝂(Ω))f\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega))^{\ast}, we consider the nonlocal convection-diffusion model problem formulated as

(86) {𝔇w𝝂(ϵ𝔊w𝝂u)+𝒃𝒢ϕ𝒏u=fin Ω,u=0in d\Ω.\left\{\begin{aligned} -\mathfrak{D}^{-\bm{\nu}}_{w}(\epsilon\mathfrak{G}^{\bm{\nu}}_{w}u)+\bm{b}\cdot\mathcal{G}_{\phi}^{\bm{n}}u&=f\quad\text{in }\ \Omega,\\ u&=0\quad\text{in }\ \mathbb{R}^{d}\backslash\Omega.\end{aligned}\right.

Equation 86 is a nonlocal analogue of the classical convection-diffusion equation, see, e.g., [14, 39, 53, 52] for related discussions. The new formulation using 𝔇w𝝂(ϵ𝔊w𝝂u)\mathfrak{D}^{-\bm{\nu}}_{w}(\epsilon\mathfrak{G}^{\bm{\nu}}_{w}u) for the nonlocal diffusion allows the possibility to explore mixed-type numerical methods for eq. 86 in the future.

Remark 6.1.

If the kernel function ww has compact support, the boundary condition in eq. 86 only needs to be imposed on a bounded domain outside Ω\Omega. For example, assume suppwBδ(𝟎)\text{supp}\ w\subset B_{\delta}(\bm{0}) for δ>0\delta>0, then for the first equation to be well-defined on Ω\Omega, we only need u=0u=0 on Ω2δ\Ω\Omega_{2\delta}\backslash\Omega where Ω2δ={𝐱d:dist(𝐱,Ω)<2δ}\Omega_{2\delta}=\{\bm{x}\in\mathbb{R}^{d}:\text{dist}(\bm{x},\Omega)<2\delta\}.

We define the bilinear form b(,):𝒮w𝝂(Ω)×𝒮w𝝂(Ω)b(\cdot,\cdot):\mathcal{S}_{w}^{\bm{\nu}}(\Omega)\times\mathcal{S}_{w}^{\bm{\nu}}(\Omega)\to\mathbb{R} associated with eq. 86 by

(87) b(u,v):=(ϵ𝔊w𝝂u,𝔊w𝝂v)L2(d;d)+(𝒃𝒢ϕ𝒏u,v)L2(d).b(u,v):=(\epsilon\mathfrak{G}^{\bm{\nu}}_{w}u,\mathfrak{G}^{\bm{\nu}}_{w}v)_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}+(\bm{b}\cdot\mathcal{G}_{\phi}^{\bm{n}}u,v)_{L^{2}(\mathbb{R}^{d})}.

Then the weak formulation is given as follows.

(88) {Find u𝒮w𝝂(Ω) such that: b(u,v)=f,v,v𝒮w𝝂(Ω).\left\{\begin{aligned} &\text{Find }u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega)\text{ such that: }\\ &b(u,v)=\langle f,v\rangle,\quad\forall v\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega).\end{aligned}\right.

The vector field 𝒏\bm{n} is given in terms of 𝒃\bm{b} by the following relation

(89) 𝒏=𝒃|𝒃|.\bm{n}=-\frac{\bm{b}}{|\bm{b}|}.

To establish the well-posedness of the model problem (86), we give an additional assumption on the velocity field 𝒃\bm{b}.

Assumption 6.1.

Assume the velocity field 𝐛L(d;d)\bm{b}\in L^{\infty}(\mathbb{R}^{d};\mathbb{R}^{d}) satisifies either one of the following assumptions:

  1. (i)

    𝒟ϕ𝒏𝒃0\mathcal{D}_{\phi}^{-\bm{n}}\ \bm{b}\leq 0, or

  2. (ii)

    |𝒟ϕ𝒏𝒃|η|\mathcal{D}_{\phi}^{-\bm{n}}\ \bm{b}|\leq\eta where η<2ϵ1/Π2\eta<2\epsilon_{1}/\Pi^{2}.

We further present a result on the convection part of the bilinear form b(u,v)b(u,v). Similar result can be found in [39].

Lemma 6.1.

Let v𝒮w𝛎(Ω)v\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega) and 𝐧\bm{n} be defined as eq. 89. Then

(𝒃𝒢ϕ𝒏v,v)L2(Ω)12(v2,𝒟ϕ𝒏𝒃)L2(Ω).(\bm{b}\cdot\mathcal{G}_{\phi}^{\bm{n}}v,v)_{L^{2}(\Omega)}\geq-\frac{1}{2}(v^{2},\mathcal{D}_{\phi}^{-\bm{n}}\bm{b})_{L^{2}(\Omega)}.
Proof.

By the integration by parts formula for vector field 𝒏\bm{n},

(𝒃𝒢ϕ𝒏v,v)L2(Ω)=(𝒃𝒢ϕ𝒏v,v)L2(d)=(v,𝒟ϕ𝒏(𝒃v))L2(d)\displaystyle\quad\ (\bm{b}\cdot\mathcal{G}_{\phi}^{\bm{n}}v,v)_{L^{2}(\Omega)}=(\bm{b}\cdot\mathcal{G}_{\phi}^{\bm{n}}v,v)_{L^{2}(\mathbb{R}^{d})}=-(v,\mathcal{D}_{\phi}^{-\bm{n}}(\bm{b}v))_{L^{2}(\mathbb{R}^{d})}
=dv(𝒙)d𝒚𝒙|𝒚𝒙|(χ𝒏(𝒚)(𝒙𝒚)𝒃(𝒚)v(𝒚)+\displaystyle=-\int_{\mathbb{R}^{d}}v(\bm{x})\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{n}(\bm{y})}(\bm{x}-\bm{y})\bm{b}(\bm{y})v(\bm{y})+
χ𝒏(𝒙)(𝒚𝒙)𝒃(𝒙)v(𝒙))ϕ(𝒚𝒙)d𝒚d𝒙\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\chi_{\bm{n}(\bm{x})}(\bm{y}-\bm{x})\bm{b}(\bm{x})v(\bm{x}))\phi(\bm{y}-\bm{x})d\bm{y}d\bm{x}
=122d(v(𝒚)v(𝒙))𝒚𝒙|𝒚𝒙|(χ𝒏(𝒚)(𝒙𝒚)𝒃(𝒚)v(𝒚)+\displaystyle=\frac{1}{2}\iint_{\mathbb{R}^{2d}}(v(\bm{y})-v(\bm{x}))\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{n}(\bm{y})}(\bm{x}-\bm{y})\bm{b}(\bm{y})v(\bm{y})+
χ𝒏(𝒙)(𝒚𝒙)𝒃(𝒙)v(𝒙))ϕ(𝒚𝒙)d𝒚d𝒙\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\chi_{\bm{n}(\bm{x})}(\bm{y}-\bm{x})\bm{b}(\bm{x})v(\bm{x}))\phi(\bm{y}-\bm{x})d\bm{y}d\bm{x}
=122d(v(𝒚)v(𝒙))v(𝒙)𝒚𝒙|𝒚𝒙|(χ𝒏(𝒚)(𝒙𝒚)𝒃(𝒚)+\displaystyle=\frac{1}{2}\iint_{\mathbb{R}^{2d}}(v(\bm{y})-v(\bm{x}))v(\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{n}(\bm{y})}(\bm{x}-\bm{y})\bm{b}(\bm{y})+
χ𝒏(𝒙)(𝒚𝒙)𝒃(𝒙))ϕ(𝒚𝒙)d𝒚d𝒙\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\chi_{\bm{n}(\bm{x})}(\bm{y}-\bm{x})\bm{b}(\bm{x}))\phi(\bm{y}-\bm{x})d\bm{y}d\bm{x}
+12dd(v(𝒚)v(𝒙))2𝒚𝒙|𝒚𝒙|χ𝒏(𝒚)(𝒙𝒚)𝒃(𝒚)ϕ(𝒚𝒙)𝑑𝒚𝑑𝒙\displaystyle\quad\quad+\frac{1}{2}\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}(v(\bm{y})-v(\bm{x}))^{2}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot\chi_{\bm{n}(\bm{y})}(\bm{x}-\bm{y})\bm{b}(\bm{y})\phi(\bm{y}-\bm{x})d\bm{y}d\bm{x}
=12dv(𝒙)2d𝒚𝒙|𝒚𝒙|(χ𝒏(𝒚)(𝒙𝒚)𝒃(𝒚)+χ𝒏(𝒙)(𝒚𝒙)𝒃(𝒙))ϕ(𝒚𝒙)𝑑𝒚𝑑𝒙\displaystyle=-\frac{1}{2}\int_{\mathbb{R}^{d}}v(\bm{x})^{2}\int_{\mathbb{R}^{d}}\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\cdot(\chi_{\bm{n}(\bm{y})}(\bm{x}-\bm{y})\bm{b}(\bm{y})+\chi_{\bm{n}(\bm{x})}(\bm{y}-\bm{x})\bm{b}(\bm{x}))\phi(\bm{y}-\bm{x})d\bm{y}d\bm{x}
+12dd(v(𝒚)v(𝒙))2𝒙𝒚|𝒚𝒙|χ𝒏(𝒙)(𝒚𝒙)𝒃(𝒙)ϕ(𝒚𝒙)𝑑𝒚𝑑𝒙\displaystyle\quad\quad+\frac{1}{2}\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}(v(\bm{y})-v(\bm{x}))^{2}\frac{\bm{x}-\bm{y}}{|\bm{y}-\bm{x}|}\cdot\chi_{\bm{n}(\bm{x})}(\bm{y}-\bm{x})\bm{b}(\bm{x})\phi(\bm{y}-\bm{x})d\bm{y}d\bm{x}
=12Ωv(𝒙)2𝒟ϕ𝒏𝒃(𝒙)𝑑𝒙\displaystyle=-\frac{1}{2}\int_{\Omega}v(\bm{x})^{2}\mathcal{D}_{\phi}^{-\bm{n}}\ \bm{b}(\bm{x})d\bm{x}
+12d{(𝒚𝒙)𝒏(𝒙)>0}(v(𝒚)v(𝒙))2𝒙𝒚|𝒚𝒙|𝒃(𝒙)ϕ(𝒚𝒙)𝑑𝒚𝑑𝒙\displaystyle\quad\quad+\frac{1}{2}\int_{\mathbb{R}^{d}}\int_{\{(\bm{y}-\bm{x})\cdot\bm{n}(\bm{x})>0\}}(v(\bm{y})-v(\bm{x}))^{2}\frac{\bm{x}-\bm{y}}{|\bm{y}-\bm{x}|}\cdot\bm{b}(\bm{x})\phi(\bm{y}-\bm{x})d\bm{y}d\bm{x}
12Ωv(𝒙)2𝒟ϕ𝒏𝒃(𝒙)𝑑𝒙.\displaystyle\geq-\frac{1}{2}\int_{\Omega}v(\bm{x})^{2}\mathcal{D}_{\phi}^{-\bm{n}}\ \bm{b}(\bm{x})d\bm{x}.

where we used eq. 89 in the last inequality. ∎

Use the above lemma, we can establish the coercivity of the bilinear form and the well-posedness of eq. 86 further by the Lax-Milgram theorem.

