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

Galerkin Scheme Using Biorthogonal Wavelets on Intervals for 2D Elliptic Interface Problems

Bin Han Department of Mathematical and Statistical Sciences, University of Alberta, Edmonton, Alberta, Canada T6G 2G1. [email protected]  and  Michelle Michelle Department of Mathematics, Purdue University, West Lafayette, IN, USA 47907. [email protected]
Abstract.

This paper introduces a wavelet Galerkin method for solving two-dimensional elliptic interface problems of the form in (au)=f-\nabla\cdot(a\nabla u)=f in Ω\Γ\Omega\backslash\Gamma, where Γ\Gamma is a smooth interface within Ω\Omega. The variable scalar coefficient a>0a>0 and source term ff may exhibit discontinuities across Γ\Gamma. By utilizing a biorthogonal wavelet basis derived from bilinear finite elements, which serves as a Riesz basis for H01(Ω)H^{1}_{0}(\Omega), we devise a strategy that achieves nearly optimal convergence rates: 𝒪(h2|log(h)|2)\mathscr{O}(h^{2}|\log(h)|^{2}) in the L2(Ω)L_{2}(\Omega)-norm and 𝒪(h|log(h)|)\mathscr{O}(h|\log(h)|) in the H1(Ω)H^{1}(\Omega)-norm with respect to the approximation order. To handle the geometry of Γ\Gamma and the singularities of the solution uu, which has a discontinuous gradient across Γ\Gamma, additional wavelet elements are introduced along the interface. The dual part of the biorthogonal wavelet basis plays a crucial role in proving these convergence rates. We develop weighted Bessel properties for wavelets, derive various inequalities in fractional Sobolev spaces, and employ finite element arguments to establish the theoretical convergence results. To achieve higher accuracy and effectively handle high-contrast coefficients aa, our method, much like meshfree approaches, relies on augmenting the number of wavelet elements throughout the domain and near the interface, eliminating the need for re-meshing as in finite element methods. Unlike all other methods for solving elliptic interface problems, the use of a wavelet Riesz basis for H01(Ω)H^{1}_{0}(\Omega) ensures that the condition numbers of the coefficient matrices remain small and uniformly bounded, regardless of the matrix size.

Key words and phrases:
Elliptic interface problems, tensor product wavelets in Sobolev spaces, spline biorthogonal wavelets, Bessel property, fractional Sobolev spaces, meshfree methods
2020 Mathematics Subject Classification:
35J15, 65T60, 42C40, 41A15
Research supported in part by Natural Sciences and Engineering Research Council (NSERC) of Canada under grant RGPIN-2024-04991, NSERC Postdoctoral Fellowship, and the Digital Research Alliance of Canada.

1. Introduction and Motivations

In this paper, we introduce a wavelet Galerkin method for solving 2D elliptic interface problems. Such problems are seen in many applications of science and engineering; for example, the modeling of fluid flow through heterogeneous porous media. Let Γ\Gamma be a smooth curve inside a problem domain Ω\Omega. Then the curve Γ\Gamma splits the domain Ω\Omega into two subregions Ω+\Omega_{+} and Ω\Omega_{-}. For example, the curve Γ\Gamma could be given by {(x,y)Ω:φ(x,y)=0}\{(x,y)\in\Omega:\varphi(x,y)=0\} through a smooth level set function φ\varphi, which splits Ω\Omega into the two subregions Ω+:={(x,y)Ω:φ(x,y)>0}\Omega_{+}:=\{(x,y)\in\Omega:\varphi(x,y)>0\} and Ω:={(x,y)Ω:φ(x,y)<0}\Omega_{-}:=\{(x,y)\in\Omega:\varphi(x,y)<0\}. Throughout the paper, for a function vv in Ω\Omega, we define v+:=vχΩ+v_{+}:=v\chi_{\Omega_{+}}, v:=vχΩv_{-}:=v\chi_{\Omega_{-}}, and

[[v]](x):=v+(x)v(x)=limyΩ+,yxv(y)limzΩ,zxv(z),xΓ,[\![v]\!](x):=v_{+}(x)-v_{-}(x)=\lim_{y\in\Omega_{+},y\to x}v(y)-\lim_{z\in\Omega_{-},z\to x}v(z),\qquad x\in\Gamma,

which is the jump of the function vv across Γ\Gamma, provided that the above jump is well defined.

The elliptic interface problem we consider in this paper is as follows:

(au)=f\displaystyle-\nabla\cdot(a\nabla u)=f in  ΩΓ\Omega\setminus\Gamma, (1.1a)
[[u]]=g\displaystyle[\![u]\!]=g\quad on  Γ\Gamma, (1.1b)
[[aun]]=gΓ\displaystyle[\![a\nabla u\cdot\vec{n}]\!]=g_{\Gamma} on  Γ\Gamma, (1.1c)
u=gb\displaystyle u=g_{b} on  Ω\partial\Omega, (1.1d)

where the variable aL(Ω)a\in L_{\infty}(\Omega) satisfies ess-infx,yΩa(x,y)>0\inf_{x,y\in\Omega}a(x,y)>0, the function fL2(Ω)f\in L_{2}(\Omega) is the source, the boundary function gbH1/2(Ω)g_{b}\in H^{1/2}(\partial\Omega) is given on Ω\partial\Omega, and gH1/2(Γ)g\in H^{1/2}(\Gamma) and gΓH1/2(Γ)g_{\Gamma}\in H^{-1/2}(\Gamma) are for the two jump conditions in (1.1b) and (1.1c). Recall that u±:=uχΩ±u_{\pm}:=u\chi_{\Omega_{\pm}}, a±:=aχΩ±a_{\pm}:=a\chi_{\Omega_{\pm}}, and f±:=fχΩ±f_{\pm}:=f\chi_{\Omega_{\pm}}. Note that [[u]]=g[\![u]\!]=g is the first jump condition (1.1b) for possible discontinuity of the solution uu across Γ\Gamma, while [[aun]]=gΓ[\![a\nabla u\cdot\vec{n}]\!]=g_{\Gamma} is the second jump condition (1.1c) for possible discontinuity of the flux across the interface Γ\Gamma, where n\vec{n} is the unit normal vector of Γ\Gamma pointing into the subregion Ω+\Omega_{+}.

In the context of partial differential equations, one considers the weak solution uu of the model problem (1.1d). Following the standard approach in finite element methods (FEMs), one often assumes g=0g=0 in (1.1b) and gb=0g_{b}=0 in (1.1d), which can be achieved by using auxiliary functions, see Section 3 for details. For the case g=0g=0 on Γ\Gamma, one can observe that the model problem (1.1d) is equivalent to

{(au)=fgΓδΓin Ω,u=gbon Ω,\begin{cases}-\nabla\cdot(a\nabla u)=f-g_{\Gamma}\delta_{\Gamma}&\text{in \quad$\Omega$},\\ u=g_{b}&\text{on \quad$\partial\Omega$},\end{cases} (1.2)

where gΓδΓg_{\Gamma}\delta_{\Gamma} is the Dirac function along the interface Γ\Gamma with weight gΓg_{\Gamma}. Consider the Sobolev space

H01(Ω):={uH1(Ω):u=0 on Ω}.H^{1}_{0}(\Omega):=\{u\in H^{1}(\Omega)\;:\;u=0\text{ on }\partial\Omega\}.

Then the weak formulation of the model problem (1.2) with gb=0g_{b}=0 seeks uH01(Ω)u\in H^{1}_{0}(\Omega) such that

a(u,v):=au,vΩ=f,vΩgΓ,vΓ,vH01(Ω).a(u,v):=\langle a\nabla u,\nabla v\rangle_{\Omega}=\langle f,v\rangle_{\Omega}-\langle g_{\Gamma},v\rangle_{\Gamma},\quad\forall v\in H^{1}_{0}(\Omega). (1.3)

The existence and uniqueness of a weak solution uH1(Ω)u\in H^{1}(\Omega) (or further requiring u+H2(Ω+)u_{+}\in H^{2}(\Omega_{+}) and uH2(Ω)u_{-}\in H^{2}(\Omega_{-})) to the model problem (1.1d) have been extensively addressed in [33, Sections 16 and 17 of Chapter 3]. For the elliptic interface problem (1.1d) with a smooth interface Γ\Gamma, we often assume that the functions aa and ff in (1.1d) are smooth in each subregion Ω±\Omega_{\pm} but could be discontinuous across the interface Γ\Gamma. Though the solution of (1.1d) is known to possess high smoothness away from the interface (i.e., u±u_{\pm} are smooth in each subregion Ω±\Omega_{\pm}), due to the jump conditions in (1.1b)-(1.1c), and due to the discontinuity of aa and ff across Γ\Gamma, the overall smoothness of the solution uu in (1.1d) in the whole domain is very low. For example, if the function gg is not identically zero in (1.1b), then the solution uu is discontinuous in Ω\Omega and has a jump discontinuity across Γ\Gamma. If gΓg_{\Gamma} is not identically zero in (1.1c), then the flux aua\nabla u is discontinuous across Γ\Gamma. Even if gΓ=0g_{\Gamma}=0 but either aa or ff is discontinuous across Γ\Gamma, the gradient u\nabla u must be discontinuous across Γ\Gamma, which produces a solution uu with low regularity. If the standard finite element method (FEM) or finite difference method (FDM) is applied without any modifications, a very low convergence rate is observed. To preserve the optimal convergence with respect to the approximation order used in discretization, various methods have been proposed.

To solve (1.1d) or (1.2), one way is to use the body-fitted FEM with its mesh generated depending on the shape of the interface and the boundary of the domain [9, 12]. This can be challenging especially when the interface has a complicated geometry, and more so for time-dependent problems [1, 41]. There is also a large class of FEMs that do not necessitate a mesh generation that conforms to the interface, which is called unfitted FEMs. Some methods that fall into this category are the immersed FEM (IFEM) [1, 20, 21, 31, 35, 37], the CutFEM [6, 30], the extended FEM (XFEM) [2, 32, 39, 40, 41], and the unfitted hphp method [5, 10, 11, 38]. After fixing a mesh that is independent of the interface, the IFEM proceeds by modifying shape functions of interface elements [21]. As a recent development in the IFEM, a high-order method that addresses nonhomogeneous (first and second) jump conditions and achieves optimal convergence was studied in [1]. Instead of modifying the shape functions near the interface, one can still choose to use the standard FEM shape functions, but employ the Nitsche’s penalty along the interface [21]. This is a key idea of the CutFEM, which was first studied by [30] and then reviewed in [6]. A related method using penalties is the discontinuous Galerkin method for elliptic interface problems [5, 7, 38]. The XFEM incorporates special basis functions near the interface in the approximation space to recover the optimal convergence rate [2, 32, 40, 41]. The shape functions in XFEM are all continuous, which is why this method is deemed to be conforming. Also, unlike other FEM-based methods, no penalties are used by XFEM. The downside is that it may lead to ill-conditioning of the linear system. However, there are further studies that deal with the stabilization for such a method so that the conditioning behaves like the standard FEM [2, 32, 40]. Some studies that assume variable piecewise coefficients aa are [2, 9, 20, 31, 32, 38], whereas the other previously mentioned studies assume piecewise constant coefficients aa.

Various FDMs for solving the model problem (1.1d) have been also studied in the literature [19, 16, 17, 18, 34, 36, 42] and references therein. One way is to use the immersed interface method introduced by [34], whose later developments were discussed in [36]. A key idea of this method is to modify the finite difference stencil that crosses the interface. Another way is to use the matched interface and boundary method [19, 42]. More recently, a sixth-order hybrid FDM for the elliptic interface problem on a rectangular domain with mixed boundary conditions was developed in [18].

Wavelets have been used to solve various differential and integral equations [8, 13, 14, 26, 28] and references therein. The basic idea of a wavelet Galerkin method for solving 2D PDEs (often without singularities) is to use a 2D wavelet basis in H01(Ω)H^{1}_{0}(\Omega). This wavelet basis comprises an affine system generated from a set of functions through scaling and shifting. More specifically, our approximated solution (trial function) takes the form of a linear combination of finitely many functions from this 2D wavelet basis. Traditionally, one often fixes the scale level, which as a result dictates the number of functions/terms in the approximated solution and in fact generates the same space as the FEM. These basis functions, which vary in scales and shifts, are positioned throughout the domain. The coefficients of this approximated solution are then obtained by solving a linear system coming from the weak formulation in (1.3) functions with elements from the same 2D wavelet basis.

1.1. Main contributions of this paper

We introduce a new second-order Galerkin scheme using the tensor product of biorthogonal wavelets on intervals for the model problem in (1.1d). As mentioned earlier, to achieve optimal convergence rates (i.e., those consistent with the approximation order of the scheme), special treatments are required to handle the interface. Our method involves adding extra wavelet elements, which touch the interface and belong to higher scale levels, to our approximated solution. This simple approach conveniently handles the complicated geometry of the interface, effectively captures the singularity along the interface, and easily handles high-contract coefficients aa, thereby enabling us to achieve near optimal convergence rates 𝒪(h2|log(h)|2)\mathscr{O}(h^{2}|\log(h)|^{2}) in the L2(Ω)L_{2}(\Omega)-norm and 𝒪(h|log(h)|)\mathscr{O}(h|\log(h)|) in the H1(Ω)H^{1}(\Omega)-norm, as stated in Theorem 2.2. More specifically, the convergence rates of our method get arbitrarily close to second-order in the L2(Ω)L_{2}(\Omega)-norm and first-order in the H1(Ω)H^{1}(\Omega)-norm as the scale level increases. We establish these near optimal convergence rates by extensively using the dual part of the biorthogonal wavelet basis, relying on the weighted Bessel property and results of wavelets in fractional Sobolev spaces, and employing standard FEM arguments. Not only does our method easily handle the interface geometry, but it can also manage high-contrast coefficients effectively as demonstrated by our numerical experiments later.

It is also worth noting that the added/augmented wavelet elements have scale levels that are at most double the maximum scale level of the other regular basis functions positioned throughout the domain. Consequently, the number of terms used in the approximated solution with these extra functions is only a fixed constant multiple of the number of terms without them. This fixed constant depends on the shape of the interface curve and the support of the wavelet elements. The new linear system corresponding to the coefficients is also a fixed constant multiple of the previous one.

Our method, in a sense, can be interpreted as a meshfree method in that we do not need to generate a mesh that depends on the domain and the interface. To obtain a more accurate solution, there is no need for re-meshing of the entire domain with a smaller mesh size. Instead, to increase accuracy, we raise the scale level of the approximated solution, which entails adding more wavelet elements throughout the domain and additional wavelet elements near the interface. Furthermore, same as the XFEM, our method is considered to be conforming, since each wavelet element is continuous.

Coefficient matrices of many FEMs are known to have condition numbers that are growing proportionally to h2h^{-2}, where hh is the mesh size (e.g., [1, 6, 28]). Our wavelet Galerkin method produces coefficient matrices whose condition numbers are relatively small and uniformly bounded regardless of its size. More precisely, we prove in (2.19) of Theorem 2.2 that the condition numbers are bounded by CwaL(Ω)a1L(Ω)C_{w}\|a\|_{L_{\infty}(\Omega)}\|a^{-1}\|_{L_{\infty}(\Omega)}, where the constant CwC_{w} only depends on the wavelet basis and the domain Ω\Omega, but CwC_{w} is independent of the interface Γ\Gamma. Additionally, the smallest singular values of the coefficient matrices are uniformly bounded away from zero. This is an advantage that we inherit directly from the fact that our 2D wavelet basis is a Riesz basis of H01(Ω)H^{1}_{0}(\Omega). Having such nice condition number properties is beneficial, when an iterative solver is employed to solve the linear system, as it often leads to a relatively small number of iterations required to reach a given tolerance level.

At present, we solely aim to lay the groundwork of our wavelet Galerkin method for solving the 2D elliptic interface problem in (1.1d). Other equally important problems like its high-order version and extensions to the 3D setting are left as a future work, since their implementations and effective calculation of quadratures are much more demanding.

1.2. Organization of this paper

In Section 2.1, we revisit some basic concepts and definitions of wavelets. In Section 2.2, we present the biorthogonal wavelet basis derived from the bilinear function, which we shall use throughout the paper. In Section 2, we formally present our wavelet Galerkin method for the model problem (1.1d) and state our main result in Theorem 2.2 on convergence rates and uniform boundedness of condition numbers. In Section 3, we discuss how we handle nonhomogeneous first jump conditions and/or Dirichlet boundary conditions, and present some numerical experiments to demonstrate the performance of our proposed method. Finally, we present the proofs of our theoretical findings in Theorem 2.2 on convergence rates in Section 4.

2. Wavelet Galerkin method for the model problem (1.1d)

In this section, we describe our Galerkin scheme using the tensor product of biorthogonal wavelets on the unit interval (0,1)(0,1) for solving the 2D elliptic interface problem in (1.1d). As usual in FEMs or traditional wavelet numerical methods, the implementation of our Galerkin scheme only employs the primal part of the biorthogonal wavelets. However, in sharp contrast to FEMs and traditional wavelet numerical methods which critically rely on the polynomial approximation and Bramble-Hilbert lemma, our proof of the nearly optimal convergence rates in Theorem 2.2 of our (nontraditional) wavelet Galerkin scheme extensively and critically takes advantages of the dual part of the biorthogonal wavelets and their weighted Bessel properties in fractional Sobolev spaces. The standard techniques available in the literature are far from sufficient to prove the nearly optimal convergence rates of our wavelet Galerkin scheme, which has to handle the complicated geometry of the interface Γ\Gamma and to capture singularities of the exact solution uu with low regularity near the interface Γ\Gamma.