Theorem 6.1.

Assume that 6.1 is satisfied. The nonlocal convection-diffusion problem (88) is well-posed. More precisely, for any f(𝒮w𝛎(Ω))f\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega))^{\ast}, there exists a unique solution u𝒮w𝛎(Ω)u\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega) such that

u𝒮w𝝂(Ω)cf(𝒮w𝝂(Ω)),\|u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)}\leq c\|f\|_{(\mathcal{S}_{w}^{\bm{\nu}}(\Omega))^{\ast}},

where c=c(ϵ,𝐛,w,ϕ,𝛎,Ω)c=c(\epsilon,\bm{b},w,\phi,\bm{\nu},\Omega) is a positive constant.

Proof.

Notice that b(,)b(\cdot,\cdot) is coercive under 6.1. Indeed, if 6.1 (i) is satisfied, then by 6.1, b(v,v)ϵ1a(v,v)=ϵ1v𝒮w𝝂(Ω)2b(v,v)\geq\epsilon_{1}a(v,v)=\epsilon_{1}\|v\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)}^{2}, v𝒮w𝝂(Ω)\forall v\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega). On the other hand, if 6.1 (ii) is satisfied, then by the nonlocal Poincaré inequality,

12|(v2,𝒟ϕ𝒏𝒃)L2(Ω)|12ηvL2(Ω)212ηΠ2𝔊w𝝂vL2(Ω;d)2,\frac{1}{2}|(v^{2},\mathcal{D}_{\phi}^{-\bm{n}}\bm{b})_{L^{2}(\Omega)}|\leq\frac{1}{2}\eta\|v\|_{L^{2}(\Omega)}^{2}\leq\frac{1}{2}\eta\Pi^{2}\|\mathfrak{G}_{w}^{\bm{\nu}}v\|_{L^{2}(\Omega;\mathbb{R}^{d})}^{2},

where Π=Π(w,𝝂,Ω)\Pi=\Pi(w,\bm{\nu},\Omega) is Poincaré constant. Hence, by 6.1,

(𝒃𝒢ϕ𝒏v,v)L2(Ω)12|(v2,𝒟ϕ𝒏𝒃)L2(Ω)|12ηΠ2𝔊w𝝂vL2(Ω;d)2.(\bm{b}\cdot\mathcal{G}_{\phi}^{\bm{n}}v,v)_{L^{2}(\Omega)}\geq-\frac{1}{2}|(v^{2},\mathcal{D}_{\phi}^{-\bm{n}}\bm{b})_{L^{2}(\Omega)}|\geq-\frac{1}{2}\eta\Pi^{2}\|\mathfrak{G}_{w}^{\bm{\nu}}v\|_{L^{2}(\Omega;\mathbb{R}^{d})}^{2}.

Therefore if 6.1 (ii) is satisfied, we have b(v,v)(ϵ112ηΠ2)v𝒮w𝝂(Ω)2b(v,v)\geq(\epsilon_{1}-\frac{1}{2}\eta\Pi^{2})\|v\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega)}^{2}, v𝒮w𝝂(Ω)\forall v\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega).

The boundedness of b(,)b(\cdot,\cdot) follows from the nonlocal Poincaré inequality and the estimate

𝒢ϕ𝒏vL2(d;d)2ϕL1(d)vL2(d),vL2(d),\|\mathcal{G}^{\bm{n}}_{\phi}v\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\leq 2\|\phi\|_{L^{1}(\mathbb{R}^{d})}\|v\|_{L^{2}(\mathbb{R}^{d})},\quad\forall v\in L^{2}(\mathbb{R}^{d}),

which is an analog of 4.1 (1) for 𝒏\bm{n} when p=2p=2. Finally, the Lax-Milgram theorem yields the well-posedness of eq. 88. ∎

6.2. Nonlocal correspondence models of isotropic linear elasticity

For 𝒖𝒮w𝝂(Ω;N)\bm{u}\in\mathcal{S}^{\bm{\nu}}_{w}(\Omega;\mathbb{R}^{N}), define distributional nonlocal vector Laplacian

(90) w𝝂𝒖:=𝔇w𝝂𝔊w𝝂𝒖.\mathcal{L}_{w}^{\bm{\nu}}\bm{u}:=\mathfrak{D}_{w}^{-\bm{\nu}}\mathfrak{G}_{w}^{\bm{\nu}}\bm{u}.

Then w𝝂𝒖(𝒮w𝝂(Ω;N))\mathcal{L}_{w}^{\bm{\nu}}\bm{u}\in(\mathcal{S}^{\bm{\nu}}_{w}(\Omega;\mathbb{R}^{N}))^{\ast} and w𝝂:𝒮w𝝂(Ω;N)(𝒮w𝝂(Ω;N))\mathcal{L}_{w}^{\bm{\nu}}:\mathcal{S}^{\bm{\nu}}_{w}(\Omega;\mathbb{R}^{N})\to(\mathcal{S}^{\bm{\nu}}_{w}(\Omega;\mathbb{R}^{N}))^{\ast} is a bounded linear operator with operator norm no more than 11 by 3.4. For the rest of the paper, we consider N=dN=d.

For a displacement field 𝒖:dd\bm{u}:\mathbb{R}^{d}\to\mathbb{R}^{d}, we study the elastic potential energy given by

(91) (𝒖)=12λ𝔇w𝝂(𝒖)L2(d)2+μew𝝂(𝒖)L2(d;d×d)2,\mathcal{E}(\bm{u})=\frac{1}{2}\lambda\|\mathfrak{D}_{w}^{-\bm{\nu}}(\bm{u})\|_{L^{2}(\mathbb{R}^{d})}^{2}+\mu\|e^{\bm{\nu}}_{w}(\bm{u})\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2},

where λ\lambda and μ\mu are Lamé coefficients such that μ>0\mu>0 and λ+2μ>0\lambda+2\mu>0 and ew𝝂(𝒖)e^{\bm{\nu}}_{w}(\bm{u}) is the nonlocal strain tensor

(92) ew𝝂(𝒖)=𝔊w𝝂𝒖+(𝔊w𝝂𝒖)T2.e^{\bm{\nu}}_{w}(\bm{u})=\frac{\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}+(\mathfrak{G}^{\bm{\nu}}_{w}\bm{u})^{T}}{2}.

We also introduce the nonlocal Naviér operator 𝒫w𝝂\mathcal{P}^{\bm{\nu}}_{w} acting on 𝒖\bm{u} as

(93) 𝒫w𝝂(𝒖):=μw𝝂𝒖(λ+μ)𝔊w𝝂𝔇w𝝂𝒖\mathcal{P}^{\bm{\nu}}_{w}(\bm{u}):=-\mu\mathcal{L}_{w}^{\bm{\nu}}\bm{u}-(\lambda+\mu)\mathfrak{G}_{w}^{\bm{\nu}}\mathfrak{D}_{w}^{-\bm{\nu}}\bm{u}

in Ω\Omega. The goal of this subsection is to show the well-posedness of the following equation

(94) {𝒫w𝝂(𝒖)=𝒇in Ω,𝒖=𝟎in d\Ω.\left\{\begin{aligned} \mathcal{P}^{\bm{\nu}}_{w}(\bm{u})&=\bm{f}\quad\text{in }\ \Omega,\\ \bm{u}&=\bm{0}\quad\text{in }\ \mathbb{R}^{d}\backslash\Omega.\end{aligned}\right.

Similarly as 6.1, when the kernel function ww is supported on Bδ(𝟎)B_{\delta}(\bm{0}), we only need the boundary condition to be imposed on Ω2δ\Ω\Omega_{2\delta}\backslash\Omega. The associated function space to the problem is

𝒮w𝝂(Ω;d)={𝒖=(u1,u2,,ud)T:ui𝒮w𝝂(Ω),i=1,2,,d}.\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d})=\{\bm{u}=(u_{1},u_{2},\cdots,u_{d})^{T}:u_{i}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega),i=1,2,\cdots,d\}.

The weak formulation of the problem is given by

(95) {Find 𝒖𝒮w𝝂(Ω;d) such that: B(𝒖,𝒗)=𝒇,𝒗,𝒗𝒮w𝝂(Ω;d),\left\{\begin{aligned} &\text{Find }\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d})\text{ such that: }\\ &B(\bm{u},\bm{v})=\langle\bm{f},\bm{v}\rangle,\quad\forall\bm{v}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}),\end{aligned}\right.

where 𝒇(𝒮w𝝂(Ω;d))\bm{f}\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}))^{\ast} and the bilinear form B(,):𝒮w𝝂(Ω;d)×𝒮w𝝂(Ω;d)B(\cdot,\cdot):\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d})\times\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d})\to\mathbb{R} is defined as

B(𝒖,𝒗):=μj=1d(𝔊w𝝂uj,𝔊w𝝂vj)L2(d;d)+(λ+μ)(𝔇w𝝂𝒖,𝔇w𝝂𝒗)L2(d).B(\bm{u},\bm{v}):=\mu\sum_{j=1}^{d}(\mathfrak{G}_{w}^{\bm{\nu}}u_{j},\mathfrak{G}_{w}^{\bm{\nu}}v_{j})_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}+(\lambda+\mu)(\mathfrak{D}_{w}^{-\bm{\nu}}\bm{u},\mathfrak{D}_{w}^{-\bm{\nu}}\bm{v})_{L^{2}(\mathbb{R}^{d})}.