2.1. Preliminaries on wavelet bases in L2()L_{2}(\mathbb{R}) and L2()L_{2}(\mathcal{I}) with :=(0,1)\mathcal{I}:=(0,1)

Let us first review some basic concepts of wavelets, which follow a similar presentation as in [28]. Let ϕ:={ϕ1,,ϕr}𝖳\phi:=\{\phi^{1},\dots,\phi^{r}\}^{\mathsf{T}} and ψ:={ψ1,,ψs}𝖳\psi:=\{\psi^{1},\dots,\psi^{s}\}^{\mathsf{T}} be square integrable functions in L2()L_{2}(\mathbb{R}). Define a wavelet affine system by

𝖠𝖲J0(ϕ;ψ):={ϕJ0;k:k,=1,,r}{ψj;k:jJ0,k,=1,,s},\operatorname{\mathsf{AS}}_{J_{0}}(\phi;\psi):=\left\{\phi^{\ell}_{J_{0};k}:k\in\mathbb{Z},\ell=1,\ldots,r\right\}\cup\left\{\psi^{\ell}_{j;k}:j\geqslant J_{0},k\in\mathbb{Z},\ell=1,\ldots,s\right\}, (2.1)

where J0J_{0}\in\mathbb{Z}, ϕJ0;k:=2J0/2ϕ(2J0k)\phi^{\ell}_{J_{0};k}:=2^{J_{0}/2}\phi^{\ell}(2^{J_{0}}\cdot-k), and ψj;k:=2j/2ψ(2jk)\psi^{\ell}_{j;k}:=2^{j/2}\psi^{\ell}(2^{j}\cdot-k). We say that 𝖠𝖲J0(ϕ;ψ)\operatorname{\mathsf{AS}}_{J_{0}}(\phi;\psi) is a Riesz basis for L2()L_{2}(\mathbb{R}) if (1) the linear span of 𝖠𝖲J0(ϕ;ψ)\operatorname{\mathsf{AS}}_{J_{0}}(\phi;\psi) is dense in L2()L_{2}(\mathbb{R}), and (2) there exist C1,C2>0C_{1},C_{2}>0 such that

C1η𝖠𝖲J0(ϕ;ψ)|cη|2η𝖠𝖲J0(ϕ;ψ)cηηL2()2C2η𝖠𝖲J0(ϕ;ψ)|cη|2C_{1}\sum_{\eta\in\operatorname{\mathsf{AS}}_{J_{0}}(\phi;\psi)}|c_{\eta}|^{2}\leqslant\Big{\|}\sum_{\eta\in\operatorname{\mathsf{AS}}_{J_{0}}(\phi;\psi)}c_{\eta}\eta\Big{\|}^{2}_{L_{2}(\mathbb{R})}\leqslant C_{2}\sum_{\eta\in\operatorname{\mathsf{AS}}_{J_{0}}(\phi;\psi)}|c_{\eta}|^{2} (2.2)

for all finitely supported sequences {cη}η𝖠𝖲J0(ϕ;ψ)\{c_{\eta}\}_{\eta\in\operatorname{\mathsf{AS}}_{J_{0}}(\phi;\psi)}. The relation in (2.2) holds for some J0J_{0}\in\mathbb{Z} if and only if it holds for all J0J_{0}\in\mathbb{Z} with identical positive constants C1C_{1} and C2C_{2} (see for example [23, Theorem 6]). As a result, we simply refer to {ϕ;ψ}\{\phi;\psi\} as a Riesz multiwavelet in L2()L_{2}(\mathbb{R}) if 𝖠𝖲0(ϕ;ψ)\operatorname{\mathsf{AS}}_{0}(\phi;\psi) is a Riesz basis for L2()L_{2}(\mathbb{R}). Further, let ϕ~:={ϕ~1,,ϕ~r}𝖳\tilde{\phi}:=\{\tilde{\phi}^{1},\dots,\tilde{\phi}^{r}\}^{\mathsf{T}} and ψ~:={ψ~1,,ψ~s}𝖳\tilde{\psi}:=\{\tilde{\psi}^{1},\dots,\tilde{\psi}^{s}\}^{\mathsf{T}} be vectors of square integrable functions in L2()L_{2}(\mathbb{R}). We say that ({ϕ~;ψ~},{ϕ;ψ})(\{\tilde{\phi};\tilde{\psi}\},\{\phi;\psi\}) is a biorthogonal multiwavelet in L2()L_{2}(\mathbb{R}) if (𝖠𝖲0(ϕ~;ψ~),𝖠𝖲0(ϕ;ψ))(\operatorname{\mathsf{AS}}_{0}(\tilde{\phi};\tilde{\psi}),\operatorname{\mathsf{AS}}_{0}(\phi;\psi)) is a biorthogonal basis in L2()L_{2}(\mathbb{R}), i.e., (1) 𝖠𝖲0(ϕ~;ψ~)\operatorname{\mathsf{AS}}_{0}(\tilde{\phi};\tilde{\psi}) and 𝖠𝖲0(ϕ;ψ)\operatorname{\mathsf{AS}}_{0}(\phi;\psi) are Riesz bases in L2()L_{2}(\mathbb{R}), and (2) 𝖠𝖲0(ϕ~;ψ~)\operatorname{\mathsf{AS}}_{0}(\tilde{\phi};\tilde{\psi}) and 𝖠𝖲0(ϕ;ψ)\operatorname{\mathsf{AS}}_{0}(\phi;\psi) are biorthogonal to each other in L2()L_{2}(\mathbb{R}).

The wavelet function ψ\psi has mm vanishing moments if xjψ(x)𝑑x=0\int_{\mathbb{R}}x^{j}\psi(x)dx=0 for all j=0,,m1j=0,\ldots,m-1. By convention, we define vm(ψ):=m\operatorname{vm}(\psi):=m with mm being the largest of such an integer.

The Fourier transform is defined by f^(ξ):=f(x)eixξ𝑑x,ξ\widehat{f}(\xi):=\int_{\mathbb{R}}f(x)e^{-\textsf{i}x\xi}dx,\xi\in\mathbb{R} for fL1()f\in L_{1}(\mathbb{R}) and is naturally extended to square integrable functions in L2()L_{2}(\mathbb{R}). Meanwhile, the Fourier series of u={u(k)}k(l0())r×su=\{u(k)\}_{k\in\mathbb{Z}}\in(l_{0}(\mathbb{Z}))^{r\times s} is defined by u^(ξ):=ku(k)eikξ\widehat{u}(\xi):=\sum_{k\in\mathbb{Z}}u(k)e^{-\textsf{i}k\xi} for ξ\xi\in\mathbb{R}, which is an r×sr\times s matrix of 2π2\pi-periodic trigonometric polynomials. By 𝜹\bm{\delta} we denote the sequence such that 𝜹(0)=1\bm{\delta}(0)=1 and 𝜹(k)=0\bm{\delta}(k)=0 if k0k\neq 0.

Now, we are ready to recall a result of biorthogonal wavelets in L2()L_{2}(\mathbb{R}).

Theorem 2.1.

([24, Theorem 6.4.6] and [23, Theorem 7]) Let ϕ,ϕ~\phi,\tilde{\phi} be r×1r\times 1 vectors of compactly supported distributions on \mathbb{R} and ψ,ψ~\psi,\tilde{\psi} be s×1s\times 1 vectors of compactly supported distributions on \mathbb{R}. Then ({ϕ~;ψ~},{ϕ;ψ})(\{\tilde{\phi};\tilde{\psi}\},\{\phi;\psi\}) is a biorthogonal wavelet in L2()L_{2}(\mathbb{R}) if and only if the following are satisfied

  1. (1)

    ϕ,ϕ~(L2())r\phi,\tilde{\phi}\in(L_{2}(\mathbb{R}))^{r} and ϕ^(0)¯𝖳ϕ~^(0)=1\overline{\widehat{\phi}(0)}^{\mathsf{T}}\widehat{\tilde{\phi}}(0)=1.

  2. (2)

    ϕ\phi and ϕ~\tilde{\phi} are biorthogonal to each other: ϕ~,ϕ(k):=ϕ~(x)ϕ(xk)¯𝖳dx=𝜹(k)Ir\langle\tilde{\phi},\phi(\cdot-k)\rangle:=\int_{\mathbb{R}}\tilde{\phi}(x)\overline{\phi(x-k)}^{\mathsf{T}}dx=\bm{\delta}(k)I_{r} k\forall k\in\mathbb{Z}.

  3. (3)

    There exist low-pass filters a,a~(l0())r×ra,\tilde{a}\in(l_{0}(\mathbb{Z}))^{r\times r} and high-pass filters b,b~(l0())s×rb,\tilde{b}\in(l_{0}(\mathbb{Z}))^{s\times r} such that

    ϕ=2ka(k)ϕ(2k),ψ=2kb(k)ϕ(2k),\displaystyle\phi=2\sum_{k\in\mathbb{Z}}a(k)\phi(2\cdot-k),\qquad\psi=2\sum_{k\in\mathbb{Z}}b(k)\phi(2\cdot-k), (2.3)
    ϕ~=2ka~(k)ϕ~(2k),ψ~=2kb~(k)ϕ~(2k),\displaystyle\tilde{\phi}=2\sum_{k\in\mathbb{Z}}\tilde{a}(k)\tilde{\phi}(2\cdot-k),\qquad\tilde{\psi}=2\sum_{k\in\mathbb{Z}}\tilde{b}(k)\tilde{\phi}(2\cdot-k), (2.4)

    and ({a~;b~},{a;b})(\{\tilde{a};\tilde{b}\},\{a;b\}) is a biorthogonal wavelet filter bank, i.e., s=rs=r and

    [a~^(ξ)a~^(ξ+π)b~^(ξ)b~^(ξ+π)][a^(ξ)¯𝖳b^(ξ)¯𝖳a^(ξ+π)¯𝖳b^(ξ+π)¯𝖳]=I2r,ξ.\left[\begin{matrix}\widehat{\tilde{a}}(\xi)&\widehat{\tilde{a}}(\xi+\pi)\\ \widehat{\tilde{b}}(\xi)&\widehat{\tilde{b}}(\xi+\pi)\end{matrix}\right]\left[\begin{matrix}\overline{\widehat{a}(\xi)}^{\mathsf{T}}&\overline{\widehat{b}(\xi)}^{\mathsf{T}}\\ \overline{\widehat{a}(\xi+\pi)}^{\mathsf{T}}&\overline{\widehat{b}(\xi+\pi)}^{\mathsf{T}}\end{matrix}\right]=I_{2r},\qquad\xi\in\mathbb{R}.
  4. (4)

    Every element in ψ\psi and ψ~\tilde{\psi} has at least one vanishing moment, i.e., ψ(x)𝑑x=ψ~(x)𝑑x=0\int_{\mathbb{R}}\psi(x)dx=\int_{\mathbb{R}}\tilde{\psi}(x)dx=0.

To solve the elliptic interface problem (1.1d), we take the tensor product of wavelets on :=(0,1)\mathcal{I}:=(0,1). Without explicitly involving the dual, the direct approach presented in [27] allows us to construct all possible locally compactly supported biorthogonal wavelets in L2(L_{2}(\mathcal{I}) satisfying prescribed boundary conditions and vanishing moments from any compactly supported biorthogonal (multi)wavelets in L2()L_{2}(\mathbb{R}). That is, our direct approach produces a biorthogonal wavelet (~J01D,J01D)(\tilde{\mathcal{B}}^{1D}_{J_{0}},\mathcal{B}^{1D}_{J_{0}}) in L2(L_{2}(\mathcal{I}), where

J01D:=ΦJ0j=J0ΨjL2(),~J01D:=Φ~J0j=J0Ψ~jL2(),\mathcal{B}^{1D}_{J_{0}}:=\Phi_{J_{0}}\cup\cup_{j=J_{0}}^{\infty}\Psi_{j}\subseteq L_{2}(\mathcal{I}),\qquad\tilde{\mathcal{B}}^{1D}_{J_{0}}:=\tilde{\Phi}_{J_{0}}\cup\cup_{j=J_{0}}^{\infty}\tilde{\Psi}_{j}\subseteq L_{2}(\mathcal{I}),

the integer J0J_{0}\in\mathbb{N} denotes the coarsest scale level, and

ΦJ0:={ϕJ0;0L}{ϕJ0;k:nl,ϕk2J0nh,ϕ}{ϕJ0;2J01R},\displaystyle\Phi_{J_{0}}:=\{\phi^{L}_{J_{0};0}\}\cup\{\phi_{J_{0};k}\;:\;n_{l,\phi}\leqslant k\leqslant 2^{J_{0}}-n_{h,\phi}\}\cup\{\phi^{R}_{J_{0};2^{J_{0}}-1}\},
Ψj:={ψj;0L}{ψj;k:nl,ψk2jnh,ψ}{ψj;2j1R},jJ0,\displaystyle\Psi_{j}:=\{\psi^{L}_{j;0}\}\cup\{\psi_{j;k}\;:\;n_{l,\psi}\leqslant k\leqslant 2^{j}-n_{h,\psi}\}\cup\{\psi^{R}_{j;2^{j}-1}\},\quad j\geqslant J_{0},

with nl,ϕ,nh,ϕ,nl,ψ,nh,ψn_{l,\phi},n_{h,\phi},n_{l,\psi},n_{h,\psi} being known integers, ϕL,ϕR\phi^{L},\phi^{R} being boundary refinable functions, and ψL,ψR\psi^{L},\psi^{R} being boundary wavelets that are finite subsets of functions in L2()L_{2}(\mathcal{I}). Recall that ψj;k:=2j/2ψ(2jk)\psi_{j;k}:=2^{j/2}\psi(2^{j}\cdot-k). We define ~J01D\tilde{\mathcal{B}}^{1D}_{J_{0}} the same way, except we add \sim to each element in J01D\mathcal{B}^{1D}_{J_{0}} for a natural bijection.

2.2. A biorthogonal wavelet basis in H01(Ω)H^{1}_{0}(\Omega) derived from bilinear finite elements

Throughout the paper, for simplicity of presentation, we consider the domain Ω=(0,1)2\Omega=(0,1)^{2}. Though many biorthogonal wavelet bases in H01(Ω)H^{1}_{0}(\Omega) can be used for numerically solving the elliptic interface problems (e.g., [27, Section 7], [28, Section 3.2] and [8, 15]), we shall restrict our attention to one specific biorthogonal wavelet basis on the bounded interval :=(0,1)\mathcal{I}:=(0,1).

Interpolating functions play a critical role in numerical PDEs, wavelet analysis, and computer aided geometric design (e.g., see [25] and references therein). The simplest example of compactly supported interpolating functions is probably the hat function ϕ(x):=max(1|x|,0)\phi(x):=\max(1-|x|,0) for xx\in\mathbb{R}, which is extensively used in numerical PDEs and approximation theory. The hat function ϕ\phi satisfies the refinement equation ϕ=12ϕ(21)+ϕ(2)+12ϕ(2+1)\phi=\frac{1}{2}\phi(2\cdot-1)+\phi(2\cdot)+\frac{1}{2}\phi(2\cdot+1) and ϕ^(0)=1\widehat{\phi}(0)=1. In what follows, we recall a biorthogonal wavelet basis in L2()L_{2}(\mathcal{I}) derived from the hat function ϕ\phi and discussed in [27, Example 7.5], which will be the only biorthogonal wavelet basis used in this paper.

Let ϕ\phi be the hat function. Consider the scalar biorthogonal wavelet ({ϕ~;ψ~},{ϕ;ψ})(\{\tilde{\phi};\tilde{\psi}\},\{\phi;\psi\}) in L2()L_{2}(\mathbb{R}) with ϕ^(0)=ϕ~^(0)=1\widehat{\phi}(0)=\widehat{\tilde{\phi}}(0)=1 and a biorthogonal wavelet filter bank ({a~;b~},{a;b})(\{\tilde{a};\tilde{b}\},\{a;b\}) given by

a=\displaystyle a= {14,12,14}[1,1],b={18,14,34,14,18}[1,3],\displaystyle\left\{\tfrac{1}{4},\tfrac{1}{2},\tfrac{1}{4}\right\}_{[-1,1]},\quad b=\left\{-\tfrac{1}{8},-\tfrac{1}{4},\tfrac{3}{4},-\tfrac{1}{4},-\tfrac{1}{8}\right\}_{[-1,3]}, (2.5)
a~=\displaystyle\tilde{a}= {18,14,34,14,18}[2,2],b~={14,12,14}[0,2].\displaystyle\left\{-\tfrac{1}{8},\tfrac{1}{4},\tfrac{3}{4},\tfrac{1}{4},-\tfrac{1}{8}\right\}_{[-2,2]},\quad\tilde{b}=\left\{-\tfrac{1}{4},\tfrac{1}{2},-\tfrac{1}{4}\right\}_{[0,2]}. (2.6)

In other words, the refinable functions ϕ,ϕ~\phi,\tilde{\phi} and the wavelet functions ψ,ψ~\psi,\tilde{\psi} are determined through the equations in (2.3) and (2.4). Note that the analytic expression of the hat function is ϕ:=(x+1)χ[1,0)+(1x)χ[0,1]\phi:=(x+1)\chi_{[-1,0)}+(1-x)\chi_{[0,1]}. As discussed in [27, Example 7.5], the boundary refinable functions and boundary wavelet functions are defined to be

ψL\displaystyle\psi^{L} =12ϕ(21)ϕ(23)+12ϕ(24),\displaystyle=\tfrac{1}{2}\phi(2\cdot-1)-\phi(2\cdot-3)+\tfrac{1}{2}\phi(2\cdot-4), (2.7)
ϕ~L\displaystyle\tilde{\phi}^{L} =[012132]ϕ~L(2)+[1212]ϕ~(23)+[3214]ϕ~(24)+[120]ϕ~(25)+[140]ϕ~(26),\displaystyle=\begin{bmatrix}0&-\tfrac{1}{2}\\ 1&\tfrac{3}{2}\\ \end{bmatrix}\tilde{\phi}^{L}(2\cdot)+\begin{bmatrix}\tfrac{1}{2}\\ \tfrac{1}{2}\end{bmatrix}\tilde{\phi}(2\cdot-3)+\begin{bmatrix}\tfrac{3}{2}\\ -\tfrac{1}{4}\end{bmatrix}\tilde{\phi}(2\cdot-4)+\begin{bmatrix}\tfrac{1}{2}\\ 0\end{bmatrix}\tilde{\phi}(2\cdot-5)+\begin{bmatrix}-\tfrac{1}{4}\\ 0\end{bmatrix}\tilde{\phi}(2\cdot-6),
ψ~L\displaystyle\tilde{\psi}^{L} =[0112]ϕ~L(2)+[10]ϕ~(23)+[012]ϕ~(24),\displaystyle=\begin{bmatrix}0&-1\\ -1&2\end{bmatrix}\tilde{\phi}^{L}(2\cdot)+\begin{bmatrix}1\\ 0\end{bmatrix}\tilde{\phi}(2\cdot-3)+\begin{bmatrix}0\\ -\tfrac{1}{2}\end{bmatrix}\tilde{\phi}(2\cdot-4),
ψR\displaystyle\psi^{R} =ϕL(1),ϕ~R=ϕ~L(1),andψ~R=ψ~L(1).\displaystyle=\phi^{L}(1-\cdot),\quad\tilde{\phi}^{R}=\tilde{\phi}^{L}(1-\cdot),\quad\text{and}\quad\tilde{\psi}^{R}=\tilde{\psi}^{L}(1-\cdot).

Furthermore, we define

Φj:={ϕj;2,ϕj;1}{ϕj;k:3k2j3}{ϕj;2j2,ϕj;2j1},\displaystyle\Phi_{j}:=\{\phi_{j;2},\phi_{j;1}\}\cup\{\phi_{j;k}:3\leqslant k\leqslant 2^{j}-3\}\cup\{\phi_{j;2^{j}-2},\phi_{j;2^{j}-1}\}, (2.8)
Ψj:={ψj;1,ψj;0L}{ψj;k:2k2j3}{ψj;2j2,ψj;2j1R},\displaystyle\Psi_{j}:=\{\psi_{j;1},\psi^{L}_{j;0}\}\cup\{\psi_{j;k}:2\leqslant k\leqslant 2^{j}-3\}\cup\{\psi_{j;2^{j}-2},\psi^{R}_{j;2^{j}-1}\},
Φ~j:={ϕ~j;0L}{ϕj;k:3k2j3}{ϕ~j;2j1R},\displaystyle\tilde{\Phi}_{j}:=\{\tilde{\phi}^{L}_{j;0}\}\cup\{\phi_{j;k}:3\leqslant k\leqslant 2^{j}-3\}\cup\{\tilde{\phi}^{R}_{j;2^{j}-1}\},
Ψ~j:={ψ~j;0L}{ψj;k:2k2j3}{ψ~j;2j1R},\displaystyle\tilde{\Psi}_{j}:=\{\tilde{\psi}^{L}_{j;0}\}\cup\{\psi_{j;k}:2\leqslant k\leqslant 2^{j}-3\}\cup\{\tilde{\psi}^{R}_{j;2^{j}-1}\},

J01D:=ΦJ0j=J0Ψj\mathcal{B}^{1D}_{J_{0}}:=\Phi_{J_{0}}\cup\cup_{j=J_{0}}^{\infty}\Psi_{j}, and ~J01D:=Φ~J0j=J0Ψ~j\tilde{\mathcal{B}}^{1D}_{J_{0}}:=\tilde{\Phi}_{J_{0}}\cup\cup_{j=J_{0}}^{\infty}\tilde{\Psi}_{j}. Then, (~J01D,J01D)(\tilde{\mathcal{B}}^{1D}_{J_{0}},\mathcal{B}^{1D}_{J_{0}}), where J03J_{0}\geqslant 3, is a biorthogonal wavelet in L2()L_{2}(\mathcal{I}). We shall use the tensor product of this one-dimensional biorthogonal wavelet in L2()L_{2}(\mathcal{I}) throughout this paper. Due to item (3) of Theorem 2.1 and the relations stated in (2.7), there exist well-defined (refinability) matrices Aj,jA_{j,j^{\prime}} and Bj,jB_{j,j^{\prime}} such that

Φj=Aj,jΦjandΨj=Bj,jΦjfor allj<j,\Phi_{j}=A_{j,j^{\prime}}\Phi_{j^{\prime}}\quad\text{and}\quad\Psi_{j}=B_{j,j^{\prime}}\Phi_{j^{\prime}}\quad\text{for all}\quad j<j^{\prime},

which may be convenient to use in the numerical implementation.

We now discuss how to obtain two-dimensional biorthogonal wavelets in L2(Ω)L_{2}(\Omega) with Ω=(0,1)2\Omega=(0,1)^{2} using the tensor product of the one-dimensional biorthogonal wavelet in L2()L_{2}(\mathcal{I}). Given 1D functions f1f_{1} and f2f_{2}, define (f1f2)(x,y):=f1(x)f2(y)(f_{1}\otimes f_{2})(x,y):=f_{1}(x)f_{2}(y) for x,yx,y\in\mathbb{R}. If F1,F2F_{1},F_{2} are sets containing 1D functions, we define F1F2:={f1f2:f1F1,f2F2}F_{1}\otimes F_{2}:=\{f_{1}\otimes f_{2}:f_{1}\in F_{1},f_{2}\in F_{2}\}. Also, define

J02D:=ΦJ02Dj=J0Ψj2D,~J02D:=Φ~J02Dj=J0Ψ~j2D,\mathcal{B}^{2D}_{J_{0}}:=\Phi^{2D}_{J_{0}}\cup\cup_{j=J_{0}}^{\infty}\Psi^{2D}_{j},\qquad\tilde{\mathcal{B}}^{2D}_{J_{0}}:=\tilde{\Phi}^{2D}_{J_{0}}\cup\cup_{j=J_{0}}^{\infty}\tilde{\Psi}^{2D}_{j}, (2.9)

where

ΦJ02D:=ΦJ0ΦJ0,Ψj2D:={ΦjΨj,ΨjΦj,ΨjΨj},\displaystyle\Phi^{2D}_{J_{0}}:=\Phi_{J_{0}}\otimes\Phi_{J_{0}},\quad\Psi^{2D}_{j}:=\{\Phi_{j}\otimes\Psi_{j},\Psi_{j}\otimes\Phi_{j},\Psi_{j}\otimes\Psi_{j}\}, (2.10)
Φ~J02D:=Φ~J0Φ~J0,Ψ~j2D:={Φ~jΨ~j,Ψ~jΦ~j,Ψ~jΨ~j},\displaystyle\tilde{\Phi}^{2D}_{J_{0}}:=\tilde{\Phi}_{J_{0}}\otimes\tilde{\Phi}_{J_{0}},\quad\tilde{\Psi}^{2D}_{j}:=\{\tilde{\Phi}_{j}\otimes\tilde{\Psi}_{j},\tilde{\Psi}_{j}\otimes\tilde{\Phi}_{j},\tilde{\Psi}_{j}\otimes\tilde{\Psi}_{j}\},

where Φj,Ψj,Φ~j,Ψ~j\Phi_{j},\Psi_{j},\tilde{\Phi}_{j},\tilde{\Psi}_{j} are defined as in (2.8), and J03J_{0}\geqslant 3. By using an argument identical to [28, Theorem 1.2], we conclude that (~J02D,J02D)(\tilde{\mathcal{B}}^{2D}_{J_{0}},\mathcal{B}^{2D}_{J_{0}}) is a biorthogonal wavelet in L2(Ω)L_{2}(\Omega) and its properly scaled version defined below

J0H01(Ω):=[2J0ΦJ02D]j=J0[2jΨj2D]\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}}:=[2^{-J_{0}}\Phi^{2D}_{J_{0}}]\cup\cup_{j=J_{0}}^{\infty}[2^{-j}\Psi^{2D}_{j}] (2.11)

is a Riesz basis of the Sobolev space H01(Ω)H^{1}_{0}(\Omega), see [28, Theorem 1.2]. That is, (1) the linear span of J0H01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}} is dense in H01(Ω)H^{1}_{0}(\Omega), and (2) there exist positive constants C,1,C,2>0C_{\mathcal{B},1},C_{\mathcal{B},2}>0 such that

C,1ηJ0H01(Ω)|cη|2ηJ0H01(Ω)cηηH1(Ω)2C,2ηJ0H01(Ω)|cη|2C_{\mathcal{B},1}\sum_{\eta\in\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}}}|c_{\eta}|^{2}\leqslant\Big{\|}\sum_{\eta\in\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}}}c_{\eta}\eta\Big{\|}^{2}_{H^{1}(\Omega)}\leqslant C_{\mathcal{B},2}\sum_{\eta\in\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}}}|c_{\eta}|^{2}

for all finitely supported sequences {cη}ηJ0H01(Ω)\{c_{\eta}\}_{\eta\in\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}}}.

2.3. Methodology for solving the model problem (1.1d)

We now describe our proposed method. As mentioned earlier, we shall always use the biorthogonal wavelet basis presented in Section 2.2.

For J0=3J_{0}=3 and JJ0J\geqslant J_{0}, we define the traditional finite-dimensional wavelet element space truncated at the scale level JJ as follows:

J0,J2D:=ΦJ02Dj=J0J1Ψj2DandJ0,JH01(Ω):=[2J0ΦJ02D]j=J0J1[2jΨj2D].\mathcal{B}^{2D}_{J_{0},J}:=\Phi^{2D}_{J_{0}}\cup\cup_{j=J_{0}}^{J-1}\Psi^{2D}_{j}\quad\mbox{and}\quad\quad\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0},J}:=[2^{-J_{0}}\Phi^{2D}_{J_{0}}]\cup\cup_{j=J_{0}}^{J-1}[2^{-j}\Psi^{2D}_{j}]. (2.12)

Obviously, J0,JH01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0},J} is a finite subset of J0H01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}}, where J0H01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}} is defined as in (2.11). Using a uniform grid, the standard FEM only uses the basis ΦJ2D\Phi^{2D}_{J} and its finite element space VJ:=span(ΦJ2D)V_{J}:=\mbox{span}(\Phi^{2D}_{J}). It is very important to notice that span(J0,J2D)=VJ\mbox{span}(\mathcal{B}^{2D}_{J_{0},J})=V_{J} and span(J0,JH01(Ω))=VJ\mbox{span}(\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0},J})=V_{J}. In other words, both J0,JH01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0},J} (or J0,J2D\mathcal{B}^{2D}_{J_{0},J}) and ΦJ2D\Phi^{2D}_{J} span the same (finite element) space VJV_{J}. The numerical solution to (1.1d) obtained by the traditional wavelet Galerkin method using only J0,J2D\mathcal{B}^{2D}_{J_{0},J} is the same as the solution obtained by using ΦJ2D\Phi^{2D}_{J} (the standard bilinear FEM). In the context of the model problem (1.1d), using only the traditional wavelet basis J0,JH01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0},J} inevitably suffers from the same convergence issue faced in the standard FEM. More specifically, the observed convergence rate will typically be well below two because the exact solution uH2(Ω)u\not\in H^{2}(\Omega) and has discontinuous gradients along the interface Γ\Gamma.

To overcome such an issue, we propose incorporating of higher-resolution wavelets defined below

𝒮j:={ηΨj2D:supp(η~)Γandη~Ψ~j2D},j\mathcal{S}_{j}:=\{\eta\in\Psi^{2D}_{j}:\operatorname{supp}(\tilde{\eta})\cap\Gamma\neq\emptyset\;\text{and}\;\tilde{\eta}\in\tilde{\Psi}^{2D}_{j}\},\qquad j\in\mathbb{N} (2.13)

to capture the geometry of the interface Γ\Gamma and the singularity of the solution uu along the interface Γ\Gamma on top of the standard wavelet elements J0,J2D\mathcal{B}^{2D}_{J_{0},J}. More specifically, we shall use

J0,JS:=J0,J2Dj=J2J2𝒮j,or equivalently,J0,JS,H01(Ω):=J0,JH01(Ω)j=J2J2[2j𝒮j],\mathcal{B}^{S}_{J_{0},J}:=\mathcal{B}^{2D}_{J_{0},J}\cup\cup_{j=J}^{2J-2}\mathcal{S}_{j},\quad\mbox{or equivalently},\quad\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}:=\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0},J}\cup\cup_{j=J}^{2J-2}[2^{-j}\mathcal{S}_{j}], (2.14)

where the superscript SS indicates that we add extra wavelet elements Sj,j=J,,2J2S_{j},j=J,\ldots,2J-2 to the traditional wavelet basis J0,J2D\mathcal{B}^{2D}_{J_{0},J}. Note that the cardinality of J0,J2D\mathcal{B}^{2D}_{J_{0},J} is 𝒪(h2)\mathscr{O}(h^{-2}), where h:=2Jh:=2^{-J} is the mesh size. Because Γ\Gamma is a 1D closed curve inside Ω\Omega, it is not difficult to observe that the cardinality of j=J2J2𝒮j\cup_{j=J}^{2J-2}\mathcal{S}_{j} is also 𝒪(h2)\mathscr{O}(h^{-2}) (with the prefactor being independent of JJ). To put differently, the cardinality of J0,JS\mathcal{B}^{S}_{J_{0},J} is still 𝒪(h2)\mathscr{O}(h^{-2}), which means that it is comparable to that of the original bases J0,J2D\mathcal{B}^{2D}_{J_{0},J} and ΦJ2D\Phi^{2D}_{J}.

By the weak formulation in (1.3) with g=0g=0 and gb=0g_{b}=0 and considering the approximated function uJ:=ηJ0,JS,H01(Ω)cηηu_{J}:=\sum_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}c_{\eta}\eta with to-be-determined unknown coefficients {cη}ηJ0,JS,H01(Ω)\{c_{\eta}\}_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}} (we shall also define uh:=uJu_{h}:=u_{J} with h:=2Jh:=2^{-J}), our wavelet Galerkin method reduces to finding all the coefficients cηc_{\eta} for ηJ0,JH1(Ω)\eta\in\mathcal{B}^{H^{1}(\Omega)}_{J_{0},J} such that

a(uJ,v):=auJ,vΩ=f,vΩgΓ,vΓ,vJ0,JS,H01(Ω).a(u_{J},v):=\langle a\nabla u_{J},\nabla v\rangle_{\Omega}=\langle f,v\rangle_{\Omega}-\langle g_{\Gamma},v\rangle_{\Gamma},\quad\forall v\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}. (2.15)

Fig. 1 visualizes the basis functions used in our approximated solution uJu_{J}. Due to the symmetry in the biorthogonal wavelet basis in Section 2.2, each term of the approximated solution can be obtained by scaling, shifting, and rotating one of the functions in panels (c)-(h) of Fig. 1.

Refer to caption
(a) ϕ\phi
Refer to caption
(b) ψ\psi (black), ψL\psi^{L} (red)
Refer to caption
(c) ϕϕ\phi\otimes\phi
Refer to caption
(d) ϕψL\phi\otimes\psi^{L}
Refer to caption
(e) ϕψ\phi\otimes\psi
Refer to caption
(f) ψLψL\psi^{L}\otimes\psi^{L}
Refer to caption
(g) ψLψ\psi^{L}\otimes\psi
Refer to caption
(h) ψψ\psi\otimes\psi
Figure 1. Panels (a)-(b) depict generators of the 1D wavelet basis J01D\mathcal{B}^{1D}_{J_{0}} with J0=3J_{0}=3. Panels (c)-(h) depict generators of the 2D wavelet basis J02D\mathcal{B}^{2D}_{J_{0}} with J0=3J_{0}=3.

Meanwhile, Fig. 2 visualizes the overlapping supports of wavelet basis functions in 3,4S,H01(Ω)j=46[2j𝒮j]\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,4}\cup\cup_{j=4}^{6}[2^{-j}\mathcal{S}_{j}], and gives us an insight as to how we add the wavelets along the interface in our approximated solution uJu_{J}.

Without loss of generality, we assume that Ω:=(0,1)2\Omega:=(0,1)^{2} is our domain. The next theorem is our main theoretical result on the convergence order of our proposed wavelet Galerkin method using J0,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J} in (2.14) as JJ\to\infty, and the uniform boundedness of the condition numbers of its coefficient matrices. We shall assume that g=0g=0 on Γ\Gamma in the first jump condition (1.1b) for avoiding discontinuous uu, and gb=0g_{b}=0 in Ω\partial\Omega for the homogeneous Dirichlet condition in accordance to the standard FEM argument. Since J0,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J} is a finite subset of the Riesz wavelet basis J0H01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}} in H01(Ω)H^{1}_{0}(\Omega), the condition numbers of coefficient matrices from (2.15) are uniformly bounded and independent of the mesh size hh and resolution level JJ. Due to the technicality, we defer the proof and its auxiliary results to Section 4.

Theorem 2.2.