To make sense of the weak formulation, one need to show that 𝔇w𝝂𝒖L2(d)\mathfrak{D}_{w}^{-\bm{\nu}}\bm{u}\in L^{2}(\mathbb{R}^{d}) for 𝒖𝒮w𝝂(Ω;d)\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}). This is proved in 3.2.

The following lemma verifies that \mathcal{E} is indeed the energy for the problem (95).

Lemma 6.2.

For 𝐮𝒮w𝛎(Ω;d)\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}),

(96) B(𝒖,𝒖)=2(𝒖).B(\bm{u},\bm{u})=2\mathcal{E}(\bm{u}).
Proof.

Note that B(𝒖,𝒖)=μ𝔊w𝝂𝒖L2(d;d×d)2+(λ+μ)𝔇w𝝂𝒖L2(d)2B(\bm{u},\bm{u})=\mu\|\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2}+(\lambda+\mu)\|\mathfrak{D}^{-\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d})}^{2}. By 3.2, there exists {𝒖(n)}n=1Cc(Ω;d)\{\bm{u}^{(n)}\}_{n=1}^{\infty}\subset C^{\infty}_{c}(\Omega;\mathbb{R}^{d}) such that 𝒢w𝝂𝒖(n)𝔊w𝝂𝒖\mathcal{G}^{\bm{\nu}}_{w}\bm{u}^{(n)}\to\mathfrak{G}^{\bm{\nu}}_{w}\bm{u} in L2(d;d×d)L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d}) and 𝒟w𝝂𝒖(n)𝔇w𝝂𝒖\mathcal{D}_{w}^{-\bm{\nu}}\bm{u}^{(n)}\to\mathfrak{D}_{w}^{-\bm{\nu}}\bm{u} in L2(d)L^{2}(\mathbb{R}^{d}) as nn\to\infty. Therefore, it suffices to show B(𝒖(n),𝒖(n))=2(𝒖(n))B(\bm{u}^{(n)},\bm{u}^{(n)})=2\mathcal{E}(\bm{u}^{(n)}) and let nn tend to infinity. Using the notation 𝒢i\mathcal{G}_{i} and eq. 51 in the proof of 3.2, we obtain

2ew𝝂(𝒖(n))L2(d;d×d)2=2i,j=1dd(12(𝒢iuj(n)(𝒙)+𝒢jui(n)(𝒙)))2𝑑𝒙\displaystyle 2\|e^{\bm{\nu}}_{w}(\bm{u}^{(n)})\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2}=2\sum_{i,j=1}^{d}\int_{\mathbb{R}^{d}}\left(\frac{1}{2}\left(\mathcal{G}_{i}u^{(n)}_{j}(\bm{x})+\mathcal{G}_{j}u^{(n)}_{i}(\bm{x})\right)\right)^{2}d\bm{x}
=\displaystyle= 𝒢w𝝂𝒖(n)L2(d;d×d)2+i,j=1dd𝒢iuj(n)(𝒙)𝒢jui(n)(𝒙)𝑑𝒙\displaystyle\|\mathcal{G}^{\bm{\nu}}_{w}\bm{u}^{(n)}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2}+\sum_{i,j=1}^{d}\int_{\mathbb{R}^{d}}\mathcal{G}_{i}u^{(n)}_{j}(\bm{x})\mathcal{G}_{j}u^{(n)}_{i}(\bm{x})d\bm{x}
=\displaystyle= 𝒢w𝝂𝒖(n)L2(d;d×d)2+𝒟w𝝂𝒖(n)L2(d)2.\displaystyle\|\mathcal{G}^{\bm{\nu}}_{w}\bm{u}^{(n)}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2}+\|\mathcal{D}^{-\bm{\nu}}_{w}\bm{u}^{(n)}\|_{L^{2}(\mathbb{R}^{d})}^{2}.

Thus,

2(𝒖(n))=λ𝒟w𝝂𝒖(n)L2(d)2+μ(𝒢w𝝂𝒖(n)L2(d;d×d)2+𝒟w𝝂𝒖(n)L2(d)2)=B(𝒖(n),𝒖(n)).\begin{split}2\mathcal{E}\left(\bm{u}^{(n)}\right)&=\lambda\|\mathcal{D}^{-\bm{\nu}}_{w}\bm{u}^{(n)}\|_{L^{2}(\mathbb{R}^{d})}^{2}+\mu\left(\|\mathcal{G}^{\bm{\nu}}_{w}\bm{u}^{(n)}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2}+\|\mathcal{D}^{-\bm{\nu}}_{w}\bm{u}^{(n)}\|_{L^{2}(\mathbb{R}^{d})}^{2}\right)\\ &=B\left(\bm{u}^{(n)},\bm{u}^{(n)}\right).\end{split}

Letting nn\to\infty finishes the proof. ∎

Now we are ready to establish the well-posedness of problem (95). In fact, an analogue of the classical Korn’s inequality holds in the nonlocal setting.

Lemma 6.3 (Nonlocal Korn’s inequality).

There exists a constant C=12min(λ+2μ,μ)>0C=\frac{1}{2}\min(\lambda+2\mu,\mu)>0 such that

(97) (𝒖)C𝔊w𝝂𝒖L2(d;d×d)2,𝒖𝒮w𝝂(Ω;d).\mathcal{E}(\bm{u})\geq C\|\mathfrak{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2},\quad\forall\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}).
Proof.

By 6.2 and 3.2, it suffices to show

B(𝒖,𝒖)min(λ+2μ,μ)𝒢w𝝂𝒖L2(d;d×d)2,𝒖Cc(Ω;d).B(\bm{u},\bm{u})\geq\min(\lambda+2\mu,\mu)\|\mathcal{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2},\quad\forall\bm{u}\in C^{\infty}_{c}(\Omega;\mathbb{R}^{d}).

Using the notations and Plancherel’s theorem as in the proof of 3.2 yields

B(𝒖,𝒖)=μ𝒢w𝝂𝒖L2(d;d×d)2+(λ+μ)𝒟w𝝂𝒖L2(d)2\displaystyle B(\bm{u},\bm{u})=\mu\|\mathcal{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2}+(\lambda+\mu)\|\mathcal{D}^{-\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d})}^{2}
=\displaystyle= μd|𝝀w𝝂(𝝃)|2|𝒖^(𝝃)|2𝑑𝝃+(λ+μ)d|𝝀w𝝂(𝝃)T𝒖^(𝝃)|2𝑑𝝃\displaystyle\mu\int_{\mathbb{R}^{d}}\left|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\right|^{2}\left|\hat{\bm{u}}(\bm{\xi})\right|^{2}d\bm{\xi}+(\lambda+\mu)\int_{\mathbb{R}^{d}}\left|\bm{\lambda}_{w}^{-\bm{\nu}}(\bm{\xi})^{T}\hat{\bm{u}}(\bm{\xi})\right|^{2}d\bm{\xi}
\displaystyle\geq min(λ+2μ,μ)d|𝝀w𝝂(𝝃)|2|𝒖^(𝝃)|2𝑑𝝃=min(λ+2μ,μ)𝒢w𝝂𝒖L2(d;d×d)2.\displaystyle\min(\lambda+2\mu,\mu)\int_{\mathbb{R}^{d}}\left|\bm{\lambda}_{w}^{\bm{\nu}}(\bm{\xi})\right|^{2}\left|\hat{\bm{u}}(\bm{\xi})\right|^{2}d\bm{\xi}=\min(\lambda+2\mu,\mu)\|\mathcal{G}^{\bm{\nu}}_{w}\bm{u}\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d\times d})}^{2}.

Theorem 6.2.

The nonlocal linear elasticity problem (95) is well-posed. More precisely, for any 𝐟(𝒮w𝛎(Ω;d))\bm{f}\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}))^{\ast}, there exists a unique solution 𝐮𝒮w𝛎(Ω;d)\bm{u}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}) such that

𝒖𝒮w𝝂(Ω;d)c𝒇(𝒮w𝝂(Ω;d)),\|\bm{u}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d})}\leq c\|\bm{f}\|_{(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{d}))^{\ast}},

where c=min(λ+2μ,μ)1c=\min(\lambda+2\mu,\mu)^{-1} is a positive constant.

Proof.

The bilinear form B(,)B(\cdot,\cdot) is coercive by 6.2 and 6.3, and is bounded by 3.2 and 3.2. Applying the Lax-Milgram theorem yields the result. ∎

6.3. Nonlocal Helmholtz decomposition

In this subsection, we always assume that d=2d=2 or d=3d=3. The nonlocal vector calculus identities in Section 3.2 will be used to obtain the nonlocal Helmholtz decomposition for d=2d=2 and d=3d=3. These results extend similar studies in [38] for periodic functions.

Theorem 6.3.

Let 𝐮(𝒮w𝛎(Ω;2))\bm{u}\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}))^{\ast}. There exist scalar potentials p𝛎,q𝛎L2(2)p^{\bm{\nu}},\,q^{\bm{\nu}}\in L^{2}(\mathbb{R}^{2}) such that

𝒖=𝔊w𝝂p𝝂+(0110)𝔊w𝝂q𝝂(𝒮w𝝂(Ω;2)).\bm{u}=\mathfrak{G}_{w}^{\bm{\nu}}p^{\bm{\nu}}+\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\mathfrak{G}_{w}^{-\bm{\nu}}q^{\bm{\nu}}\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}))^{\ast}.

In addition, there exists a constant CC depending on the Poincaré constant Π\Pi such that

p𝝂L2(2)+q𝝂L2(2)C𝒖(𝒮w𝝂(Ω;2)).\|p^{\bm{\nu}}\|_{L^{2}(\mathbb{R}^{2})}+\|q^{\bm{\nu}}\|_{L^{2}(\mathbb{R}^{2})}\leq C\|\bm{u}\|_{(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}))^{\ast}}.
Proof.