Under the standard assumptions g=0g=0 and gb=0g_{b}=0 in finite element methods, let uH01(Ω)u\in H^{1}_{0}(\Omega) be the exact solution of the model problem (1.1d) with variable functions a,f,gΓa,f,g_{\Gamma} such that

u+:=uχΩ+H2(Ω+) and u:=uχΩH2(Ω).u_{+}:=u\chi_{\Omega_{+}}\in H^{2}(\Omega_{+})\quad\mbox{ and }\quad u_{-}:=u\chi_{\Omega_{-}}\in H^{2}(\Omega_{-}). (2.16)

We assume that the interface Γ\Gamma is of class 𝒞2\mathscr{C}^{2}. For JJ0J\geqslant J_{0}, define h:=2Jh:=2^{-J} as the mesh size and NJN_{J} as the cardinality of the set of the basis J0,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}. Define Vhwav:=span(J0,JS,H01(Ω))V_{h}^{wav}:=\mbox{span}(\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}). Let uh=uJ:=ηJ0,JS,H01(Ω)cηηVhwavu_{h}=u_{J}:=\sum_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}c_{\eta}\eta\in V_{h}^{wav} be the numerical solution obtained from (2.15) (i.e., the weak formulation of (1.1d) in the wavelet subspace VhwavV_{h}^{wav}) by using the basis J0,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J} in (2.14). Then for all JJ0J\geqslant J_{0}, there exists a positive constant CC, independent of all JJ, hh and NJN_{J}, such that

uhuH1(Ω)Ch|log(h)|,uJuH1(Ω)CNJ1/2J,\|u_{h}-u\|_{H^{1}(\Omega)}\leqslant Ch|\log(h)|,\qquad\|u_{J}-u\|_{H^{1}(\Omega)}\leqslant CN_{J}^{-1/2}J, (2.17)

and

uhuL2(Ω)Ch2|log(h)|2,uJuL2(Ω)CNJ1J2,\|u_{h}-u\|_{L_{2}(\Omega)}\leqslant Ch^{2}|\log(h)|^{2},\qquad\|u_{J}-u\|_{L_{2}(\Omega)}\leqslant CN_{J}^{-1}J^{2}, (2.18)

where log()\log(\cdot) is the natural logarithm and in fact, the above generic constant CC in (2.17) and (2.18) is bounded by c(u+H2(Ω+)2+uH2(Ω)2)1/2c(\|u_{+}\|^{2}_{H^{2}(\Omega_{+})}+\|u_{-}\|^{2}_{H^{2}(\Omega_{-})})^{1/2} with a positive constant cc only depending on the domain Ω\Omega, the interface Γ\Gamma and the wavelet basis. Moreover, the condition number must satisfy

κ([a(α,β)]α,βJ0,JS,H01(Ω))CwaL(Ω)a1L(Ω),for all JJ0,\kappa\Big{(}[a(\alpha,\beta)]_{\alpha,\beta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}\Big{)}\leqslant C_{w}\|a\|_{L_{\infty}(\Omega)}\|a^{-1}\|_{L_{\infty}(\Omega)},\quad\text{for all }J\geqslant J_{0}, (2.19)

where κ\kappa denotes the condition number of the coefficient matrix and CwC_{w} is a positive constant that only depends on the wavelet basis and the domain Ω\Omega, but CwC_{w} is independent of the interface Γ\Gamma.

We shall prove Theorem 2.2 under the abstract assumption (2.16) on u+u_{+} and uu_{-}, which can be satisfied by specifying concrete conditions on variable functions a,fa,f and gΓg_{\Gamma}. For example, according to [33, Theorem 10.1 and Section 16], the assumption (2.16) on uu is satisfied if a+:=aχΩ+C1(Ω+¯)a_{+}:=a\chi_{\Omega_{+}}\in C^{1}(\overline{\Omega_{+}}), a:=aχΩC1(Ω¯)a_{-}:=a\chi_{\Omega_{-}}\in C^{1}(\overline{\Omega_{-}}), fL2(Ω)f\in L_{2}(\Omega), and gΓH1/2(Γ)g_{\Gamma}\in H^{1/2}(\Gamma). Of course, we assume ess-infx,yΩa(x,y)>0\inf_{x,y\in\Omega}a(x,y)>0 but the variable functions aL(Ω)a\in L_{\infty}(\Omega) and fL2(Ω)f\in L_{2}(\Omega) could be discontinuous across the interface Γ\Gamma. It is also important to notice that NJN_{J}, the cardinality of the set J0,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}, satisfies h2NJCΓh2h^{-2}\leqslant N_{J}\leqslant C_{\Gamma}h^{-2} with h:=2Jh:=2^{-J} for a positive constant CΓC_{\Gamma} only depending on the interface curve Γ\Gamma, in particular, the length of Γ\Gamma. Our proof of Theorem 2.2 extensively uses the dual part of the biorthogonal wavelet basis and relies on the weighted Bessel properties and results of wavelets in fractional Sobolev spaces, plus standard FEM arguments and various inequalities in fractional Sobolev spaces.

Refer to caption
(a) 3,42D\mathcal{B}^{2D}_{3,4}
Refer to caption
(b) 𝒮4\mathcal{S}_{4}
Refer to caption
(c) 𝒮5\mathcal{S}_{5}
Refer to caption
(d) 𝒮6\mathcal{S}_{6}
Figure 2. For simplicity, we assume that the interface curve, Γ\Gamma, is a circle. Panel (a) depicts the overlapping supports of wavelets in 3,42D\mathcal{B}^{2D}_{3,4}. Panels (b)-(d) depict the overlapping supports of extra wavelets added along the interface Γ\Gamma, which make up the set j=46[2j𝒮j]\cup_{j=4}^{6}[2^{-j}\mathcal{S}_{j}].

3. Numerical Experiments

In this section, we present some numerical experiments to demonstrate the performance of our wavelet Galerkin method. In each table below, JJ corresponds to the scale level in (2.12) with the coarsest scale level J0=3J_{0}=3. Additionally, NJN_{J} stands for the number of terms (freedom) at the scale level JJ used in the approximated solution, which is equal to the cardinality of 3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} as defined in (2.14). If the exact solution uu is known, the quantities under ‘order’ (for L2(Ω)L_{2}(\Omega) convergence) are computed as follows

order=2log2(uuJ12uuJ2)(log2(NJNJ1))1.\text{order}=2\log_{2}\left(\frac{\|u-u_{J-1}\|_{2}}{\|u-u_{J}\|_{2}}\right)\left(\log_{2}\left(\frac{N_{J}}{N_{J-1}}\right)\right)^{-1}. (3.1)

On the other hand, if the exact solution uu is unknown, we replace the error uuJ1u-u_{J-1} with uJuJ1u_{J}-u_{J-1} and the error uuJu-u_{J} with uJ+1uJu_{J+1}-u_{J} in the above formula. The convergence in terms of H1(Ω)H^{1}(\Omega) semi-norm is similarly calculated, except we replace the solution with its gradient. To approximate the L2(Ω)L_{2}(\Omega) error, we compute all errors using the l2l_{2} norm on a fine grid of size 2132^{-13} in each direction. The condition number κ\kappa of the coefficient matrix of size NJ×NJN_{J}\times N_{J} is calculated by dividing its largest singular value with its smallest singular value. For each example, we compare the errors and the convergence rates of the approximated solution formed by 3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} with the one formed by the traditional wavelet method using 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} only. Because both 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} and ΦJ2D\Phi^{2D}_{J} span the same finite element space VJ=span(ΦJ2D)V_{J}=\mbox{span}(\Phi^{2D}_{J}), the numerical solutions obtained by the traditional wavelet method using 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} and the FEM ΦJ2D\Phi^{2D}_{J} are the same.

3.1. Handling nonhomogeneous first jump and/or Dirichlet boundary conditions

Some of the following examples have nonhomogeneous first jump condition and/or Dirichlet boundary condition. To handle them, we shall exploit the geometry of our interface curve and unit square domain with (1/2,1/2)(1/2,1/2) as its center. For the sake of discussion, we assume that the first jump condition is parameterized in terms of angle and Ω\Omega_{-} is away from Ω\partial\Omega. Since the interface curve is smooth, we are able to radially extend the first jump condition outward and treat its restriction in Ω+\Omega_{+} as an auxiliary solution. More specifically, for (x,y)2Ω(x,y)\in\mathbb{R}^{2}\setminus\Omega_{-}, we define

g~(x,y):=g(Θ(x,y)),Θ(x,y):={arctan(y1/2x1/2),ifx>1/2,arctan(y1/2x1/2)+π,ifx<1/2,y1/2,arctan(y1/2x1/2)π,ifx<1/2,y<1/2,π/2,ifx=1/2,y>1/2,π/2,ifx=1/2,y<1/2.\tilde{g}(x,y):=g\left(\Theta(x,y)\right),\quad\Theta(x,y):=\begin{cases}\arctan\left(\tfrac{y-1/2}{x-1/2}\right),&\text{if}\;x>1/2,\\ \arctan\left(\tfrac{y-1/2}{x-1/2}\right)+\pi,&\text{if}\;x<1/2,\;y\geqslant 1/2,\\ \arctan\left(\tfrac{y-1/2}{x-1/2}\right)-\pi,&\text{if}\;x<1/2,\;y<1/2,\\ \pi/2,&\text{if}\;x=1/2,\;y>1/2,\\ -\pi/2,&\text{if}\;x=1/2,\;y<1/2.\end{cases} (3.2)

To handle this nonhomogeneous Dirichlet boundary condition, we build two more auxiliary solutions

u~LR\displaystyle\tilde{u}_{LR} :=(gb(0,y)g~(0,y))(1x)+(gb(1,y)g~(1,y))x,\displaystyle:=(g_{b}(0,y)-\tilde{g}(0,y))(1-x)+(g_{b}(1,y)-\tilde{g}(1,y))x,
u~BT\displaystyle\tilde{u}_{BT} :=(gb(x,0)g~(x,0)u~LR(x,0))(1y)+(gb(x,1)g~(x,1)u~LR(x,1))y.\displaystyle:=(g_{b}(x,0)-\tilde{g}(x,0)-\tilde{u}_{LR}(x,0))(1-y)+(g_{b}(x,1)-\tilde{g}(x,1)-\tilde{u}_{LR}(x,1))y.

Define the function GG such that

G+=GχΩ+:=g~χΩ++u~LR+u~BTandG=GχΩ:=u~LR+u~BT.G_{+}=G\chi_{\Omega_{+}}:=\tilde{g}\chi_{\Omega_{+}}+\tilde{u}_{LR}+\tilde{u}_{BT}\quad\text{and}\quad G_{-}=G\chi_{\Omega_{-}}:=\tilde{u}_{LR}+\tilde{u}_{BT}.

Next, we aim to find u~J:=ηJ0,JS,H01(Ω)cηη\tilde{u}_{J}:=\sum_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}c_{\eta}\eta such that

au~J,vΩ=f,vΩgΓ,vΓaG,v,vJ0,JS,H01(Ω).\langle a\nabla\tilde{u}_{J},\nabla v\rangle_{\Omega}=\langle f,v\rangle_{\Omega}-\langle g_{\Gamma},v\rangle_{\Gamma}-\langle a\nabla G,\nabla v\rangle,\quad\forall v\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}.

Finally, we define our approximated solution as uJ:=u~J+g~χΩ++u~LR+u~BTu_{J}:=\tilde{u}_{J}+\tilde{g}\chi_{\Omega_{+}}+\tilde{u}_{LR}+\tilde{u}_{BT}.

3.2. Examples with known exact solutions

We present four examples here, where the exact solutions are known. Theorem 2.2 guarantees that the condition numbers satisfy κCwaL(Ω)a1L(Ω)\kappa\leqslant C_{w}\|a\|_{L_{\infty}(\Omega)}\|a^{-1}\|_{L_{\infty}(\Omega)} in (2.19). In all the numerical examples, we indicate the numerically estimated constant CwC_{w} in (2.19).

Example 3.1.

Consider the model problem (1.1d), where a+{102,106}a_{+}\in\{10^{2},10^{6}\}, a=1a_{-}=1,

Γ={(x,y)Ω:x(θ)=14cos(θ)+12,y(θ)=14cos(θ)+12,θ[0,2π)},\Gamma=\{(x,y)\in\Omega\;:\;x(\theta)=\tfrac{1}{4}\cos(\theta)+\tfrac{1}{2},\;y(\theta)=\tfrac{1}{4}\cos(\theta)+\tfrac{1}{2},\;\theta\in[0,2\pi)\},

and f,g,gΓf,g,g_{\Gamma} are chosen such that the exact solution, uu, is

u+=a+1((x12)2+(y12)2)3/2+26(a1a+1)andu=a1((x12)2+(y12)2)3/2.u_{+}=a_{+}^{-1}((x-\tfrac{1}{2})^{2}+(y-\tfrac{1}{2})^{2})^{3/2}+2^{-6}\left(a_{-}^{-1}-a_{+}^{-1}\right)\quad\text{and}\quad u_{-}=a_{-}^{-1}((x-\tfrac{1}{2})^{2}+(y-\tfrac{1}{2})^{2})^{3/2}.

This makes g=gΓ=0g=g_{\Gamma}=0 on Γ\Gamma, and the Dirichlet boundary condition, gbg_{b}, nonzero on Ω\partial\Omega.

a+=102a_{+}=10^{2}, a=1a_{-}=1
3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (traditional) or ΦJ2D\Phi^{2D}_{J} (FEM)
JJ NJN_{J} κ\kappa uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order NJN_{J} uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order
4 2345 3.52E+2 8.57E-2 3.06E-1 225 5.54E-1 7.44E-1
5 10401 4.66E+2 2.30E-2 1.76 1.59E-2 0.878 961 3.20E-1 0.755 5.85E-1 0.332
6 43449 5.67E+2 6.14E-3 1.85 8.10E-2 0.944 3969 1.60E-1 0.980 4.15E-1 0.482
7 177169 6.42E+2 1.60E-3 1.91 3.79E-2 1.08 16129 8.23E-2 0.946 3.01E-1 0.458
a+=106,a=1a_{+}=10^{6},a_{-}=1
3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (traditional) or ΦJ2D\Phi^{2D}_{J} (FEM)
JJ NJN_{J} κ\kappa uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order NJN_{J} uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order
4 2345 5.83E+6 1.64E-1 3.98E-1 225 7.11E-1 8.97E-1
5 10401 8.94E+6 3.75E-2 1.98 2.01E-1 0.917 961 4.24E-1 0.711 7.14E-1 0.315
6 43449 1.24E+7 8.38E-3 2.10 1.00E-1 0.970 3969 2.30E-1 0.863 5.42E-1 0.388
7 177169 1.56E+7 2.35E-3 1.81 5.16E-2 0.947 16129 1.06E-1 1.10 3.71E-1 0.539
Table 1. Numerical results for Example 3.1. The estimated constant CwC_{w} in (2.19) is less than 1616.
3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) ΦJ2D\Phi^{2D}_{J} (FEM)
JJ 4 5 6 7 6 7 8 9
NJN_{J} 2345 10401 43449 177169 3969 16129 65025 261121
#\# of iterations 121 166 189 207 334 689 1382 2639
Table 2. The number of GMRES iterations required to reach the tolerance level of 10810^{-8} for Example 3.1 with a+=102a_{+}=10^{2} and a=1a_{-}=1.

See Tables 1 and 2 for numerical results, and Fig. 3 for plots. This example aims to show that the high contrast in the diffusion coefficient aa results in large condition numbers, but they are still uniformly bounded, which is consistent with our main result Theorem 2.2. If we compare the degrees of freedom of 3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} and 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (or equivalently ΦJ2D\Phi^{2D}_{J}) at each scale level, we observe that the former is only a fixed constant multiple of the latter for all scale levels and this constant is independent of the scale level. Table 2 demonstrates that the number of GMRES iterations required to reach the tolerance level 10810^{-8} is smaller compared to the standard FEM case and is uniformly bounded irrespective of the matrix size. This is due to fact that the wavelet coefficient matrices have small condition numbers that are uniformly bounded. On the other hand, in the case of standard FEM, the number of GMRES iterations required to reach a tolerance level of 10810^{-8} doubles with each increase in the scale level.

\begin{overpic}[width=108.405pt]{circle_gamma.pdf} \put(23.0,68.0){$a_{+}\in\{10^{2},10^{6}\}$} \put(40.0,38.0){$a_{-}=1$} \end{overpic}
Refer to caption
Refer to caption
Figure 3. Example 3.1. Left: the plot of Γ\Gamma. Middle: the plot of the approximated solution at J=7J=7, where a+=106a_{+}=10^{6}. Right: the plot of the error at J=7J=7, where a+=106a_{+}=10^{6}.
Example 3.2.

Consider the model problem (1.1d), where a+=2+sin(5(x1/2))sin(5(y1/2))a_{+}=2+\sin(5(x-1/2))\sin(5(y-1/2)), a=103a+a_{-}=10^{3}a_{+},

Γ={(x,y)Ω:x(θ)=(15+225sin(5θ))cos(θ)+12,y(θ)=(15+225sin(5θ))sin(θ)+12,θ[0,2π)},\Gamma=\{(x,y)\in\Omega\;:\;x(\theta)=(\tfrac{1}{5}+\tfrac{2}{25}\sin(5\theta))\cos(\theta)+\tfrac{1}{2},\;y(\theta)=(\tfrac{1}{5}+\tfrac{2}{25}\sin(5\theta))\sin(\theta)+\tfrac{1}{2},\;\theta\in[0,2\pi)\},

and f,g,gΓf,g,g_{\Gamma} are chosen such that the exact solution, uu, is

u+\displaystyle u_{+} =sin(10x5)sin(10y5)((x12)2+(y12)2(15+225sin(5Θ(x,y)))2)+1,\displaystyle=\sin(10x-5)\sin(10y-5)\left((x-\tfrac{1}{2})^{2}+(y-\tfrac{1}{2})^{2}-(\tfrac{1}{5}+\tfrac{2}{25}\sin(5\Theta(x,y)))^{2}\right)+1,
u\displaystyle u_{-} =103sin(10x5)sin(10y5)((x12)2+(y12)2(15+225sin(5Θ(x,y)))2)+31,\displaystyle=10^{-3}\sin(10x-5)\sin(10y-5)\left((x-\tfrac{1}{2})^{2}+(y-\tfrac{1}{2})^{2}-(\tfrac{1}{5}+\tfrac{2}{25}\sin(5\Theta(x,y)))^{2}\right)+31,

where Θ\Theta is defined as in (3.2) for x,yΩx,y\in\Omega and Θ(1/2,1/2):=0\Theta(1/2,1/2):=0. This makes g0g\neq 0, gΓ=0g_{\Gamma}=0 on Γ\Gamma, and the Dirichlet boundary condition, gbg_{b}, nonzero on Ω\partial\Omega. See Table 3 for numerical results and Fig. 4 for plots.

3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (traditional) or ΦJ2D\Phi^{2D}_{J} (FEM)
JJ NJN_{J} κ\kappa uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order NJN_{J} uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order
4 2833 3.07E+4 3.85E-2 2.03E-1 225 5.68E-2 2.34E-1
5 13589 4.04E+4 1.04E-2 1.68 1.08E-1 0.810 961 2.50E-2 1.13 1.26E-1 0.857
6 57317 4.50E+4 2.70E-3 1.87 5.55E-2 0.929 3969 1.38E-2 0.841 7.35E-2 0.755
7 233583 4.54E+4 6.88E-4 1.95 2.75E-2 1.00 16129 7.68E-3 0.833 4.58E-2 0.677
Table 3. Numerical results for Example 3.2. The estimated constant CwC_{w} in (2.19) is less than 1616.
\begin{overpic}[width=108.405pt]{flower5_gamma.pdf} \put(34.0,42.0){\tiny$a_{-}=10^{3}a_{+}$} \put(10.0,73.0){\tiny$a_{+}=$} \put(10.0,66.0){\tiny$2+\sin(5(x-\tfrac{1}{2}))\sin(5(y-\tfrac{1}{2}))$} \end{overpic}
Refer to caption
Refer to caption
Figure 4. Example 3.2. Left: the plot of Γ\Gamma. Middle: the plot of the approximated solution at J=7J=7. Right: the plot of the error at J=7J=7.
Example 3.3.

Consider the model problem (1.1d), where a+=1a_{+}=1, a=103a_{-}=10^{-3},

Γ={(x,y)Ω:x(θ)=14(π3+25sin(8θ))cos(θ)+12,y(θ)=14(π3+25sin(8θ))sin(θ)+12,θ[0,2π)},\Gamma=\{(x,y)\in\Omega\;:\;x(\theta)=\tfrac{1}{4}(\tfrac{\pi}{3}+\tfrac{2}{5}\sin(8\theta))\cos(\theta)+\tfrac{1}{2},\;y(\theta)=\tfrac{1}{4}(\tfrac{\pi}{3}+\tfrac{2}{5}\sin(8\theta))\sin(\theta)+\tfrac{1}{2},\;\theta\in[0,2\pi)\},

and f,g,gΓf,g,g_{\Gamma} are chosen such that the exact solution, uu, is

u+=cos(4x2),u=103sin(4y2)+1500.u_{+}=\cos(4x-2),\quad u_{-}=10^{3}\sin(4y-2)+1500.

This makes g,gΓ0g,g_{\Gamma}\neq 0 on Γ\Gamma, and the Dirichlet boundary condition, gbg_{b}, nonzero on Ω\partial\Omega. See Table 4 for numerical results and Fig. 5 for plots.

3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (traditional) or ΦJ2D\Phi^{2D}_{J} (FEM)
JJ NJN_{J} κ\kappa uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order NJN_{J} uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order
4 4585 5.33E+3 1.27E-1 3.85E-1 225 2.47E-1 7.19E-1
5 22857 8.23E+3 3.01E-2 1.79 1.98E-1 0.828 961 1.82E-1 0.417 5.30E-1 0.420
6 97497 9.71E+3 7.79E-3 1.86 9.27E-2 1.05 3969 9.62E-2 0.900 3.51E-1 0.580
7 398713 1.08E+4 1.58E-3 2.26 4.71E-2 0.961 16129 5.24E-2 0.866 2.44E-1 0.519
Table 4. Numerical results for Example 3.3. The estimated constant CwC_{w} in (2.19) is less than 1111.
\begin{overpic}[width=108.405pt]{flower8_gamma.pdf} \put(40.0,40.0){\tiny$a_{-}=10^{-3}$} \put(40.0,71.0){$a_{+}=1$} \end{overpic}
Refer to caption
Refer to caption
Figure 5. Example 3.3. Left: the plot of Γ\Gamma. Middle: the plot of the approximated solution at J=7J=7. Right: the plot of the error at J=7J=7.
Example 3.4.

Consider the model problem (1.1d), where a+=104a_{+}=10^{4}, a=1a_{-}=1,

Γ={(x,y)Ω:\displaystyle\Gamma=\{(x,y)\in\Omega\;: x(θ)=101/2(1+25sin(6θ))1/4cos(θ)+12,\displaystyle x(\theta)=10^{-1/2}(1+\tfrac{2}{5}\sin(6\theta))^{-1/4}\cos(\theta)+\tfrac{1}{2},
y(θ)=101/2(1+25sin(6θ))1/4sin(θ)+12,θ[0,2π)},\displaystyle y(\theta)=10^{-1/2}(1+\tfrac{2}{5}\sin(6\theta))^{-1/4}\sin(\theta)+\tfrac{1}{2},\;\theta\in[0,2\pi)\},

and f,g,gΓf,g,g_{\Gamma} are chosen such that the exact solution, uu, is

u+=a+1(((x12)2+(y12)2)2(1+25sin(6Θ(x,y)))102),u=a1a+u+,u_{+}=a_{+}^{-1}(((x-\tfrac{1}{2})^{2}+(y-\tfrac{1}{2})^{2})^{2}(1+\tfrac{2}{5}\sin(6\Theta(x,y)))-10^{-2}),\quad u_{-}=a_{-}^{-1}a_{+}u_{+},

where Θ\Theta is defined as in (3.2) for x,yΩx,y\in\Omega and Θ(1/2,1/2):=0\Theta(1/2,1/2):=0. This makes g=gΓ=0g=g_{\Gamma}=0 on Γ\Gamma, and the Dirichlet boundary condition, gbg_{b}, nonzero on Ω\partial\Omega. See Table 5 for numerical results and Fig. 6 for plots.

3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (traditional) or ΦJ2D\Phi^{2D}_{J} (FEM)
JJ NJN_{J} κ\kappa uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order NJN_{J} uuJ2u2\frac{\|u-u_{J}\|_{2}}{\|u\|_{2}} order uuJ2u2\frac{\|\nabla u-\nabla u_{J}\|_{2}}{\|\nabla u\|_{2}} order
4 3401 7.26E+4 1.19E-1 3.77E-1 225 6.02E-1 8.37E-1
5 14361 9.47E+4 2.87E-2 1.96 1.97E-1 0.901 961 3.90E-1 0.598 7.10E-1 0.228
6 59361 1.21E+5 7.92E-3 1.82 1.04E-1 0.903 3969 2.04E-1 0.914 5.16E-1 0.449
7 241409 1.36E+5 2.08E-3 1.91 4.97E-2 1.04 16129 9.68E-2 1.06 3.60E-1 0.515
Table 5. Numerical results for Example 3.4. The estimated constant CwC_{w} in (2.19) is less than 1414.
\begin{overpic}[width=108.405pt]{flower6_gamma.pdf} \put(40.0,40.0){$a_{-}=1$} \put(35.0,70.0){$a_{+}=10^{4}$} \end{overpic}
Refer to caption
Refer to caption
Figure 6. Example 3.4. Left: the plot of Γ\Gamma. Middle: the plot of the approximated solution at J=7J=7. Right: the plot of the error at J=7J=7.

3.3. Examples with unknown exact solutions

We present two examples, where the exact solutions are unknown.

Example 3.5.

Consider the model problem (1.1d), where a+=1a_{+}=1, a=10a_{-}=10,

Γ={(x,y)Ω:x(θ)=14cos(θ)+12,y(θ)=14cos(θ)+12,θ[0,2π)},\Gamma=\{(x,y)\in\Omega\;:\;x(\theta)=\tfrac{1}{4}\cos(\theta)+\tfrac{1}{2},\;y(\theta)=\tfrac{1}{4}\cos(\theta)+\tfrac{1}{2},\;\theta\in[0,2\pi)\},

f=1f=1, and g=gΓ=gb=0g=g_{\Gamma}=g_{b}=0. The exact solution, uu, is unknown. See Table 6 for numerical results and Fig. 7 for plots.

3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (traditional) or ΦJ2D\Phi^{2D}_{J} (FEM)
JJ NJN_{J} κ\kappa uJ+1uJ2\|u_{J+1}-u_{J}\|_{2} order uJ+1uJ2\|\nabla u_{J+1}-\nabla u_{J}\|_{2} order NJN_{J} uJ+1uJ2\|u_{J+1}-u_{J}\|_{2} order uJ+1uJ2\|\nabla u_{J+1}-\nabla u_{J}\|_{2} order
4 2345 1.42E+2 24.4 1.49E+3 225 48.8 1.95E+3
5 10401 1.53E+2 6.40 1.80 7.80E+2 0.865 961 26.6 0.836 1.26E+3 0.604
6 43449 1.65E+2 1.64 1.91 3.91E+2 0.964 3969 12.9 1.02 8.08E+2 0.625
Table 6. Numerical results Example 3.5. The estimated constant CwC_{w} in (2.19) is less than 1717.
\begin{overpic}[width=108.405pt]{circle_gamma.pdf} \put(37.0,68.0){$a_{+}=10$} \put(40.0,38.0){$a_{-}=1$} \end{overpic}
Refer to caption
Refer to caption
Figure 7. Example 3.5. Left: the plot of Γ\Gamma. Middle: the plot of the approximated solution at J=7J=7. Right: the plot of the error |u6u7||u_{6}-u_{7}|.
Example 3.6.

Consider the model problem (1.1d), where a+=103(2+cos(4x2)cos(4y2))a_{+}=10^{3}(2+\cos(4x-2)\cos(4y-2)) and a=2+cos(4x2)cos(4y2)a_{-}=2+\cos(4x-2)\cos(4y-2),

Γ={(x,y)Ω:\displaystyle\Gamma=\{(x,y)\in\Omega\;: x(θ)=12(12+14sin(3θ))cos(θ)+12,\displaystyle x(\theta)=\tfrac{1}{2}(\tfrac{1}{2}+\tfrac{1}{4}\sin(3\theta))\cos(\theta)+\tfrac{1}{2},
y(θ)=12(12+14sin(3θ))sin(θ)+12,θ[0,2π)},\displaystyle y(\theta)=\tfrac{1}{2}(\tfrac{1}{2}+\tfrac{1}{4}\sin(3\theta))\sin(\theta)+\tfrac{1}{2},\;\theta\in[0,2\pi)\},

f+=16sin(π(4x2))sin(π(4y2))f_{+}=-16\sin(\pi(4x-2))\sin(\pi(4y-2)), f=16cos(π(4x2))cos(π(4y2))f_{-}=-16\cos(\pi(4x-2))\cos(\pi(4y-2)), gb=0g_{b}=0, and g=sin(θ)1g=-\sin(\theta)-1, gΓ=cos(θ)g_{\Gamma}=\cos(\theta) for θ[0,2π)\theta\in[0,2\pi). The exact solution, uu, is unknown. See Table 7 for numerical results and Fig. 8 for plots.

3,JS,H01(Ω)\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{3,J} (ours) 3,J2D,H01(Ω)\mathcal{B}^{2D,H^{1}_{0}(\Omega)}_{3,J} (traditional) or ΦJ2D\Phi^{2D}_{J} (FEM)
JJ NJN_{J} κ\kappa uJ+1uJ2\|u_{J+1}-u_{J}\|_{2} order uJ+1uJ2\|\nabla u_{J+1}-\nabla u_{J}\|_{2} order NJN_{J} uJ+1uJ2\|u_{J+1}-u_{J}\|_{2} order uJ+1uJ2\|\nabla u_{J+1}-\nabla u_{J}\|_{2} order
4 3383 5.86E+3 40.0 2.64E+3 225 2.16E+2 5.84E+3
5 14775 7.63E+3 11.0 1.74 1.34E+3 0.913 961 8.53E+1 1.28 3.38E+3 0.756
6 61277 8.65E+3 3.02 1.82 6.99E+2 0.917 3969 5.92E+1 0.515 3.18E+3 0.086
Table 7. Numerical results Example 3.6. The estimated constant CwC_{w} in (2.19) is less than 33.
\begin{overpic}[width=108.405pt]{flower3_QF_gamma.pdf} \put(47.0,40.0){$a_{-}$} \put(47.0,70.0){$a_{+}$} \end{overpic}
Refer to caption
Refer to caption
Figure 8. Example 3.6. Left: the plot of Γ\Gamma. Middle: the plot of the approximated solution at J=7J=7. Right: the plot of the error |u6u7||u_{6}-u_{7}|.

4. Proof of Theorem 2.2

In this section, we shall prove Theorem 2.2 for the convergence rate of the Galerkin scheme using the biorthogonal wavelet on the unit interval (0,1)(0,1) described in Subsection 2.2.

Throughout this section, the functions ϕ,ψ\phi,\psi and ϕ~,ψ~\tilde{\phi},\tilde{\psi} are given in (2.3) and (2.4) with their biorthogonal wavelet filter bank ({a~;b~},{a;b})(\{\tilde{a};\tilde{b}\},\{a;b\}) in (2.5) and (2.6). Then ({ϕ~;ψ~},{ϕ;ψ})(\{\tilde{\phi};\tilde{\psi}\},\{\phi;\psi\}) forms a biorthogonal wavelet in L2()L_{2}(\mathbb{R}). It is important to notice that ϕ~,ψ~Hτ()\tilde{\phi},\tilde{\psi}\in H^{\tau}(\mathbb{R}) with τ<0.440765\tau<0.440765, while ϕ,ψHτ()\phi,\psi\in H^{\tau}(\mathbb{R}) with τ<1.5\tau<1.5. Moreover, both wavelet functions ψ\psi and ψ~\tilde{\psi} have order two vanishing moments.

To prove Theorem 2.2, we need three auxiliary results. The first auxiliary result deals with the weighted Bessel property in the fractional Sobolev space Hτ(2)H^{\tau}(\mathbb{R}^{2}) with τ\tau\in\mathbb{R} for the wavelet system generated by the dual wavelet function ψ~\tilde{\psi} and the dual refinable function ϕ~\tilde{\phi} in Subsection 2.2. To prove the first auxiliary result, we recall the bracket product for functions f,g:2f,g:\mathbb{R}^{2}\rightarrow\mathbb{C} as follows:

[f,g](ξ):=k2f(ξ+2πk)g(ξ+2πk)¯,ξ2[f,g](\xi):=\sum_{k\in\mathbb{Z}^{2}}f(\xi+2\pi k)\overline{g(\xi+2\pi k)},\quad\xi\in\mathbb{R}^{2}

provided that the series converges absolutely for almost every ξ2\xi\in\mathbb{R}^{2}.

Using the ideas in [24, Theorem 4.6.5] and [29, Theorem 2.3], we can establish the following result.

Theorem 4.1.

Let η~{ϕ~ψ~,ψ~ϕ~,ψ~ψ~}\tilde{\eta}\in\{\tilde{\phi}\otimes\tilde{\psi},\tilde{\psi}\otimes\tilde{\phi},\tilde{\psi}\otimes\tilde{\psi}\} with ϕ~,ψ~\tilde{\phi},\tilde{\psi} in (2.4) and masks in (2.6). Let 0<τ1<τ2<20<\tau_{1}<\tau_{2}<2. For any τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}], there exists a positive constant CC, which is independent of τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}] but may depend on τ1\tau_{1} and τ2\tau_{2}, such that

j=0k222τj|v,η~j,k|2CvHτ(2)2,for all vHτ(2),\sum_{j=0}^{\infty}\sum_{k\in\mathbb{Z}^{2}}2^{2\tau j}|\langle v,\tilde{\eta}_{j,k}\rangle|^{2}\leqslant C\|v\|_{H^{\tau}(\mathbb{R}^{2})}^{2},\qquad\mbox{for all }\quad v\in H^{\tau}(\mathbb{R}^{2}), (4.1)

where η~j,k:=2jη~(2jk)\tilde{\eta}_{j,k}:=2^{j}\tilde{\eta}(2^{j}\cdot-k), which is the dilated and shifted version of the bivariate function η~\tilde{\eta}.

Proof.

Because η~L2(2)\tilde{\eta}\in L_{2}(\mathbb{R}^{2}) has compact support, we have [η~^,η~^](ξ)=k2η~,η~(k)eikξ[\widehat{\tilde{\eta}},\widehat{\tilde{\eta}}](\xi)=\sum_{k\in\mathbb{Z}^{2}}\langle\tilde{\eta},\tilde{\eta}(\cdot-k)\rangle e^{-ik\cdot\xi} (e.g., see [24, Lemma 4.4.1]), which is a bivariate 2π22\pi\mathbb{Z}^{2}-periodic trigonometric polynomial. Hence, [η~^,η~^]L(𝕋2)[\widehat{\tilde{\eta}},\widehat{\tilde{\eta}}]\in L_{\infty}(\mathbb{T}^{2}), which can be also deduced from [29, Proposition 2.6] or [24, Lemma 6.3.2].