Applying 6.1 with ϵ=1\epsilon=1 and 𝒃=𝟎\bm{b}=\bm{0} componentwise, it follows that there exists a unique function 𝒇𝒮w𝝂(Ω;2)\bm{f}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}) such that w𝝂𝒇=𝒖-\mathcal{L}_{w}^{\bm{\nu}}\bm{f}=\bm{u} with

𝒇𝒮w𝝂(Ω;2)c𝒖(𝒮w𝝂(Ω;2)),\|\bm{f}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2})}\leq c\|\bm{u}\|_{(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}))^{\ast}},

where c=c(w,𝝂,Ω)>0c=c(w,\bm{\nu},\Omega)>0. Let

p𝝂=𝔇w𝝂𝒇 and q𝝂=𝔇w𝝂[(0110)𝒇].p^{\bm{\nu}}=-\mathfrak{D}_{w}^{-\bm{\nu}}\bm{f}\ \text{ and }\ q^{\bm{\nu}}=\mathfrak{D}_{w}^{\bm{\nu}}\left[\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\bm{f}\right].

By 3.2, we have p𝝂,q𝝂L2(2)p^{\bm{\nu}},q^{\bm{\nu}}\in L^{2}(\mathbb{R}^{2}), and p𝝂L2(2)+q𝝂L2(2)C~𝒇𝒮w𝝂(Ω;2)C𝒖(𝒮w𝝂(Ω;2))\|p^{\bm{\nu}}\|_{L^{2}(\mathbb{R}^{2})}+\|q^{\bm{\nu}}\|_{L^{2}(\mathbb{R}^{2})}\leq\tilde{C}\|\bm{f}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2})}\leq C\|\bm{u}\|_{(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{2}))^{\ast}}. Then by 2.6 we obtain

𝒖\displaystyle\bm{u} =w𝝂𝒇=𝔊w𝝂p𝝂+(0110)𝔊w𝝂q𝝂.\displaystyle=-\mathcal{L}_{w}^{\bm{\nu}}\bm{f}=\mathfrak{G}_{w}^{\bm{\nu}}p^{\bm{\nu}}+\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\mathfrak{G}_{w}^{-\bm{\nu}}q^{\bm{\nu}}.

This finishes the proof. ∎

Theorem 6.4.

Let 𝐮(𝒮w𝛎(Ω;3))\bm{u}\in(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}))^{\ast}. There exist a scalar potential p𝛎L2(3)p^{\bm{\nu}}\in L^{2}(\mathbb{R}^{3}) and a vector potential 𝐯𝛎L2(3;3)\bm{v}^{\bm{\nu}}\in L^{2}(\mathbb{R}^{3};\mathbb{R}^{3}) such that

(98) 𝒖=𝔊w𝝂p𝝂+w𝝂𝒗𝝂,\bm{u}=\mathfrak{G}_{w}^{\bm{\nu}}p^{\bm{\nu}}+\mathfrak{C}_{w}^{-\bm{\nu}}\bm{v}^{\bm{\nu}},

with

(99) 𝔇w𝝂𝒗𝝂=0,\mathfrak{D}_{w}^{\bm{\nu}}\bm{v}^{\bm{\nu}}=0,

where the above equations are understood in (𝒮w𝛎(Ω;3))(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}))^{\ast} and (𝒮w𝛎(Ω))(\mathcal{S}_{w}^{-\bm{\nu}}(\Omega))^{\ast}, respectively. In addition, there exists a constant CC depending on the Poincaré constant Π\Pi such that

p𝝂L2(3)+𝒗𝝂L2(3;3)C𝒖(𝒮w𝝂(Ω;3)).\|p^{\bm{\nu}}\|_{L^{2}(\mathbb{R}^{3})}+\|\bm{v}^{\bm{\nu}}\|_{L^{2}(\mathbb{R}^{3};\mathbb{R}^{3})}\leq C\|\bm{u}\|_{(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}))^{\ast}}.
Proof.

As in the proof of Theorem 6.3, there exists 𝒇𝒮w𝝂(Ω;3)\bm{f}\in\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}) such that w𝝂𝒇=𝒖-\mathcal{L}_{w}^{\bm{\nu}}\bm{f}=\bm{u} and

(100) 𝒇𝒮w𝝂(Ω;3)c𝒖(𝒮w𝝂(Ω;3)).\|\bm{f}\|_{\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3})}\leq c\|\bm{u}\|_{(\mathcal{S}_{w}^{\bm{\nu}}(\Omega;\mathbb{R}^{3}))^{\ast}}.

We choose

p𝝂=𝔇w𝝂𝒇 and 𝒗𝝂=w𝝂𝒇,p^{\bm{\nu}}=-\mathfrak{D}_{w}^{-\bm{\nu}}\bm{f}\ \text{ and }\ \bm{v}^{\bm{\nu}}=\mathfrak{C}_{w}^{\bm{\nu}}\bm{f},

and use eq. 35 to derive eq. 98. The computation is staightforward and thus omitted. By 3.4, eq. 99 holds. Similar to the proof of Theorem 6.3, the final estimate follows from 3.2, 3.3 and eq. 100. ∎

Remark 6.2.

If the kernel function ww is integrable, then by 4.1, for any uL2(d)u\in L^{2}(\mathbb{R}^{d}) and 𝐯L2(d;d)\bm{v}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}), we have 𝔊w𝛎u=𝒢w𝛎uL2(d;d)\mathfrak{G}_{w}^{\bm{\nu}}u=\mathcal{G}_{w}^{\bm{\nu}}u\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}) and w𝛎𝐯=𝒞w𝛎𝐯L2(d;d)\mathfrak{C}_{w}^{\bm{\nu}}\bm{v}=\mathcal{C}_{w}^{\bm{\nu}}\bm{v}\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). Therefore, the two components in the Helmholtz decomposition in 6.3 or 6.4 are orthogonal in L2(d;d)L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}) as a result of integration by parts together with 3.2 (d=2d=2) or 3.4 (d=3d=3).

Remark 6.3.

If the kernel function ww has compact support, then the potentials in 6.3 and 6.4 vanish outside a compact set. More specifically, if suppwBδ(𝟎)\text{supp}\ w\subset B_{\delta}(\bm{0}) for δ>0\delta>0, then the p𝛎,q𝛎L2(Ωδ)p^{\bm{\nu}},q^{\bm{\nu}}\in L^{2}(\Omega_{\delta}) and 𝐯𝛎L2(Ωδ;3)\bm{v}^{\bm{\nu}}\in L^{2}(\Omega_{\delta};\mathbb{R}^{3}) where Ωδ={𝐱d:dist(𝐱,Ω)<δ}\Omega_{\delta}=\{\bm{x}\in\mathbb{R}^{d}:\text{dist}(\bm{x},\Omega)<\delta\}.

7. Conclusion

In this paper, we have studied nonlocal half-ball gradient, divergence and curl operators with a rather general class of kernels. These nonlocal operators can be generalized to distributional operators upon which a Sobolev-type space is defined. For this function space, the set of smooth functions with compact support is proved to be dense. Moreover, a nonlocal Poincaré inequality on bounded domains is established, which is crucial to study the well-posedness of nonlocal Dirichlet boundary value problems such as nonlocal convection-diffusion and nonlocal correspondence model of linear elasticity and to prove a nonlocal Helmholtz decomposition.

This work provides a rigorous mathematical analysis on the stability of some linear nonlocal problems with homogeneous Dirichlet boundary, thus generalizes the analytical results in [38] where the domains are periodic cells. While we mainly focused on the analysis of these nonlocal problems, standard Galerkin approximations to these problems are also natural based on the Poincaré inequality and the density result. It would also be interesting to investigate Petrov-Galerkin methods for the nonlocal convection-diffusion problems [39], as well as mixed-type methods for them [17]. Other problems such as nonlocal elasticity models in heterogeneous media and the Stokes system in [22, 38] may also be studied in the future. As for the analysis, our approach relies heavily on Fourier analysis which is powerful but limited to L2L^{2} formulation. The nonlocal LpL^{p} Poincaré inequality for half-ball gradient operator on bounded domains is still open to investigation. In addition, Poincaré inequality for Neumann type boundary is also interesting to be explored in the future. We note that in this work the dependence of the Poincaré constant on the kernel function is implicit as a result of argument by contradiction. Further investigation on how the constant depends on the kernel function is needed, and following [41], a sharper version of Poincaré inequality may be considered by establishing compactness results analogous to those in [10]. Last but not least, it remains of great interest to develop nonlocal exterior calculus and geometric structures that connect the corresponding discrete theories and continuous local theories [3, 7, 32, 36].

Acknowledgements

This research was supported in part by NSF grants DMS-2111608 and DMS-2240180. The authors thank Qiang Du, Tadele Mengesha, and James Scott for their helpful discussions. The authors would also like to thank the anonymous reviewers for their valuable comments and suggestions.

Appendix A

Proof of 2.2.

Let uW1,p(d)u\in W^{1,p}(\mathbb{R}^{d}). To use Lebesgue dominated convergence theorem to show the principal value integral coincide with the usual Lebesgue integral, we construct the majorizing function

g𝒙(𝒚):=|u(𝒚)u(𝒙)|w(𝒚𝒙),𝒚d,g_{\bm{x}}(\bm{y}):=|u(\bm{y})-u(\bm{x})|w(\bm{y}-\bm{x}),\quad\bm{y}\in\mathbb{R}^{d},

and show that g𝒙L1(d)g_{\bm{x}}\in L^{1}(\mathbb{R}^{d}) for a.e. 𝒙d\bm{x}\in\mathbb{R}^{d}. This follows from the fact that the function 𝒙dg𝒙(𝒚)𝑑𝒚Lp(d)\bm{x}\mapsto\int_{\mathbb{R}^{d}}g_{\bm{x}}(\bm{y})d\bm{y}\in L^{p}(\mathbb{R}^{d}). When p=p=\infty, this is obvious. To show this fact for 1p<1\leq p<\infty, first note that

d|dg𝒙(𝒚)𝑑𝒚|p𝑑𝒙=d||𝒚𝒙|<1g𝒙(𝒚)𝑑𝒚+|𝒚𝒙|>1g𝒙(𝒚)𝑑𝒚|p𝑑𝒙2p1d(|𝒚𝒙|<1g𝒙(𝒚)𝑑𝒚)pd𝒙+d(|𝒚𝒙|>1g𝒙(𝒚)𝑑𝒚)p𝑑𝒙.\begin{split}\int_{\mathbb{R}^{d}}\left|\int_{\mathbb{R}^{d}}g_{\bm{x}}(\bm{y})d\bm{y}\right|^{p}d\bm{x}&=\int_{\mathbb{R}^{d}}\left|\int_{|\bm{y}-\bm{x}|<1}g_{\bm{x}}(\bm{y})d\bm{y}+\int_{|\bm{y}-\bm{x}|>1}g_{\bm{x}}(\bm{y})d\bm{y}\right|^{p}d\bm{x}\\ \leq 2^{p-1}\int_{\mathbb{R}^{d}}&\left(\int_{|\bm{y}-\bm{x}|<1}g_{\bm{x}}(\bm{y})d\bm{y}\right)^{p}d\bm{x}+\int_{\mathbb{R}^{d}}\left(\int_{|\bm{y}-\bm{x}|>1}g_{\bm{x}}(\bm{y})d\bm{y}\right)^{p}d\bm{x}.\end{split}

Then by Hölder’s inequality,

d(|𝒚𝒙|<1g𝒙(𝒚)𝑑𝒚)p𝑑𝒙=d(|𝒛|<1|u(𝒙+𝒛)u(𝒙)||𝒛||𝒛|w(𝒛)𝑑𝒛)p𝑑𝒙(|𝒛|<1|𝒛|w(𝒛)𝑑𝒛)p1d|𝒛|<1|𝒛|w(𝒛)|u(𝒙+𝒛)u(𝒙)|p|𝒛|p𝑑𝒛𝑑𝒙(Mw1)puLp(d)p,\begin{split}&\int_{\mathbb{R}^{d}}\left(\int_{|\bm{y}-\bm{x}|<1}g_{\bm{x}}(\bm{y})d\bm{y}\right)^{p}d\bm{x}\\ =&\int_{\mathbb{R}^{d}}\left(\int_{|\bm{z}|<1}\frac{|u(\bm{x}+\bm{z})-u(\bm{x})|}{|\bm{z}|}|\bm{z}|w(\bm{z})d\bm{z}\right)^{p}d\bm{x}\\ \leq&\left(\int_{|\bm{z}|<1}|\bm{z}|w(\bm{z})d\bm{z}\right)^{p-1}\int_{\mathbb{R}^{d}}\int_{|\bm{z}|<1}|\bm{z}|w(\bm{z})\frac{|u(\bm{x}+\bm{z})-u(\bm{x})|^{p}}{|\bm{z}|^{p}}d\bm{z}d\bm{x}\\ \leq&(M_{w}^{1})^{p}\|\nabla u\|_{L^{p}(\mathbb{R}^{d})}^{p},\end{split}

where we used inequality (see Proposition 9.3 in [11])

d|u(𝒙+𝒛)u(𝒙)|p|𝒛|p𝑑𝒙uLp(d)p,𝒛d\{𝟎}.\int_{\mathbb{R}^{d}}\frac{|u(\bm{x}+\bm{z})-u(\bm{x})|^{p}}{|\bm{z}|^{p}}d\bm{x}\leq\|\nabla u\|_{L^{p}(\mathbb{R}^{d})}^{p},\quad\bm{z}\in\mathbb{R}^{d}\backslash\{\bm{0}\}.

Applying the same techniques it follows that

d(|𝒚𝒙|>1g𝒙(𝒚)𝑑𝒚)p𝑑𝒙2p1d(|𝒚𝒙|>1|u(𝒚)|w(𝒚𝒙)𝑑𝒚)p+(|𝒚𝒙|>1|u(𝒙)|w(𝒚𝒙)𝑑𝒚)pd𝒙2p(Mw2)puLp(d)p.\begin{split}&\int_{\mathbb{R}^{d}}\left(\int_{|\bm{y}-\bm{x}|>1}g_{\bm{x}}(\bm{y})d\bm{y}\right)^{p}d\bm{x}\\ \leq&2^{p-1}\int_{\mathbb{R}^{d}}\left(\int_{|\bm{y}-\bm{x}|>1}|u(\bm{y})|w(\bm{y}-\bm{x})d\bm{y}\right)^{p}+\left(\int_{|\bm{y}-\bm{x}|>1}|u(\bm{x})|w(\bm{y}-\bm{x})d\bm{y}\right)^{p}d\bm{x}\\ \leq&2^{p}(M_{w}^{2})^{p}\|u\|_{L^{p}(\mathbb{R}^{d})}^{p}.\end{split}

Combining the above estimates, there exists a constant C>0C>0 depending on pp such that

(101) (d|dg𝒙(𝒚)𝑑𝒚|p𝑑𝒙)1pC(Mw1uLp(d)+Mw2uLp(d)),uW1,p(d).\left(\int_{\mathbb{R}^{d}}\left|\int_{\mathbb{R}^{d}}g_{\bm{x}}(\bm{y})d\bm{y}\right|^{p}d\bm{x}\right)^{\frac{1}{p}}\leq C\left(M_{w}^{1}\|\nabla u\|_{L^{p}(\mathbb{R}^{d})}+M_{w}^{2}\|u\|_{L^{p}(\mathbb{R}^{d})}\right),\quad u\in W^{1,p}(\mathbb{R}^{d}).

Therefore, g𝒙L1(d)g_{\bm{x}}\in L^{1}(\mathbb{R}^{d}) for a.e. 𝒙d\bm{x}\in\mathbb{R}^{d} and by Lebesgue dominated convergence theorem, equalities (5) hold for a.e. xdx\in\mathbb{R}^{d}. Since |𝒢w𝝂u(𝒙)|dg𝒙(𝒚)𝑑𝒚|\mathcal{G}^{\bm{\nu}}_{w}u(\bm{x})|\leq\int_{\mathbb{R}^{d}}g_{\bm{x}}(\bm{y})d\bm{y}, the estimate (11) follows from (101). Similar proofs hold for 𝒟w𝝂\mathcal{D}^{\bm{\nu}}_{w} and 𝒞w𝝂\mathcal{C}^{\bm{\nu}}_{w} and are omitted. ∎

Proof of 2.1.
  1. (1)

    Since w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(d×d)w(\bm{x}-\bm{y})\left|\bm{u}(\bm{x})-\bm{u}(\bm{y})\right|\in L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}) and 𝒗Cc1(d;d×N)\bm{v}\in C^{1}_{c}(\mathbb{R}^{d};\mathbb{R}^{d\times N}), one can show by Lebesgue dominated convergence theorem that

    (102) d𝒢w𝝂u(𝒙):𝒗(𝒙)d𝒙=limϵ0ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|(𝒖(𝒚)𝒖(𝒙)):𝒗(𝒙)d𝒚d𝒙,\int_{\mathbb{R}^{d}}\mathcal{G}^{\bm{\nu}}_{w}u(\bm{x}):\bm{v}(\bm{x})d\bm{x}=\lim_{\epsilon\to 0}\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes(\bm{u}(\bm{y})-\bm{u}(\bm{x})):\bm{v}(\bm{x})d\bm{y}d\bm{x},

    where ϵ2d:=d×d\{(𝒙,𝒚)2d:|𝒙𝒚|ϵ}\mathbb{R}^{2d}_{\epsilon}:=\mathbb{R}^{d}\times\mathbb{R}^{d}\backslash\{(\bm{x},\bm{y})\in\mathbb{R}^{2d}:|\bm{x}-\bm{y}|\leq\epsilon\}. Similarly,

    d𝒖(𝒙)𝒟w𝝂𝒗(𝒙)𝑑𝒙=d𝒖(𝒙)limϵ0gϵ(𝒙)d𝒙\displaystyle-\int_{\mathbb{R}^{d}}\bm{u}(\bm{x})\cdot\mathcal{D}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}\bm{u}(\bm{x})\cdot\lim_{\epsilon\to 0}g_{\epsilon}(\bm{x})d\bm{x}
    =\displaystyle= limϵ0ϵ2d𝒖(𝒙)[𝒚T𝒙T|𝒚𝒙|(χ𝝂(𝒙𝒚)𝒗(𝒚)+χ𝝂(𝒚𝒙)𝒗(𝒙))]Tw(𝒚𝒙)𝑑𝒚𝑑𝒙,\displaystyle-\lim_{\epsilon\to 0}\iint_{\mathbb{R}^{2d}_{\epsilon}}\bm{u}(\bm{x})\cdot\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{v}(\bm{x}))\right]^{T}w(\bm{y}-\bm{x})d\bm{y}d\bm{x},

    where

    gϵ(𝒙):\displaystyle g_{\epsilon}(\bm{x}): =d\Bϵ(𝒙)[𝒚T𝒙T|𝒚𝒙|(χ𝝂(𝒙𝒚)(𝒗(𝒚)𝒗(𝒙)))]Tw(𝒚𝒙)𝑑𝒚,\displaystyle=\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\chi_{\bm{\nu}}(\bm{x}-\bm{y})(\bm{v}(\bm{y})-\bm{v}(\bm{x})))\right]^{T}w(\bm{y}-\bm{x})d\bm{y},
    =d\Bϵ(𝒙)[𝒚T𝒙T|𝒚𝒙|(χ𝝂(𝒙𝒚)𝒗(𝒚)+χ𝝂(𝒚𝒙)𝒗(𝒙))]Tw(𝒚𝒙)𝑑𝒚\displaystyle=\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{v}(\bm{x}))\right]^{T}w(\bm{y}-\bm{x})d\bm{y}
    d\Bϵ(𝒙)[𝒚T𝒙T|𝒚𝒙|𝒗(𝒙)]Tw(𝒚𝒙)𝑑𝒚\displaystyle\qquad-\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}\bm{v}(\bm{x})\right]^{T}w(\bm{y}-\bm{x})d\bm{y}
    =d\Bϵ(𝒙)[𝒚T𝒙T|𝒚𝒙|(χ𝝂(𝒙𝒚)𝒗(𝒚)+χ𝝂(𝒚𝒙)𝒗(𝒙))]Tw(𝒚𝒙)𝑑𝒚\displaystyle=\int_{\mathbb{R}^{d}\backslash B_{\epsilon}(\bm{x})}\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{v}(\bm{x}))\right]^{T}w(\bm{y}-\bm{x})d\bm{y}

    where we have used χ𝝂(𝒙𝒚)+χ𝝂(𝒚𝒙)=1\chi_{\bm{\nu}}(\bm{x}-\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})=1. The change of order of limitation and integration is again justified by Lebesgue dominated convergence theorem due to 𝒖L1(d;N)\bm{u}\in L^{1}(\mathbb{R}^{d};\mathbb{R}^{N}) and 𝒗Cc1(d;d×N)\bm{v}\in C_{c}^{1}(\mathbb{R}^{d};\mathbb{R}^{d\times N}).