Let k:=(k1,k2)2k:=(k_{1},k_{2})\in\mathbb{Z}^{2}. We observe that

[π,π]2[v^(2j),η~^](ξ)eikξdξ\displaystyle\int_{[-\pi,\pi]^{2}}[\widehat{v}(2^{j}\cdot),\widehat{\tilde{\eta}}](\xi)e^{ik\cdot\xi}d\xi =ππππ[v^(2j,2j),η~^](ξ1,ξ2)ei(k1ξ1+k2ξ2)dξ1dξ2\displaystyle=\int_{-\pi}^{\pi}\int_{-\pi}^{\pi}[\widehat{v}(2^{j}\cdot,2^{j}\cdot),\widehat{\tilde{\eta}}](\xi_{1},\xi_{2})e^{i(k_{1}\xi_{1}+k_{2}\xi_{2})}d\xi_{1}d\xi_{2}
=v^(2jξ1,2jξ2)η~^(ξ1,ξ2)¯ei(k1ξ1+k2ξ2)𝑑ξ1𝑑ξ2\displaystyle=\int_{\mathbb{R}}\int_{\mathbb{R}}\widehat{v}(2^{j}\xi_{1},2^{j}\xi_{2})\overline{\widehat{\tilde{\eta}}(\xi_{1},\xi_{2})}e^{i(k_{1}\xi_{1}+k_{2}\xi_{2})}d\xi_{1}d\xi_{2}
=122jv^,η~^(2j,2j)ei(k1ξ1+k2ξ2)=(2π)222jv,η~j,k,\displaystyle=\frac{1}{2^{2j}}\langle\widehat{v},\widehat{\tilde{\eta}}(2^{-j}\cdot,2^{-j}\cdot)e^{-i(k_{1}\xi_{1}+k_{2}\xi_{2})}\rangle=\frac{(2\pi)^{2}}{2^{2j}}\langle v,\tilde{\eta}_{j,k}\rangle,

which due to Parseval’s identity yields

k2|v,η~j,k|2=22j(2π)2[π,π]2|[v^(2j),η~^](ξ)|2dξ.\sum_{k\in\mathbb{Z}^{2}}|\langle v,\tilde{\eta}_{j,k}\rangle|^{2}=\frac{2^{2j}}{(2\pi)^{2}}\int_{[-\pi,\pi]^{2}}|[\widehat{v}(2^{j}\cdot),\widehat{\tilde{\eta}}](\xi)|^{2}d\xi.

Since (x+y)22(x2+y2)(x+y)^{2}\leqslant 2(x^{2}+y^{2}) for all x,yx,y\in\mathbb{R}, we have

|[v^(2j),η~^](ξ)|2\displaystyle|[\widehat{v}(2^{j}\cdot),\widehat{\tilde{\eta}}](\xi)|^{2} 2|v^(2jξ)η~^(ξ)¯|2+2|k2\{(0,0)}v^(2j(ξ+2πk))η~^(ξ+2πk)¯|2.\displaystyle\leqslant 2|\widehat{v}(2^{j}\xi)\overline{\widehat{\tilde{\eta}}(\xi)}|^{2}+2\left|\sum_{k\in\mathbb{Z}^{2}\backslash\{(0,0)\}}\widehat{v}(2^{j}(\xi+2\pi k))\overline{\widehat{\tilde{\eta}}(\xi+2\pi k)}\right|^{2}.

Hence, it follows that

k2|v,η~j,k|222j1π2[π,π]2|v^(2jξ)η~^(ξ)¯|2𝑑ξ+22j1π2[π,π]2|k2\{(0,0)}v^(2j(ξ+2πk))η~^(ξ+2πk)¯|2𝑑ξ\displaystyle\sum_{k\in\mathbb{Z}^{2}}|\langle v,\tilde{\eta}_{j,k}\rangle|^{2}\leqslant\frac{2^{2j-1}}{\pi^{2}}\int_{[-\pi,\pi]^{2}}|\widehat{v}(2^{j}\xi)\overline{\widehat{\tilde{\eta}}(\xi)}|^{2}d\xi+\frac{2^{2j-1}}{\pi^{2}}\int_{[-\pi,\pi]^{2}}\left|\sum_{k\in\mathbb{Z}^{2}\backslash\{(0,0)\}}\widehat{v}(2^{j}(\xi+2\pi k))\overline{\widehat{\tilde{\eta}}(\xi+2\pi k)}\right|^{2}d\xi
12π22|v^(ξ)η~^(2jξ)|2χ[π,π]2(2jξ)𝑑ξ\displaystyle\leqslant\frac{1}{2\pi^{2}}\int_{\mathbb{R}^{2}}|\widehat{v}(\xi)\widehat{\tilde{\eta}}(2^{-j}\xi)|^{2}\chi_{[-\pi,\pi]^{2}}(2^{-j}\xi)d\xi
+22j1π2[π,π]2[k2\{(0,0)}|v^(2j(ξ+2πk))|2][k2\{(0,0)}|η~^(ξ+2πk)|2]𝑑ξ\displaystyle\qquad\qquad+\frac{2^{2j-1}}{\pi^{2}}\int_{[-\pi,\pi]^{2}}\Big{[}\sum_{k\in\mathbb{Z}^{2}\backslash\{(0,0)\}}|\widehat{v}(2^{j}(\xi+2\pi k))|^{2}\Big{]}\Big{[}\sum_{k\in\mathbb{Z}^{2}\backslash\{(0,0)\}}|\widehat{\tilde{\eta}}(\xi+2\pi k)|^{2}\Big{]}d\xi
12π22|v^(ξ)η~^(2jξ)|2χ[π,π]2(2jξ)𝑑ξ+22jCη~2π2[π,π]2k2\{(0,0)}|v^(2j(ξ+2πk))|2dξ\displaystyle\leqslant\frac{1}{2\pi^{2}}\int_{\mathbb{R}^{2}}|\widehat{v}(\xi)\widehat{\tilde{\eta}}(2^{-j}\xi)|^{2}\chi_{[-\pi,\pi]^{2}}(2^{-j}\xi)d\xi+\frac{2^{2j}C_{\tilde{\eta}}}{2\pi^{2}}\int_{[-\pi,\pi]^{2}}\sum_{k\in\mathbb{Z}^{2}\backslash\{(0,0)\}}|\widehat{v}(2^{j}(\xi+2\pi k))|^{2}d\xi
12π22|v^(ξ)η~^(2jξ)|2χ[π,π]2(2jξ)𝑑ξ+Cη~2π22|v^(ξ)|2χ2\[π,π]2(2jξ)𝑑ξ,\displaystyle\leqslant\frac{1}{2\pi^{2}}\int_{\mathbb{R}^{2}}|\widehat{v}(\xi)\widehat{\tilde{\eta}}(2^{-j}\xi)|^{2}\chi_{[-\pi,\pi]^{2}}(2^{-j}\xi)d\xi+\frac{C_{\tilde{\eta}}}{2\pi^{2}}\int_{\mathbb{R}^{2}}|\widehat{v}(\xi)|^{2}\chi_{\mathbb{R}^{2}\backslash[-\pi,\pi]^{2}}(2^{-j}\xi)d\xi,

where we already proved that Cη~:=[η~^,η~^]L(𝕋2)<C_{\tilde{\eta}}:=\|[\widehat{\tilde{\eta}},\widehat{\tilde{\eta}}]\|_{L_{\infty}(\mathbb{T}^{2})}<\infty. As a result, we have

j=0k222τj|v,η~j,k|214π22|v^(ξ)|2(1+ξ2)τ(2B1(ξ)+2Cη~B2(ξ))𝑑ξ,\sum_{j=0}^{\infty}\sum_{k\in\mathbb{Z}^{2}}2^{2\tau j}|\langle v,\tilde{\eta}_{j,k}\rangle|^{2}\leqslant\frac{1}{4\pi^{2}}\int_{\mathbb{R}^{2}}|\widehat{v}(\xi)|^{2}(1+\|\xi\|^{2})^{\tau}(2B_{1}(\xi)+2C_{\tilde{\eta}}B_{2}(\xi))d\xi, (4.2)

where

B1(ξ)\displaystyle B_{1}(\xi) :=(1+ξ2)τj=022τj|η~^(2jξ)|2χ[π,π]2(2jξ),\displaystyle:=(1+\|\xi\|^{2})^{-\tau}\sum_{j=0}^{\infty}2^{2\tau j}|\widehat{\tilde{\eta}}(2^{-j}\xi)|^{2}\chi_{[-\pi,\pi]^{2}}(2^{-j}\xi),
B2(ξ)\displaystyle B_{2}(\xi) :=(1+ξ2)τj=022τjχ2\[π,π]2(2jξ).\displaystyle:=(1+\|\xi\|^{2})^{-\tau}\sum_{j=0}^{\infty}2^{2\tau j}\chi_{\mathbb{R}^{2}\backslash[-\pi,\pi]^{2}}(2^{-j}\xi).

We first estimate B1(ξ)B_{1}(\xi). Recall that η~\tilde{\eta} takes the form of ϕ~ψ~\tilde{\phi}\otimes\tilde{\psi}, ψ~ϕ~\tilde{\psi}\otimes\tilde{\phi}, or ψ~ψ~\tilde{\psi}\otimes\tilde{\psi}. Thus, one of the following inequalities holds: |ϕ~^(ξ1)ψ~^(ξ2)|2Cvm2|ξ2|4|\widehat{\tilde{\phi}}(\xi_{1})\widehat{\tilde{\psi}}(\xi_{2})|^{2}\leqslant C_{\operatorname{vm}}^{2}|\xi_{2}|^{4}, |ψ~^(ξ1)ϕ~^(ξ2)|2Cvm2|ξ1|4|\widehat{\tilde{\psi}}(\xi_{1})\widehat{\tilde{\phi}}(\xi_{2})|^{2}\leqslant C_{\operatorname{vm}}^{2}|\xi_{1}|^{4}, or |ψ~^(ξ1)ψ~^(ξ2)|2Cvm2|ξ1ξ2|2|\widehat{\tilde{\psi}}(\xi_{1})\widehat{\tilde{\psi}}(\xi_{2})|^{2}\leqslant C_{\operatorname{vm}}^{2}|\xi_{1}\xi_{2}|^{2} for some positive constant CvmC_{\operatorname{vm}} and for almost every ξ1,ξ2[π,π]\xi_{1},\xi_{2}\in[-\pi,\pi]. That is, |η~^(ξ)|2=|η~^(ξ1,ξ2)|2Cvm2(|ξ1|2+|ξ2|2)2=Cvm2ξ4|\widehat{\tilde{\eta}}(\xi)|^{2}=|\widehat{\tilde{\eta}}(\xi_{1},\xi_{2})|^{2}\leqslant C_{\operatorname{vm}}^{2}(|\xi_{1}|^{2}+|\xi_{2}|^{2})^{2}=C_{\operatorname{vm}}^{2}\|\xi\|^{4} for almost every ξ[π,π]2\xi\in[-\pi,\pi]^{2}. Define

Jξ:=max{0,log2(ξ/π)}.J_{\xi}:=\max\{0,\lceil\log_{2}(\|\xi\|/\pi)\rceil\}.

Then, for all τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}], because 0<τ1<τ2<20<\tau_{1}<\tau_{2}<2 and τ2<0\tau-2<0, we have

B1(ξ)\displaystyle B_{1}(\xi) Cvm2(1+ξ2)τξ4j=Jξ22j(τ2)Cvm2(1+ξ2)τξ422Jξ(τ2)122(τ2)\displaystyle\leqslant C_{\operatorname{vm}}^{2}(1+\|\xi\|^{2})^{-\tau}\|\xi\|^{4}\sum_{j=J_{\xi}}^{\infty}2^{2j(\tau-2)}\leqslant C_{\operatorname{vm}}^{2}(1+\|\xi\|^{2})^{-\tau}\|\xi\|^{4}\frac{2^{2J_{\xi}(\tau-2)}}{1-2^{2(\tau-2)}}
Cvm2(1+ξ2)τξ4ξ2(2τ)π2(2τ)122(τ2)=Cvm2(ξ21+ξ2)τπ2(2τ)122(τ2)π4122(τ22)Cvm2,\displaystyle\leqslant C_{\operatorname{vm}}^{2}(1+\|\xi\|^{2})^{-\tau}\frac{\|\xi\|^{4}}{\|\xi\|^{2(2-\tau)}}\frac{\pi^{2(2-\tau)}}{1-2^{2(\tau-2)}}=C_{\operatorname{vm}}^{2}\Big{(}\frac{\|\xi\|^{2}}{1+\|\xi\|^{2}}\Big{)}^{\tau}\frac{\pi^{2(2-\tau)}}{1-2^{2(\tau-2)}}\leqslant\frac{\pi^{4}}{1-2^{2(\tau_{2}-2)}}C_{\operatorname{vm}}^{2},

which implies that B1(ξ)L(2)B_{1}(\xi)\in L_{\infty}(\mathbb{R}^{2}). Define jξ:=max{0,log2(ξ/π)}j_{\xi}:=\max\{0,\lfloor\log_{2}(\|\xi\|/\pi)\rfloor\}. Meanwhile, for τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}], by τ>τ1>0\tau>\tau_{1}>0, we have

B2(ξ)\displaystyle B_{2}(\xi) (1+ξ2)τj=0jξ22τj(1+ξ2)τ22τ(jξ+1)22τ1(22τπ2τ22τ1)(ξ21+ξ2)τ22τ122τ11,\displaystyle\leqslant(1+\|\xi\|^{2})^{-\tau}\sum_{j=0}^{j_{\xi}}2^{2\tau j}\leqslant(1+\|\xi\|^{2})^{-\tau}\frac{2^{2\tau(j_{\xi}+1)}}{2^{2\tau}-1}\leqslant\left(\frac{2^{2\tau}\pi^{-2\tau}}{2^{2\tau}-1}\right)\Big{(}\frac{\|\xi\|^{2}}{1+\|\xi\|^{2}}\Big{)}^{\tau}\leqslant\frac{2^{2\tau_{1}}}{2^{2\tau_{1}}-1},

which implies that B2(ξ)L(2)B_{2}(\xi)\in L_{\infty}(\mathbb{R}^{2}). Continuing from (4.2), we have

j=0k222τj|v,η~j,k|2C14π22|v^(ξ)|2(1+ξ2)τ𝑑ξ,\sum_{j=0}^{\infty}\sum_{k\in\mathbb{Z}^{2}}2^{2\tau j}|\langle v,\tilde{\eta}_{j,k}\rangle|^{2}\leqslant C\frac{1}{4\pi^{2}}\int_{\mathbb{R}^{2}}|\widehat{v}(\xi)|^{2}(1+\|\xi\|^{2})^{\tau}d\xi,

where C:=2B1L(2)+2Cη~B2L(2)C:=2\|B_{1}\|_{L_{\infty}(\mathbb{R}^{2})}+2C_{\tilde{\eta}}\|B_{2}\|_{L_{\infty}(\mathbb{R}^{2})} with

B1L(2)π4122(τ22)Cvm2andB2L(2)22τ122τ11\|B_{1}\|_{L_{\infty}(\mathbb{R}^{2})}\leqslant\frac{\pi^{4}}{1-2^{2(\tau_{2}-2)}}C_{\operatorname{vm}}^{2}\quad\text{and}\quad\|B_{2}\|_{L_{\infty}(\mathbb{R}^{2})}\leqslant\frac{2^{2\tau_{1}}}{2^{2\tau_{1}}-1}

for all τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}] with 0<τ1<τ2<20<\tau_{1}<\tau_{2}<2. We obtain the desired conclusion. ∎

In preparation for the next auxiliary result, we introduce a few notations and present a few observations. Recall that Ψ~j2D={Φ~jΨ~j,Ψ~jΦ~j,Ψ~jΨ~j}\tilde{\Psi}^{2D}_{j}=\{\tilde{\Phi}_{j}\otimes\tilde{\Psi}_{j},\tilde{\Psi}_{j}\otimes\tilde{\Phi}_{j},\tilde{\Psi}_{j}\otimes\tilde{\Psi}_{j}\}. We can split the set Ψ~j2D\tilde{\Psi}^{2D}_{j} into two groups:

Gjx:=[Ψ~jΦ~j][Ψ~jΨ~j]andGjy:=Φ~jΨ~j.G_{j}^{x}:=[\tilde{\Psi}_{j}\otimes\tilde{\Phi}_{j}]\cup[\tilde{\Psi}_{j}\otimes\tilde{\Psi}_{j}]\quad\text{and}\quad G_{j}^{y}:=\tilde{\Phi}_{j}\otimes\tilde{\Psi}_{j}. (4.3)

Note that any α~jΨ~j2D\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j} must belong to either GjxG_{j}^{x} or GjyG_{j}^{y}. By construction, all elements in Ψ~j\tilde{\Psi}_{j} must have order two vanishing moments, i.e., 01η(x)𝑑x=0\int_{0}^{1}\eta(x)dx=0 for all ηΨ~j\eta\in\tilde{\Psi}_{j}. We define an integration operation along one axis as follows:

α~̊j:={2jx1α~j(t)𝑑t,α~jΨ~j,2jx1α~j(t,y)𝑑t,α~jGjx,2jy1α~j(x,t)𝑑t,α~jGjy.\mathring{\tilde{\alpha}}_{j}:=\begin{cases}2^{j}\int_{x}^{1}\tilde{\alpha}_{j}(t)dt,&\tilde{\alpha}_{j}\in\tilde{\Psi}_{j},\\ 2^{j}\int_{x}^{1}\tilde{\alpha}_{j}(t,y)dt,&\tilde{\alpha}_{j}\in G_{j}^{x},\\ 2^{j}\int_{y}^{1}\tilde{\alpha}_{j}(x,t)dt,&\tilde{\alpha}_{j}\in G_{j}^{y}.\end{cases}

Since the tensor product part of α~jGjx\tilde{\alpha}_{j}\in G_{j}^{x} in the xx-coordinate has order two vanishing moments and α~jH01(Ω)\tilde{\alpha}_{j}\in H_{0}^{1}(\Omega), we conclude that α~̊jH01(Ω)\mathring{\tilde{\alpha}}_{j}\in H_{0}^{1}(\Omega) and it must have order one vanishing moment. Moreover, if α~jGjx\tilde{\alpha}_{j}\in G_{j}^{x}, then

α~̊j{Ψ~˘jΦj,Ψ~˘jΨj},for alljJ0,\mathring{\tilde{\alpha}}_{j}\in\{\breve{\tilde{\Psi}}_{j}\otimes\Phi_{j},\breve{\tilde{\Psi}}_{j}\otimes\Psi_{j}\},\qquad\mbox{for all}\quad j\geqslant J_{0}, (4.4)

where Ψ~˘j={ψ~̊j;0L}{ψ~̊j;k: 2k2j3}{ψ~̊j;2j1R}\breve{\tilde{\Psi}}_{j}=\{\mathring{\tilde{\psi}}^{L}_{j;0}\}\cup\{\mathring{\tilde{\psi}}_{j;k}\;:\;2\leqslant k\leqslant 2^{j}-3\}\cup\{\mathring{\tilde{\psi}}^{R}_{j;2^{j}-1}\}. Similarly, since the tensor product part of α~jGjy\tilde{\alpha}_{j}\in G^{y}_{j} in the yy-coordinate has order two vanishing moments and α~jH01(Ω)\tilde{\alpha}_{j}\in H^{1}_{0}(\Omega), we conclude that α~̊jH01(Ω)\mathring{\tilde{\alpha}}_{j}\in H^{1}_{0}(\Omega) and it must have order one vanishing moment. Moreover, if α~jGjy\tilde{\alpha}_{j}\in G^{y}_{j}, then α~̊jΦjΨ~˘j\mathring{\tilde{\alpha}}_{j}\in\Phi_{j}\otimes\breve{\tilde{\Psi}}_{j} for all jJ0j\geqslant J_{0}, where Ψ~˘j\breve{\tilde{\Psi}}_{j} is defined the same way as before.