    Therefore, it suffices to prove that

    ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|(𝒖(𝒚)𝒖(𝒙)):𝒗(𝒙)d𝒚d𝒙\displaystyle\quad\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes(\bm{u}(\bm{y})-\bm{u}(\bm{x})):\bm{v}(\bm{x})d\bm{y}d\bm{x}
    =ϵ2d𝒖(𝒙)[𝒚T𝒙T|𝒚𝒙|(χ𝝂(𝒙𝒚)𝒗(𝒚)+χ𝝂(𝒚𝒙)𝒗(𝒙))]Tw(𝒚𝒙)𝑑𝒚𝑑𝒙.\displaystyle=-\iint_{\mathbb{R}^{2d}_{\epsilon}}\bm{u}(\bm{x})\cdot\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{v}(\bm{x}))\right]^{T}w(\bm{y}-\bm{x})d\bm{y}d\bm{x}.

    Applying Fubini’s theorem completes the proof as

    ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|(𝒖(𝒚)𝒖(𝒙)):𝒗(𝒙)d𝒚d𝒙\displaystyle\quad\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes(\bm{u}(\bm{y})-\bm{u}(\bm{x})):\bm{v}(\bm{x})d\bm{y}d\bm{x}
    =ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|𝒖(𝒚):𝒗(𝒙)d𝒚d𝒙\displaystyle=\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes\bm{u}(\bm{y}):\bm{v}(\bm{x})d\bm{y}d\bm{x}
    ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|𝒖(𝒙):𝒗(𝒙)d𝒚d𝒙\displaystyle\quad-\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes\bm{u}(\bm{x}):\bm{v}(\bm{x})d\bm{y}d\bm{x}
    =ϵ2dχ𝝂(𝒙𝒚)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|𝒖(𝒙):𝒗(𝒚)d𝒚d𝒙\displaystyle=-\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{x}-\bm{y})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes\bm{u}(\bm{x}):\bm{v}(\bm{y})d\bm{y}d\bm{x}
    ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|𝒖(𝒙):𝒗(𝒙)d𝒚d𝒙\displaystyle\quad-\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes\bm{u}(\bm{x}):\bm{v}(\bm{x})d\bm{y}d\bm{x}
    =ϵ2d𝒖(𝒙)[𝒚T𝒙T|𝒚𝒙|(χ𝝂(𝒙𝒚)𝒗(𝒚)+χ𝝂(𝒚𝒙)𝒗(𝒙))]Tw(𝒚𝒙)𝑑𝒚𝑑𝒙.\displaystyle=-\iint_{\mathbb{R}^{2d}_{\epsilon}}\bm{u}(\bm{x})\cdot\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\chi_{\bm{\nu}}(\bm{x}-\bm{y})\bm{v}(\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})\bm{v}(\bm{x}))\right]^{T}w(\bm{y}-\bm{x})d\bm{y}d\bm{x}.
  2. (2)

    Since w(𝒙𝒚)|𝒖(𝒙)𝒖(𝒚)|L1(d×d)w(\bm{x}-\bm{y})|\bm{u}(\bm{x})-\bm{u}(\bm{y})|\in L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}) and 𝒗Cc1(d;N)\bm{v}\in C^{1}_{c}(\mathbb{R}^{d};\mathbb{R}^{N}), by Lebesgue dominated convergence theorem one can show that

    (103) d𝒟w𝝂𝒖(𝒙)𝒗(𝒙)𝑑𝒙=limϵ0ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)[𝒚T𝒙T|𝒚𝒙|(𝒖(𝒚)𝒖(𝒙))]T𝒗(𝒙)𝑑𝒚𝑑𝒙,\begin{split}&\int_{\mathbb{R}^{d}}\mathcal{D}^{\bm{\nu}}_{w}\bm{u}(\bm{x})\cdot\bm{v}(\bm{x})d\bm{x}\\ =&\lim_{\epsilon\to 0}\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\bm{u}(\bm{y})-\bm{u}(\bm{x}))\right]^{T}\cdot\bm{v}(\bm{x})d\bm{y}d\bm{x},\end{split}

    where ϵ2d:=d×d\{(𝒙,𝒚)2d:|𝒙𝒚|ϵ}\mathbb{R}^{2d}_{\epsilon}:=\mathbb{R}^{d}\times\mathbb{R}^{d}\backslash\{(\bm{x},\bm{y})\in\mathbb{R}^{2d}:|\bm{x}-\bm{y}|\leq\epsilon\}. Similarly, by the same reasoning as in the proof of 2.1(1),

    d𝒖(𝒙):𝒢w𝝂𝒗(𝒙)d𝒙\displaystyle-\int_{\mathbb{R}^{d}}\bm{u}(\bm{x}):\mathcal{G}^{\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}
    =\displaystyle= limϵ0ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒖(𝒙):𝒚𝒙|𝒚𝒙|(𝒗(𝒚)𝒗(𝒙))d𝒚d𝒙.\displaystyle-\lim_{\epsilon\to 0}\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\bm{u}(\bm{x}):\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes(\bm{v}(\bm{y})-\bm{v}(\bm{x}))d\bm{y}d\bm{x}.

    Therefore, it suffices to prove that

    ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)[𝒚T𝒙T|𝒚𝒙|(𝒖(𝒚)𝒖(𝒙))]T𝒗(𝒙)𝑑𝒚𝑑𝒙\displaystyle\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\left[\frac{\bm{y}^{T}-\bm{x}^{T}}{|\bm{y}-\bm{x}|}(\bm{u}(\bm{y})-\bm{u}(\bm{x}))\right]^{T}\cdot\bm{v}(\bm{x})d\bm{y}d\bm{x}
    =\displaystyle= ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒖(𝒙):𝒚𝒙|𝒚𝒙|(𝒗(𝒚)𝒗(𝒙))d𝒚d𝒙.\displaystyle-\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\bm{u}(\bm{x}):\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\otimes(\bm{v}(\bm{y})-\bm{v}(\bm{x}))d\bm{y}d\bm{x}.

    Applying Fubini’s theorem as in the proof of 2.1(1) gives the desired result.

  3. (3)

    By similar reasoning as the proof of 2.1(1) and 2.1(2), we have

    d𝒞w𝝂𝒖(𝒙)𝒗(𝒙)𝑑𝒙=limϵ0ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|×(𝒖(𝒚)𝒖(𝒙))𝒗(𝒙)𝑑𝒚𝑑𝒙,\int_{\mathbb{R}^{d}}\mathcal{C}^{\bm{\nu}}_{w}\bm{u}(\bm{x})\cdot\bm{v}(\bm{x})d\bm{x}=\lim_{\epsilon\to 0}\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\times(\bm{u}(\bm{y})-\bm{u}(\bm{x}))\cdot\bm{v}(\bm{x})d\bm{y}d\bm{x},

    and

    d𝒖(𝒙)𝒞w𝝂𝒗(𝒙)𝑑𝒙=limϵ0ϵ2dχ𝝂(𝒙𝒚)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|×(𝒗(𝒚)𝒗(𝒙))𝒖(𝒙)𝑑𝒚𝑑𝒙.\int_{\mathbb{R}^{d}}\bm{u}(\bm{x})\cdot\mathcal{C}^{-\bm{\nu}}_{w}\bm{v}(\bm{x})d\bm{x}=\lim_{\epsilon\to 0}\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{x}-\bm{y})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\times(\bm{v}(\bm{y})-\bm{v}(\bm{x}))\cdot\bm{u}(\bm{x})d\bm{y}d\bm{x}.

    Using Fubini’s theorem and the two identities χ𝝂(𝒙𝒚)+χ𝝂(𝒚𝒙)=1\chi_{\bm{\nu}}(\bm{x}-\bm{y})+\chi_{\bm{\nu}}(\bm{y}-\bm{x})=1 and 𝒂(𝒃×𝒄)=𝒄(𝒃×𝒂)\bm{a}\cdot(\bm{b}\times\bm{c})=-\bm{c}\cdot(\bm{b}\times\bm{a}), one can show

    ϵ2dχ𝝂(𝒚𝒙)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|×(𝒖(𝒚)𝒖(𝒙))𝒗(𝒙)𝑑𝒚𝑑𝒙=ϵ2dχ𝝂(𝒙𝒚)w(𝒚𝒙)𝒚𝒙|𝒚𝒙|×(𝒗(𝒚)𝒗(𝒙))𝒖(𝒙)𝑑𝒚𝑑𝒙,\begin{split}&\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{y}-\bm{x})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\times(\bm{u}(\bm{y})-\bm{u}(\bm{x}))\cdot\bm{v}(\bm{x})d\bm{y}d\bm{x}\\ =&\iint_{\mathbb{R}^{2d}_{\epsilon}}\chi_{\bm{\nu}}(\bm{x}-\bm{y})w(\bm{y}-\bm{x})\frac{\bm{y}-\bm{x}}{|\bm{y}-\bm{x}|}\times(\bm{v}(\bm{y})-\bm{v}(\bm{x}))\cdot\bm{u}(\bm{x})d\bm{y}d\bm{x},\end{split}

    for any ϵ>0\epsilon>0, and therefore the desired result is implied.

Proof of 3.2.