To prove Theorem 2.2, we shall also need the following second auxiliary result.

Theorem 4.2.

Let Ψ~j2D\tilde{\Psi}_{j}^{2D} with jJ0j\geqslant J_{0} be defined in (2.10). Then there exists a positive constant CC such that

j=Jα~jΨ~j2D22j|v,α~j|222JC|v|H2(Ω)2,for all JJ0,vH2(Ω)H01(Ω),\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j}}2^{2j}|\langle v,\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-2J}C|v|^{2}_{H^{2}(\Omega)},\qquad\mbox{for all }J\geqslant J_{0},v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega), (4.5)

where |v|H2(Ω)|v|_{H^{2}(\Omega)} is the semi-norm of vv in H2(Ω)H^{2}(\Omega), i.e., |v|H2(Ω)2:=|μ|=2μvL2(Ω)2|v|_{H^{2}(\Omega)}^{2}:=\sum_{|\mu|=2}\|\partial^{\mu}v\|_{L_{2}(\Omega)}^{2}, where μ:=(μ1,μ2)\mu:=(\mu_{1},\mu_{2}) and μ:=μ1+μ2xμ1yμ2\partial^{\mu}:=\tfrac{\partial^{\mu_{1}+\mu_{2}}}{\partial x^{\mu_{1}}\partial y^{\mu_{2}}} with μ1,μ2{0}\mu_{1},\mu_{2}\in\mathbb{N}\cup\{0\}.

Proof.

Define VJ=span(ΦJ2D)V_{J}=\text{span}(\Phi^{2D}_{J}) and h:=2Jh:=2^{-J}. From Section 2.3, we know that VJV_{J} is just the finite element space of the bilinear elements with the mesh size h=2Jh=2^{-J}. Also, note that VJ=span(J0,J2D)V_{J}=\text{span}(\mathcal{B}^{2D}_{J_{0},J}).

For vH2(Ω)H01(Ω)v\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega), we define IhvI_{h}v to be the interpolation function of vv on the grid of Ω\Omega using the mesh size hh, i.e., on the grid Ω2J2\Omega\cap 2^{-J}\mathbb{Z}^{2} such that [Ihv](p)=v(p)[I_{h}v](p)=v(p) for all pΩ2J2p\in\Omega\cap 2^{-J}\mathbb{Z}^{2}. By [4, Theorem 4.6.14], there exists a positive constant C0C_{0}, only depends on the hat function ϕ\phi, such that

|vIhv|H1(Ω):=(vIhv)L2(Ω)C0h|v|H2(Ω)=C02J|v|H2(Ω).|v-I_{h}v|_{H^{1}(\Omega)}:=\|\nabla(v-I_{h}v)\|_{L_{2}(\Omega)}\leqslant C_{0}h|v|_{H^{2}(\Omega)}=C_{0}2^{-J}|v|_{H^{2}(\Omega)}. (4.6)

Note that IhvH01(Ω)I_{h}v\in H^{1}_{0}(\Omega). Because IhvVJ=span(J0,J2D)I_{h}v\in V_{J}=\text{span}(\mathcal{B}^{2D}_{J_{0},J}) and (~J02D,J02D)(\tilde{\mathcal{B}}_{J_{0}}^{2D},\mathcal{B}_{J_{0}}^{2D}) is a biorthogonal wavelet in L2(Ω)L_{2}(\Omega), by J0,J2DJ02D\mathcal{B}^{2D}_{J_{0},J}\subset\mathcal{B}^{2D}_{J_{0}}, it is critical to notice the following perpendicular condition:

Ihv,α~j=0for all α~jΨ~j2D with jJ.\langle I_{h}v,\tilde{\alpha}_{j}\rangle=0\qquad\mbox{for all }\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j}\mbox{ with }j\geqslant J. (4.7)

We now estimate (4.5). We first handle the case, where α~jGjx\tilde{\alpha}_{j}\in G_{j}^{x}. Integrating by parts with respect to xx and using the fact that vIhvH01(Ω)v-I_{h}v\in H^{1}_{0}(\Omega) satisfies the homogeneous Dirichlet condition, we obtain

2jvIhv,α~j=x(vIhv),α~̊j,α~jGjx.2^{j}\langle v-I_{h}v,\tilde{\alpha}_{j}\rangle=\langle\tfrac{\partial}{\partial x}(v-I_{h}v),\mathring{\tilde{\alpha}}_{j}\rangle,\qquad\tilde{\alpha}_{j}\in G_{j}^{x}.

Therefore, we conclude that

j=Jα~jGjx22j|v,α~j|2\displaystyle\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in G_{j}^{x}}2^{2j}|\langle v,\tilde{\alpha}_{j}\rangle|^{2} =j=Jα~jGjx22j|vIhv,α~j|2=j=Jα~jGjx|x(vIhv),α~̊j|2\displaystyle=\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in G_{j}^{x}}2^{2j}|\langle v-I_{h}v,\tilde{\alpha}_{j}\rangle|^{2}=\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in G_{j}^{x}}|\langle\tfrac{\partial}{\partial x}(v-I_{h}v),\mathring{\tilde{\alpha}}_{j}\rangle|^{2}
j=Jα~̊j[Ψ~˘jΦj][Ψ~˘jΨj]|x(vIhv),α~̊j|2,\displaystyle\leqslant\sum_{j=J}^{\infty}\sum_{\mathring{\tilde{\alpha}}_{j}\in[\breve{\tilde{\Psi}}_{j}\otimes\Phi_{j}]\cup[\breve{\tilde{\Psi}}_{j}\otimes\Psi_{j}]}|\langle\tfrac{\partial}{\partial x}(v-I_{h}v),\mathring{\tilde{\alpha}}_{j}\rangle|^{2},

where we used (4.4) and the fact that Gjx[Ψ~˘jΦj][Ψ~˘jΨj]G_{j}^{x}\subseteq[\breve{\tilde{\Psi}}_{j}\otimes\Phi_{j}]\cup[\breve{\tilde{\Psi}}_{j}\otimes\Psi_{j}] to arrive at the last line. Because vIhvH01(Ω)v-I_{h}v\in H_{0}^{1}(\Omega), we have x(vIhv)L2(Ω)\tfrac{\partial}{\partial x}(v-I_{h}v)\in L_{2}(\Omega). Note that all the elements in [Ψ~˘jΦj][Ψ~˘jΨj][\breve{\tilde{\Psi}}_{j}\otimes\Phi_{j}]\cup[\breve{\tilde{\Psi}}_{j}\otimes\Psi_{j}] are compactly supported functions in L2(2)L_{2}(\mathbb{R}^{2}) and have at least one vanishing moment. Now by the Bessel property in [28, Lemma 6.1] and [22, Theorems 2.2 and 2.3], there must exist a positive constant C1C_{1}, independent of JJ and only depending on the wavelet, such that

j=Jα~jGj122j|v,α~j|2j=Jα~̊j[Ψ~˘jΦj][Ψ~˘jΨj]|x(vIhv),α~̊j|2C1x(vIhv)L2(Ω)2.\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in G_{j}^{1}}2^{2j}|\langle v,\tilde{\alpha}_{j}\rangle|^{2}\leqslant\sum_{j=J}^{\infty}\sum_{\mathring{\tilde{\alpha}}_{j}\in[\breve{\tilde{\Psi}}_{j}\otimes\Phi_{j}]\cup[\breve{\tilde{\Psi}}_{j}\otimes\Psi_{j}]}|\langle\tfrac{\partial}{\partial x}(v-I_{h}v),\mathring{\tilde{\alpha}}_{j}\rangle|^{2}\leqslant C_{1}\|\tfrac{\partial}{\partial x}(v-I_{h}v)\|^{2}_{L_{2}(\Omega)}. (4.8)

We now handle the case, where α~jGjy\tilde{\alpha}_{j}\in G_{j}^{y}. By using a similar calculation as before, we obtain

2jvIhv,α~j=y(vIhv),α~̊j,α~jGjy.2^{j}\langle v-I_{h}v,\tilde{\alpha}_{j}\rangle=\langle\tfrac{\partial}{\partial y}(v-I_{h}v),\mathring{\tilde{\alpha}}_{j}\rangle,\qquad\tilde{\alpha}_{j}\in G_{j}^{y}.

Furthermore, there exists a positive constant C2C_{2}, independent of JJ, such that

j=Jα~jGjy22j|v,α~j|2j=Jα~̊jΦjΨ~˘j|y(vIhv),α~̊j|2C2y(vIhv)L2(Ω)2.\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in G_{j}^{y}}2^{2j}|\langle v,\tilde{\alpha}_{j}\rangle|^{2}\leqslant\sum_{j=J}^{\infty}\sum_{\mathring{\tilde{\alpha}}_{j}\in\Phi_{j}\otimes\breve{\tilde{\Psi}}_{j}}|\langle\tfrac{\partial}{\partial y}(v-I_{h}v),\mathring{\tilde{\alpha}}_{j}\rangle|^{2}\leqslant C_{2}\|\tfrac{\partial}{\partial y}(v-I_{h}v)\|^{2}_{L_{2}(\Omega)}.

Combining the above estimates with (4.8), we conclude that

j=Jα~jΨ~j2D22j|v,α~j|2\displaystyle\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D}}2^{2j}|\langle v,\tilde{\alpha}_{j}\rangle|^{2} =j=Jα~jGjxGjy22j|v,α~j|2\displaystyle=\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in G_{j}^{x}\cup G_{j}^{y}}2^{2j}|\langle v,\tilde{\alpha}_{j}\rangle|^{2}
C1x(vIhv)L2(Ω)2+C2y(vIhv)L2(Ω)2\displaystyle\leqslant C_{1}\|\tfrac{\partial}{\partial x}(v-I_{h}v)\|^{2}_{L_{2}(\Omega)}+C_{2}\|\tfrac{\partial}{\partial y}(v-I_{h}v)\|^{2}_{L_{2}(\Omega)}
max(C1,C2)(vIhv)L2(Ω)2=max(C1,C2)|vIhv|H1(Ω)2,\displaystyle\leqslant\max(C_{1},C_{2})\|\nabla(v-I_{h}v)\|^{2}_{L_{2}(\Omega)}=\max(C_{1},C_{2})|v-I_{h}v|^{2}_{H^{1}(\Omega)},

which combined with the approximation result in (4.6) further yields

j=Jα~jΨ~j2D22j|v,α~j|2max(C1,C2)|vIhv|H1(Ω)2max(C1,C2)C02h2|v|H2(Ω)2.\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D}}2^{2j}|\langle v,\tilde{\alpha}_{j}\rangle|^{2}\leqslant\max(C_{1},C_{2})|v-I_{h}v|^{2}_{H^{1}(\Omega)}\leqslant\max(C_{1},C_{2})C_{0}^{2}h^{2}|v|^{2}_{H^{2}(\Omega)}.

This proves (4.5) with C:=max(C1,C2)C02<C:=\max(C_{1},C_{2})C_{0}^{2}<\infty, where h=2Jh=2^{-J}. ∎

We also need the following lemma in the proof of Theorem 2.2.

Lemma 4.3.

Let Ω\Omega_{-} be a bounded open domain with a smooth boundary Γ\Gamma. Then

χΩH1/2ε(2)2CΓ2ε1, 0<ε<1/4,\|\chi_{\Omega_{-}}\|^{2}_{H^{1/2-\varepsilon}(\mathbb{R}^{2})}\leqslant C_{\Gamma}^{2}\varepsilon^{-1},\qquad\forall\;0<\varepsilon<1/4, (4.9)

for a positive constant CΓC_{\Gamma} that only depends on Ω\Omega_{-} but is independent of 0<ε<1/40<\varepsilon<1/4.

Proof.

For a compactly supported function FL2(2)F\in L_{2}(\mathbb{R}^{2}), recall that its modulus of smoothness in the L2L_{2} norm is defined by

ω(F,s)2:=sup|t|sF(+t)FL2(2),s>0.\omega(F,s)_{2}:=\sup_{|t|\leqslant s}\|F(\cdot+t)-F\|_{L_{2}(\mathbb{R}^{2})},\qquad s>0. (4.10)

For any 0<τ<10<\tau<1, the semi-norm FHτ(2)F\in H^{\tau}(\mathbb{R}^{2}) is

|F|Hτ(Ω):=(0[sτω(F,s)2]2dss)1/2.|F|_{H^{\tau}(\Omega)}:=\left(\int_{0}^{\infty}[s^{-\tau}\omega(F,s)_{2}]^{2}\frac{ds}{s}\right)^{1/2}. (4.11)

For t2t\in\mathbb{R}^{2}, we define

Et:={q2:q[(2\Ω)+t]Ωorq[Ω+t][2\Ω]},E_{t}:=\{q\in\mathbb{R}^{2}\quad:\quad q\in[(\mathbb{R}^{2}\backslash\Omega_{-})+t]\cap\Omega_{-}\quad\mbox{or}\quad q\in[\Omega_{-}+t]\cap[\mathbb{R}^{2}\backslash\Omega_{-}]\},

which has a measure of order 𝒪(t)\mathcal{O}(\|t\|), because Γ=Ω¯2\Ω¯\Gamma=\overline{\Omega_{-}}\cap\overline{\mathbb{R}^{2}\backslash\Omega_{-}} is a closed smooth curve. Define F:=χΩF:=\chi_{\Omega_{-}}. Then for all t2t\in\mathbb{R}^{2}, we have

F(pt)F(p)L2(2)2=2|F(pt)F(p)|2𝑑p=Et𝑑pC12t,\|F(p-t)-F(p)\|^{2}_{L_{2}(\mathbb{R}^{2})}=\int_{\mathbb{R}^{2}}|F(p-t)-F(p)|^{2}dp=\int_{E_{t}}dp\leqslant C^{2}_{1}\|t\|,

for a positive constant C1C_{1} only depending on Ω\partial\Omega_{-}. Consequently, we have

ω(F,s)2C1s1/2,s(0,).\omega(F,s)_{2}\leqslant C_{1}s^{1/2},\qquad\forall s\in(0,\infty). (4.12)

Note ω(F,s)22FL2(2)=2|Ω|1/2\omega(F,s)_{2}\leqslant 2\|F\|_{L_{2}(\mathbb{R}^{2})}=2|\Omega_{-}|^{1/2} by the triangle inequality. Hence, for 0<ε<1/40<\varepsilon<1/4, we have

|F|H1/2ε(2)2\displaystyle|F|^{2}_{H^{1/2-\varepsilon}(\mathbb{R}^{2})} =0[s(1/2ε)ω(F,s)2]2dss=01[s(1/2ε)ω(F,s)2]2dss+1[s(1/2ε)ω(F,s)2]2dss\displaystyle=\int_{0}^{\infty}[s^{-(1/2-\varepsilon)}\omega(F,s)_{2}]^{2}\frac{ds}{s}=\int_{0}^{1}[s^{-(1/2-\varepsilon)}\omega(F,s)_{2}]^{2}\frac{ds}{s}+\int_{1}^{\infty}[s^{-(1/2-\varepsilon)}\omega(F,s)_{2}]^{2}\frac{ds}{s}
C1201s2ε1sdss+4|Ω|1s2ε1dss=C122ε+412ε|Ω|C122ε+8|Ω|,\displaystyle\leqslant C_{1}^{2}\int_{0}^{1}s^{2\varepsilon-1}s\frac{ds}{s}+4|\Omega_{-}|\int_{1}^{\infty}s^{2\varepsilon-1}\frac{ds}{s}=\frac{C_{1}^{2}}{2\varepsilon}+\frac{4}{1-2\varepsilon}|\Omega_{-}|\leqslant\frac{C_{1}^{2}}{2\varepsilon}+8|\Omega_{-}|,

where we used (4.12) for the first inequality. Since FH1/2ε(2)2=FL2(2)2+|F|H1/2ε(2)2\|F\|^{2}_{H^{1/2-\varepsilon}(\mathbb{R}^{2})}=\|F\|^{2}_{L_{2}(\mathbb{R}^{2})}+|F|^{2}_{H^{1/2-\varepsilon}(\mathbb{R}^{2})} and FL2(2)2=|Ω|2\|F\|^{2}_{L_{2}(\mathbb{R}^{2})}=|\Omega_{-}|^{2}, we proved the claim with CΓ2:=12C12+2|Ω|+14|Ω|2C_{\Gamma}^{2}:=\frac{1}{2}C_{1}^{2}+2|\Omega_{-}|+\frac{1}{4}|\Omega_{-}|^{2} for all 0<ε<1/40<\varepsilon<1/4. ∎

We are now ready to present the proof of Theorem 2.2.