Note that τ𝒂uL2(d)\tau_{\bm{a}}u\in L^{2}(\mathbb{R}^{d}) is obvious. To show τ𝒂u𝒮w𝝂(d;d)\tau_{\bm{a}}u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d};\mathbb{R}^{d}), it suffices to show 𝔊w𝝂(τ𝒂u)L2(d;d)\mathfrak{G}^{\bm{\nu}}_{w}(\tau_{\bm{a}}u)\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). We claim that 𝔊w𝝂(τ𝒂u)=τ𝒂(𝔊w𝝂u)L2(d;d)\mathfrak{G}^{\bm{\nu}}_{w}(\tau_{\bm{a}}u)=\tau_{\bm{a}}(\mathfrak{G}^{\bm{\nu}}_{w}u)\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). Indeed, for any ϕCc(d;d)\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}),

dτ𝒂u(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙\displaystyle\int_{\mathbb{R}^{d}}\tau_{\bm{a}}u(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x} =du(𝒙)(𝒟w𝝂ϕ)(𝒙𝒂)𝑑𝒙\displaystyle=\int_{\mathbb{R}^{d}}u(\bm{x})(\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi})(\bm{x}-\bm{a})d\bm{x}
=du(𝒙)(τ𝒂𝒟w𝝂ϕ)(𝒙)𝑑𝒙\displaystyle=\int_{\mathbb{R}^{d}}u(\bm{x})(\tau_{-\bm{a}}\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi})(\bm{x})d\bm{x}
=du(𝒙)(𝒟w𝝂(τ𝒂ϕ))(𝒙)𝑑𝒙\displaystyle=\int_{\mathbb{R}^{d}}u(\bm{x})(\mathcal{D}^{-\bm{\nu}}_{w}(\tau_{-\bm{a}}\bm{\phi}))(\bm{x})d\bm{x}
=d𝔊w𝝂u(𝒙)τ𝒂ϕ(𝒙)𝑑𝒙\displaystyle=-\int_{\mathbb{R}^{d}}\mathfrak{G}^{\bm{\nu}}_{w}u(\bm{x})\cdot\tau_{-\bm{a}}\bm{\phi}(\bm{x})d\bm{x}
=dτ𝒂(𝔊w𝝂u)(𝒙)ϕ(𝒙)𝑑𝒙,\displaystyle=-\int_{\mathbb{R}^{d}}\tau_{\bm{a}}(\mathfrak{G}^{\bm{\nu}}_{w}u)(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x},

where 𝒟w𝝂(τ𝒂ϕ)=τ𝒂𝒟w𝝂ϕ\mathcal{D}^{-\bm{\nu}}_{w}(\tau_{-\bm{a}}\bm{\phi})=\tau_{-\bm{a}}\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi} can be easily checked. Therefore, the claim is true and thus τ𝒂u𝒮w𝝂(d)\tau_{\bm{a}}u\in\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d}).

To show the continuity, first note that

lim|𝒂|0τ𝒂uuL2(d)=0\lim_{|\bm{a}|\to 0}\|\tau_{\bm{a}}u-u\|_{L^{2}(\mathbb{R}^{d})}=0

by continuity of translation in L2(d)L^{2}(\mathbb{R}^{d}). Then using the claim above and the continuity of translation in L2(d;d)L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}), we have

𝔊w𝝂(τ𝒂u)𝔊w𝝂uL2(d;d)=τ𝒂(𝔊w𝝂u)𝔊w𝝂uL2(d;d)0,|𝒂|0.\|\mathfrak{G}^{\bm{\nu}}_{w}(\tau_{\bm{a}}u)-\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}=\|\tau_{\bm{a}}(\mathfrak{G}^{\bm{\nu}}_{w}u)-\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}\to 0,\quad|\bm{a}|\to 0.

Hence,

lim|𝒂|0τ𝒂uu𝒮w𝝂(d)=0.\lim_{|\bm{a}|\to 0}\|\tau_{\bm{a}}u-u\|_{\mathcal{S}_{w}^{\bm{\nu}}(\mathbb{R}^{d})}=0.

Proof of 3.3.

Since uL2(d)u\in L^{2}(\mathbb{R}^{d}), by the property of mollification, ηϵuL2(d)\eta_{\epsilon}*u\in L^{2}(\mathbb{R}^{d}) and

limϵ0ηϵuuL2(d)=0.\lim_{\epsilon\to 0}\|\eta_{\epsilon}*u-u\|_{L^{2}(\mathbb{R}^{d})}=0.

We claim that

𝔊w𝝂(ηϵu)=ηϵ𝔊w𝝂uL2(d;d).\mathfrak{G}^{\bm{\nu}}_{w}(\eta_{\epsilon}*u)=\eta_{\epsilon}*\mathfrak{G}^{\bm{\nu}}_{w}u\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}).

To show the claim, we need to prove that

(104) d(ηϵ𝔊w𝝂u)(𝒙)ϕ(𝒙)𝑑𝒙=d(ηϵu)(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙,ϕCc(d;d).\int_{\mathbb{R}^{d}}(\eta_{\epsilon}*\mathfrak{G}^{\bm{\nu}}_{w}u)(\bm{x})\bm{\phi}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}(\eta_{\epsilon}*u)(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x},\quad\forall\bm{\phi}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d}).

For the right-hand side, we use Fubini’s theorem to get

d(ηϵu)(𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙\displaystyle-\int_{\mathbb{R}^{d}}(\eta_{\epsilon}*u)(\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x} =ddηϵ(𝒙𝒚)u(𝒚)𝑑𝒚𝒟w𝝂ϕ(𝒙)𝑑𝒙\displaystyle=-\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}\eta_{\epsilon}(\bm{x}-\bm{y})u(\bm{y})d\bm{y}\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x}
=du(𝒚)dηϵ(𝒚𝒙)𝒟w𝝂ϕ(𝒙)𝑑𝒙𝑑𝒚\displaystyle=-\int_{\mathbb{R}^{d}}u(\bm{y})\int_{\mathbb{R}^{d}}\eta_{\epsilon}(\bm{y}-\bm{x})\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi}(\bm{x})d\bm{x}d\bm{y}
=du(𝒙)(ηϵ𝒟w𝝂ϕ)(𝒙)𝑑𝒙,\displaystyle=-\int_{\mathbb{R}^{d}}u(\bm{x})(\eta_{\epsilon}*\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi})(\bm{x})d\bm{x},

For the left-hand side, use Fubini’s theorem again to obtain

d(ηϵ𝔊w𝝂u)(𝒙)ϕ(𝒙)𝑑𝒙\displaystyle\int_{\mathbb{R}^{d}}(\eta_{\epsilon}*\mathfrak{G}^{\bm{\nu}}_{w}u)(\bm{x})\cdot\bm{\phi}(\bm{x})d\bm{x} =ddηϵ(𝒙𝒚)𝔊w𝝂u(𝒚)𝑑𝒚ϕ(𝒙)𝑑𝒙\displaystyle=\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}\eta_{\epsilon}(\bm{x}-\bm{y})\mathfrak{G}^{\bm{\nu}}_{w}u(\bm{y})d\bm{y}\cdot\bm{\phi}(\bm{x})d\bm{x}
=d𝔊w𝝂u(𝒚)dηϵ(𝒚𝒙)ϕ(𝒙)𝑑𝒙𝑑𝒚\displaystyle=\int_{\mathbb{R}^{d}}\mathfrak{G}^{\bm{\nu}}_{w}u(\bm{y})\cdot\int_{\mathbb{R}^{d}}\eta_{\epsilon}(\bm{y}-\bm{x})\bm{\phi}(\bm{x})d\bm{x}d\bm{y}
=d𝔊w𝝂u(𝒚)(ηϵϕ)(𝒚)𝑑𝒚\displaystyle=\int_{\mathbb{R}^{d}}\mathfrak{G}^{\bm{\nu}}_{w}u(\bm{y})\cdot(\eta_{\epsilon}*\bm{\phi})(\bm{y})d\bm{y}
=du(𝒚)𝒟w𝝂(ηϵϕ)(𝒚)𝑑𝒚.\displaystyle=-\int_{\mathbb{R}^{d}}u(\bm{y})\mathcal{D}^{-\bm{\nu}}_{w}(\eta_{\epsilon}*\bm{\phi})(\bm{y})d\bm{y}.

One can check that 𝒟w𝝂(ηϵϕ)(𝒙)=(ηϵ𝒟w𝝂ϕ)(𝒙)\mathcal{D}^{-\bm{\nu}}_{w}(\eta_{\epsilon}*\bm{\phi})(\bm{x})=(\eta_{\epsilon}*\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi})(\bm{x}) and therefore

d(ηϵ𝔊w𝝂u)(𝒙)ϕ(𝒙)𝑑𝒙=du(𝒚)(ηϵ𝒟w𝝂ϕ)(𝒚)𝑑𝒚,\int_{\mathbb{R}^{d}}(\eta_{\epsilon}*\mathfrak{G}^{\bm{\nu}}_{w}u)(\bm{x})\bm{\phi}(\bm{x})d\bm{x}=-\int_{\mathbb{R}^{d}}u(\bm{y})(\eta_{\epsilon}*\mathcal{D}^{-\bm{\nu}}_{w}\bm{\phi})(\bm{y})d\bm{y},

Comparing the left-hand and right-hand side, eq. 104 is proved and 𝔊w𝝂(ηϵu)=ηϵ𝔊w𝝂uL2(d;d)\mathfrak{G}^{\bm{\nu}}_{w}(\eta_{\epsilon}*u)=\eta_{\epsilon}*\mathfrak{G}^{\bm{\nu}}_{w}u\in L^{2}(\mathbb{R}^{d};\mathbb{R}^{d}). Therefore

limϵ0𝔊w𝝂(ηϵu)𝔊w𝝂uL2(d;d)=limϵ0ηϵ𝔊w𝝂u𝔊w𝝂uL2(d;d)=0,\lim_{\epsilon\to 0}\|\mathfrak{G}^{\bm{\nu}}_{w}(\eta_{\epsilon}*u)-\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}=\lim_{\epsilon\to 0}\|\eta_{\epsilon}*\mathfrak{G}^{\bm{\nu}}_{w}u-\mathfrak{G}^{\bm{\nu}}_{w}u\|_{L^{2}(\mathbb{R}^{d};\mathbb{R}^{d})}=0,

and thus the lemma is proved.

References

  • [1] R. A. Adams and J. F. Fournier. Sobolev Spaces. Academic Press, San Diego, CA, second edition, 2003.
  • [2] M. Allen, L. Caffarelli, and A. Vasseur. A parabolic problem with a fractional time derivative. Archive for Rational Mechanics and Analysis, 221(2):603–630, 2016.
  • [3] D. Arnold. Finite Element Exterior Calculus. Number 93 in CBMS-NSF Regional Conference Series in Applied Mathematics. SIAM, Philadelphia, PA, 2018.
  • [4] E. Askari, F. Bobaru, R. Lehoucq, M. Parks, S. Silling, and O. Weckner. Peridynamics for multiscale materials modeling. In Journal of Physics: Conference Series, volume 125, page 012078. IOP Publishing, 2008.
  • [5] O. G. Bakunin. Turbulence and Diffusion: Scaling versus Equations. Springer Science & Business Media, 2008.
  • [6] J. Ball and F. Murat. W1,p{W}^{1,p}-quasiconvexity and variational problems for multiple integrals. Journal of Functional Analysis, 58(3):225–253, 1984.
  • [7] L. Bartholdi, T. Schick, N. Smale, and S. Smale. Hodge theory on metric spaces. Foundations of Computational Mathematics, 12(1):1–48, 2012.
  • [8] J. C. Bellido, J. Cueto, and C. Mora-Corral. Non-local gradients in bounded domains motivated by continuum mechanics: Fundamental theorem of calculus and embeddings. Advances in Nonlinear Analysis, 12(1), 2023.
  • [9] D. A. Benson, S. W. Wheatcraft, and M. M. Meerschaert. Application of a fractional advection-dispersion equation. Water resources research, 36(6):1403–1412, 2000.
  • [10] J. Bourgain, H. Brezis, and P. Mironescu. Another look at Sobolev spaces. In Optimal Control and Partial Differential Equation. Conference, Paris , FRANCE (04/12/2000), pages 439–455. IOS Press, Amsterdam, 2001.
  • [11] H. Brezis. Functional Analysis, Sobolev Spaces and Partial Differential Equations. Universitext. Springer, New York, NY, 2011.
  • [12] E. Brué, M. Calzi, G. E. Comi, and G. Stefani. A distributional approach to fractional Sobolev spaces and fractional variation: Asymptotics II. Comptes Rendus. Mathématique, 360(G6):589–626, 2022.
  • [13] A. Buades, B. Coll, and J.-M. Morel. Image denoising methods. A new nonlocal principle. SIAM review, 52(1):113–147, 2010.
  • [14] M. D’Elia, Q. Du, M. Gunzburger, and R. Lehoucq. Nonlocal convection-diffusion problems on bounded domains and finite-range jump processes. Computational Methods in Applied Mathematics, 17(4):707–722, 2017.
  • [15] M. D’Elia, M. Gulian, T. Mengesha, and J. M. Scott. Connections between nonlocal operators: From vector calculus identities to a fractional Helmholtz decomposition. Fractional Calculus and Applied Analysis, pages 1–44, 2022.
  • [16] M. D’Elia, M. Gulian, H. Olson, and G. E. Karniadakis. Towards a unified theory of fractional and nonlocal vector calculus. Fractional Calculus and Applied Analysis, 24(5):1301–1355, 2021.
  • [17] L. Demkowicz, T. Führer, N. Heuer, and X. Tian. The double adaptivity paradigm:(How to circumvent the discrete inf–sup conditions of Babuška and Brezzi). Computers & Mathematics with Applications, 95:41–66, 2021.
  • [18] Q. Du. Nonlocal Modeling, Analysis, and Computation, volume 94. SIAM, 2019.
  • [19] Q. Du, M. Gunzburger, R. B. Lehoucq, and K. Zhou. Analysis and approximation of nonlocal diffusion problems with volume constraints. SIAM review, 54(4):667–696, 2012.
  • [20] Q. Du, M. Gunzburger, R. B. Lehoucq, and K. Zhou. A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws. Mathematical Models and Methods in Applied Sciences, 23(03):493–540, 2013.
  • [21] Q. Du and X. Tian. Stability of nonlocal dirichlet integrals and implications for peridynamic correspondence material modeling. SIAM Journal on Applied Mathematics, 78(3):1536–1552, 2018.
  • [22] Q. Du and X. Tian. Mathematics of smoothed particle hydrodynamics: A study via nonlocal Stokes equations. Foundations of Computational Mathematics, 20:801–826, 2020.
  • [23] Q. Du, X. Tian, and Z. Zhou. Nonlocal diffusion models with consistent local and fractional limits. arXiv preprint arXiv:2203.00167, 2022.
  • [24] L. Evans and R. Gariepy. Measure Theory and Fine Properties of Functions, Revised Edition. Chapman and Hall/CRC, 2015.
  • [25] A. Fiscella, R. Servadei, and E. Valdinoci. Density properties for fractional Sobolev spaces. Ann. Acad. Sci. Fenn. Math, 40(1):235–253, 2015.
  • [26] M. Fuentes, M. Kuperman, and V. Kenkre. Nonlocal interaction effects on pattern formation in population dynamics. Physical review letters, 91(15):158104, 2003.
  • [27] G. F. F. Gounoue. L2-Theory for Nonlocal Operators on Domains. PhD thesis, Bielefeld university, 2020.
  • [28] L. Grafakos. Classical Fourier Analysis. Graduate Texts in Mathematics. Springer New York, 2014.
  • [29] M. Gunzburger and R. B. Lehoucq. A nonlocal vector calculus with application to nonlocal boundary value problems. Multiscale Modeling & Simulation, 8(5):1581–1598, 2010.
  • [30] Y. D. Ha and F. Bobaru. Characteristics of dynamic brittle fracture captured with peridynamics. Engineering Fracture Mechanics, 78(6):1156–1168, 2011.
  • [31] A. Haar and P. Radu. A new nonlocal calculus framework. Helmholtz decompositions, properties, and convergence for nonlocal operators in the limit of the vanishing horizon. Partial Differential Equations and Applications, 3(3):1–20, 2022.
  • [32] A. N. Hirani. Discrete Exterior Calculus. PhD thesis, California Institute of Technology, 2003.
  • [33] M. Kassmann and A. Mimica. Intrinsic scaling properties for nonlocal operators. Journal of the European Mathematical Society, 19(4):983–1011, 2017.
  • [34] A. A. Kilbas, O. Marichev, and S. Samko. Fractional Integrals and Derivatives: Theory and Applications, 1993.
  • [35] J. Klafter and I. M. Sokolov. Anomalous diffusion spreads its wings. Physics world, 18(8):29, 2005.
  • [36] T. D. Le. Nonlocal Exterior Calculus on Riemannian Manifolds. PhD thesis, Pennsylvania State University, 2013.
  • [37] H. Lee and Q. Du. Asymptotically compatible SPH-like particle discretizations of one dimensional linear advection models. SIAM Journal on Numerical Analysis, 57(1):127–147, 2019.
  • [38] H. Lee and Q. Du. Nonlocal gradient operators with a nonspherical interaction neighborhood and their applications. ESAIM: Mathematical Modelling and Numerical Analysis, 54(1):105–128, 2020.
  • [39] Y. Leng, X. Tian, L. Demkowicz, H. Gomez, and J. T. Foster. A Petrov-Galerkin method for nonlocal convection-dominated diffusion problems. Journal of Computational Physics, 452:110919, 2022.
  • [40] M. M. Meerschaert, J. Mortensen, and S. W. Wheatcraft. Fractional vector calculus for fractional advection–dispersion. Physica A: Statistical Mechanics and its Applications, 367:181–190, 2006.
  • [41] T. Mengesha and Q. Du. The bond-based peridynamic system with Dirichlet-type volume constraint. Proceedings of the Royal Society of Edinburgh. Section A. Mathematics, 144(1):161–186, 2014.
  • [42] T. Mengesha and Q. Du. Nonlocal constrained value problems for a linear peridynamic Navier equation. Journal of Elasticity, 116(1):27–51, 2014.
  • [43] T. Mengesha and Q. Du. Characterization of function spaces of vector fields and an application in nonlinear peridynamics. Nonlinear Analysis, 140:82–111, 2016.
  • [44] F. Sabzikar, M. M. Meerschaert, and J. Chen. Tempered fractional calculus. Journal of Computational Physics, 293:14–28, 2015.
  • [45] E. Scalas, R. Gorenflo, and F. Mainardi. Fractional calculus and continuous-time finance. Physica A: Statistical Mechanics and its Applications, 284(1-4):376–384, 2000.
  • [46] T.-T. Shieh and D. E. Spector. On a new class of fractional partial differential equations. Advances in Calculus of Variations, 8(4):321–336, 2015.
  • [47] T.-T. Shieh and D. E. Spector. On a new class of fractional partial differential equations II. Advances in Calculus of Variations, 11(3):289–307, 2018.
  • [48] S. A. Silling. Reformulation of elasticity theory for discontinuities and long-range forces. Journal of the Mechanics and Physics of Solids, 48(1):175–209, 2000.
  • [49] S. A. Silling. Stability of peridynamic correspondence material models and their particle discretizations. Computer Methods in Applied Mechanics and Engineering, 322:42–57, 2017.
  • [50] S. A. Silling, M. Epton, O. Weckner, J. Xu, and E. Askari. Peridynamic states and constitutive modeling. Journal of Elasticity, 88(2):151–184, 2007.
  • [51] P. R. Stinga and M. Vaughan. One-sided fractional derivatives, fractional laplacians, and weighted sobolev spaces. Nonlinear Analysis, 193:111505, 2020.
  • [52] H. Tian, L. Ju, and Q. Du. Nonlocal convection–diffusion problems and finite element approximations. Computer Methods in Applied Mechanics and Engineering, 289:60–78, 2015.
  • [53] H. Tian, L. Ju, and Q. Du. A conservative nonlocal convection–diffusion model and asymptotically compatible finite difference discretization. Computer Methods in Applied Mechanics and Engineering, 320:46–67, 2017.