Proof of Theorem 2.2.

We split the analysis and estimation in three regions: u+u_{+}, uu_{-}, and uΓu_{\Gamma}; the last of the three is the neighborhood of the interface curve Γ\Gamma.

Proving the H1(Ω)H^{1}(\Omega) convergence. Because the interface Γ\Gamma is of class 𝒞2\mathscr{C}^{2} and u+H2(Ω+)u_{+}\in H^{2}(\Omega_{+}) by our assumption (2.16), we can extend the function u+u_{+} from Ω+\Omega_{+} to the domain Ω\Omega and obtain a function v+H2(Ω)H01(Ω)v_{+}\in H^{2}(\Omega)\cap H_{0}^{1}(\Omega) such that v+=u+v_{+}=u_{+} in Ω+\Omega_{+} and v+H2(Ω)C0u+H2(Ω+)\|v_{+}\|_{H^{2}(\Omega)}\leqslant C_{0}\|u_{+}\|_{H^{2}(\Omega_{+})}. Therefore, by Theorem 4.2, there exists a positive constant C+C_{+} such that

j=Jα~jΨ~j2D22j|v+,α~j|222JC+|v+|H2(Ω)2,for all JJ0.\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j}}2^{2j}|\langle v_{+},\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-2J}C_{+}|v_{+}|^{2}_{H^{2}(\Omega)},\qquad\mbox{for all }J\geqslant J_{0}. (4.13)

If α~jΨ~j2D\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j} and supp(α~j)\operatorname{supp}(\tilde{\alpha}_{j}) is completely inside Ω+\Omega_{+}, due to v+=u+v_{+}=u_{+} on Ω+\Omega_{+}, we must have

v+,α~j=u+,α~j=u,α~j,ifsupp(α~j)Ω+.\langle v_{+},\tilde{\alpha}_{j}\rangle=\langle u_{+},\tilde{\alpha}_{j}\rangle=\langle u,\tilde{\alpha}_{j}\rangle,\qquad\mbox{if}\quad\operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{+}.

Consequently, we conclude from (4.13) and v+H2(Ω)C0u+H2(Ω+)\|v_{+}\|_{H^{2}(\Omega)}\leqslant C_{0}\|u_{+}\|_{H^{2}(\Omega_{+})} that

j=Jα~jΨ~j2D,supp(α~j)Ω+22j|u,α~j|2j=Jα~jΨ~j2D22j|v+,α~j|222JC1|u+|H2(Ω+)2,JJ0,\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{+}\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\leqslant\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j}}2^{2j}|\langle v_{+},\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-2J}C_{1}|u_{+}|^{2}_{H^{2}(\Omega_{+})},\quad\forall\,J\geqslant J_{0}, (4.14)

where C1:=C+C02C_{1}:=C_{+}C_{0}^{2}. Similarly, by assumption (2.16), uu_{-} in Ω\Omega_{-} can be extended into a function vH2(Ω)H01(Ω)v_{-}\in H^{2}(\Omega)\cap H_{0}^{1}(\Omega) such that v=uv_{-}=u_{-} in Ω\Omega_{-} and vH2(Ω)C0uH2(Ω)\|v_{-}\|_{H^{2}(\Omega)}\leqslant C_{0}\|u_{-}\|_{H^{2}(\Omega_{-})}. Therefore, by Theorem 4.2 and the same argument, there exists a positive constant C2C_{2} such that for all JJ0J\geqslant J_{0},

j=Jα~jΨ~j2D,supp(α~j)Ω22j|u,α~j|2j=Jα~jΨ~j2D22j|v,α~j|222JC2|u|H2(Ω)2,JJ0.\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{-}\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\leqslant\sum_{j=J}^{\infty}\sum_{\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j}}2^{2j}|\langle v_{-},\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-2J}C_{2}|u_{-}|^{2}_{H^{2}(\Omega_{-})},\quad\forall\,J\geqslant J_{0}. (4.15)

We now handle the solution uu in a neighborhood of the interface Γ\Gamma. Because the closed curve Γ\Gamma is completely inside Ω\Omega, we can assume that there exist two open neighborhoods Ω0\Omega_{0} and Ωρ\Omega_{\rho} of Γ\Gamma such that Ω0ΩρΩ\Omega_{0}\subseteq\Omega_{\rho}\subseteq\Omega, the closure of Ωρ\Omega_{\rho} is contained inside Ω\Omega and the closure of Ω0\Omega_{0} is inside Ωρ\Omega_{\rho}. Since Γ\Gamma is a curve, we can take a compactly supported smooth function ρ\rho supported inside Ωρ\Omega_{\rho} such that ρ=1\rho=1 in Ω0\Omega_{0}. Define a bivariate function w:=ρuw:=\rho u. Obviously, supp(w)Ωρ\operatorname{supp}(w)\subseteq\Omega_{\rho} and hence ww can be regarded as a function in the whole space 2\mathbb{R}^{2} by the zero extension outside Ω\Omega. Therefore, applying Theorem 4.1 with τ1=5/4\tau_{1}=5/4 and τ2=7/4\tau_{2}=7/4, for any τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}], there exists a positive constant C3C_{3}, independent of τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}], such that

η~{ϕ~ψ~,ψ~ϕ~,ψ~ψ~}j=0k222τj|w,η~j,k|223/2C3wHτ(2)2,τ[τ1,τ2]:=[5/4,7/4].\sum_{\tilde{\eta}\in\{\tilde{\phi}\otimes\tilde{\psi},\tilde{\psi}\otimes\tilde{\phi},\tilde{\psi}\otimes\tilde{\psi}\}}\sum_{j=0}^{\infty}\sum_{k\in\mathbb{Z}^{2}}2^{2\tau j}|\langle w,\tilde{\eta}_{j,k}\rangle|^{2}\leqslant 2^{-3/2}C_{3}\|w\|^{2}_{H^{\tau}(\mathbb{R}^{2})},\quad\forall\,\tau\in[\tau_{1},\tau_{2}]:=[5/4,7/4]. (4.16)

It is important to notice that ρ\rho is supported inside Ωρ\Omega_{\rho} and ρ=1\rho=1 in Ω0\Omega_{0}. Take any element

α~jΨ~j2Dandsupp(α~j)Γ\tilde{\alpha}_{j}\in\tilde{\Psi}^{2D}_{j}\quad\mbox{and}\quad\operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset (4.17)

for jJ0j\geqslant J_{0}. Because Γ\Gamma is away from the boundary Ω\partial\Omega and because the support of α~j\tilde{\alpha}_{j} becomes smaller and smaller and closer to the interface Γ\Gamma for jj large enough, any element α~j\tilde{\alpha}_{j} in (4.17) cannot be the boundary wavelets, i.e., we must have α~j=η~j,k:=2jη~(2jk)\tilde{\alpha}_{j}=\tilde{\eta}_{j,k}:=2^{j}\tilde{\eta}(2^{j}\cdot-k) for some k2k\in\mathbb{Z}^{2} and η~{ϕ~ψ~,ψ~ϕ~,ψ~ψ~}\tilde{\eta}\in\{\tilde{\phi}\otimes\tilde{\psi},\tilde{\psi}\otimes\tilde{\phi},\tilde{\psi}\otimes\tilde{\psi}\}. In addition, ρ\rho takes value 11 in Ω0\Omega_{0} and the support of α~j\tilde{\alpha}_{j} will be contained inside Ω0\Omega_{0} for large enough jj. In conclusion, there must exist a positive integer J̊\mathring{J} such that any element α~j\tilde{\alpha}_{j} in (4.17) with jJ̊j\geqslant\mathring{J} must satisfy

supp(α~j)Ω0andα~j=η~j,kfor some k2,\operatorname{supp}(\tilde{\alpha}_{j})\subset\Omega_{0}\quad\mbox{and}\quad\tilde{\alpha}_{j}=\tilde{\eta}_{j,k}\quad\mbox{for some }k\in\mathbb{Z}^{2},

where η~{ϕ~ψ~,ψ~ϕ~,ψ~ψ~}\tilde{\eta}\in\{\tilde{\phi}\otimes\tilde{\psi},\tilde{\psi}\otimes\tilde{\phi},\tilde{\psi}\otimes\tilde{\psi}\}. Moreover, because ρ=1\rho=1 on Ω0\Omega_{0} and supp(α~j)Ω0\operatorname{supp}(\tilde{\alpha}_{j})\subset\Omega_{0}, we have

w,α~j=ρu,α~j=u,α~j=u,η~j,k\langle w,\tilde{\alpha}_{j}\rangle=\langle\rho u,\tilde{\alpha}_{j}\rangle=\langle u,\tilde{\alpha}_{j}\rangle=\langle u,\tilde{\eta}_{j,k}\rangle

for some unique k2k\in\mathbb{Z}^{2}, where we used the definition w=ρuw=\rho u.

For simplicity of discussion, without loss of any generality, we can assume J̊=J0\mathring{J}=J_{0}, because we are only interested in large JJ for proving the convergence rate. Hence, for any α~j\tilde{\alpha}_{j} satisfies (4.17) with jJ0j\geqslant J_{0}, the above discussion implies that for τ[τ1,τ2]:=[5/4,7/4]\tau\in[\tau_{1},\tau_{2}]:=[5/4,7/4] and JJ0J\geqslant J_{0}, we have

j=Jα~jΨ~j2D,supp(α~j)Γ22τj|u,α~j|2\displaystyle\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2\tau j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2} =j=Jα~jΨ~j2D,supp(α~j)Γ22τj|w,α~j|2η~{ϕ~ψ~,ψ~ϕ~,ψ~ψ~}j=0k222τj|w,η~j,k|2.\displaystyle=\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2\tau j}|\langle w,\tilde{\alpha}_{j}\rangle|^{2}\leqslant\sum_{\tilde{\eta}\in\{\tilde{\phi}\otimes\tilde{\psi},\tilde{\psi}\otimes\tilde{\phi},\tilde{\psi}\otimes\tilde{\psi}\}}\sum_{j=0}^{\infty}\sum_{k\in\mathbb{Z}^{2}}2^{2\tau j}|\langle w,\tilde{\eta}_{j,k}\rangle|^{2}.

Now we conclude from the inequality (4.16) and the above estimation that

j=Jα~jΨ~j2D,supp(α~j)Γ22τj|u,α~j|223/2C3wHτ(Ω)2,for all JJ0,τ[τ1,τ2].\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2\tau j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-3/2}C_{3}\|w\|_{H^{\tau}(\Omega)}^{2},\quad\text{for all }J\geqslant J_{0},\tau\in[\tau_{1},\tau_{2}]. (4.18)

In particular, for τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}], we have

j=2J1α~jΨ~j2D,supp(α~j)Γ22j|u,α~j|2=j=2J1α~jΨ~j2D,supp(α~j)Γ22(τ1)j22τj|u,α~j|2\displaystyle\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}=\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{-2(\tau-1)j}2^{2\tau j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}
23/224J(τ1)j=2J1α~jΨ~j2D,supp(α~j)Γ22τj|u,α~j|224J(τ1)C3wHτ(Ω)2,\displaystyle\quad\leqslant 2^{3/2}2^{-4J(\tau-1)}\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2\tau j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-4J(\tau-1)}C_{3}\|w\|_{H^{\tau}(\Omega)}^{2},

where we used 22(τ1)22(τ21)23/22^{2(\tau-1)}\leqslant 2^{2(\tau_{2}-1)}\leqslant 2^{3/2} due to τ2=7/4\tau_{2}=7/4. Consider τ:=3/22ε[5/4,3/2)\tau:=3/2-2\varepsilon\in[5/4,3/2) with 0<ε<1/80<\varepsilon<1/8. Then obviously, τ[τ1,τ2]:=[5/4,7/4]\tau\in[\tau_{1},\tau_{2}]:=[5/4,7/4]. Since 2(τ1)=2(1/22ε)=14ε2(\tau-1)=2(1/2-2\varepsilon)=1-4\varepsilon, the above estimate can be equivalently re-expressed as follows:

j=2J1α~jΨ~j2D,supp(α~j)Γ22j|u,α~j|222J(14ε)C3wH3/22ε(Ω)2,for all 0<ε<1/8.\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-2J(1-4\varepsilon)}C_{3}\|w\|_{H^{3/2-2\varepsilon}(\Omega)}^{2},\quad\text{for all }0<\varepsilon<1/8. (4.19)

In the following, we estimate the quantity wH3/22ε(Ω)\|w\|_{H^{3/2-2\varepsilon}(\Omega)} for 0<ε<1/80<\varepsilon<1/8, specifically for ε0+\varepsilon\to 0^{+}. Define w+:=wχΩ+=ρu+w_{+}:=w\chi_{\Omega_{+}}=\rho u_{+} and w:=wχΩ=ρuw_{-}:=w\chi_{\Omega_{-}}=\rho u_{-}. Then w+H2(Ω+)w_{+}\in H^{2}(\Omega_{+}) and wH2(Ω+)w_{-}\in H^{2}(\Omega_{+}). Moreover, ρv+\rho v_{+} and ρv\rho v_{-} are extensions of w+w_{+} and ww_{-}, respectively. Because w=ρuH01(Ω)w=\rho u\in H_{0}^{1}(\Omega), we consider w\nabla w. To estimate wH3/22ε(Ω)\|w\|_{H^{3/2-2\varepsilon}(\Omega)} for 0<ε<1/80<\varepsilon<1/8, it suffices to estimate wH1/22ε(Ω)\|\nabla w\|_{H^{1/2-2\varepsilon}(\Omega)}. For simplicity of discussion, we only handle xw\frac{\partial}{\partial x}w and we assume that Ω\Omega_{-} is inside Ω\Omega and ΩΩ=\partial\Omega_{-}\cap\partial\Omega=\emptyset. Because w=ρuH01(Ω)w=\rho u\in H_{0}^{1}(\Omega), we have xwL2(Ω)\frac{\partial}{\partial x}w\in L_{2}(\Omega). Noting that ρv+H2(2)\rho v_{+}\in H^{2}(\mathbb{R}^{2}), we can rewrite

wx:=xw=x[ρv+]χΩ++x[ρv]χΩ=x[ρv+]+FχΩwithF:=x[ρv]x[ρv+],w_{x}:=\frac{\partial}{\partial x}w=\frac{\partial}{\partial x}[\rho v_{+}]\chi_{\Omega_{+}}+\frac{\partial}{\partial x}[\rho v_{-}]\chi_{\Omega_{-}}=\frac{\partial}{\partial x}[\rho v_{+}]+F\chi_{\Omega_{-}}\quad\mbox{with}\quad F:=\frac{\partial}{\partial x}[\rho v_{-}]-\frac{\partial}{\partial x}[\rho v_{+}],

because w=ρu+=ρv+w=\rho u_{+}=\rho v_{+} in Ω+\Omega_{+} and w=ρu=ρvw=\rho u_{-}=\rho v_{-} in Ω\Omega_{-}. Note that FH1(2)F\in H^{1}(\mathbb{R}^{2}) and FF has compact support by v+,vH2(Ω)H01(Ω)v_{+},v_{-}\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega). Consequently, there exists a positive constant C4C_{4} such that

FH1(2)C4(u+H2(Ω+)+uH2(Ω)),\|F\|_{H^{1}(\mathbb{R}^{2})}\leqslant C_{4}(\|u_{+}\|_{H^{2}(\Omega_{+})}+\|u_{-}\|_{H^{2}(\Omega_{-})}), (4.20)

where C4C_{4} only depends on the smooth function ρ\rho and the positive constant C0C_{0} appeared in v+H2(Ω)C0u+H2(Ω)\|v_{+}\|_{H^{2}(\Omega)}\leqslant C_{0}\|u_{+}\|_{H^{2}(\Omega)} and vH2(Ω)C0uH2(Ω)\|v_{-}\|_{H^{2}(\Omega)}\leqslant C_{0}\|u_{-}\|_{H^{2}(\Omega)}. We still consider τ:=3/22ε\tau:=3/2-2\varepsilon with 0<ε<1/80<\varepsilon<1/8. Then τε:=τ1=1/22ε[1/4,1/2)\tau_{\varepsilon}:=\tau-1=1/2-2\varepsilon\in[1/4,1/2). Hence, by wx=x[ρv+]+FχΩw_{x}=\frac{\partial}{\partial x}[\rho v_{+}]+F\chi_{\Omega_{-}}, we have

wxHτε(2)[ρv+]xHτε(2)+FχΩHτε(2)CρC0u+H2(Ω)+FχΩHτε(2),\|w_{x}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}\leqslant\|[\rho v_{+}]_{x}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}+\|F\chi_{\Omega_{-}}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}\leqslant C_{\rho}C_{0}\|u_{+}\|_{H^{2}(\Omega)}+\|F\chi_{\Omega_{-}}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}, (4.21)

where τε[1/4,1/2]\tau_{\varepsilon}\in[1/4,1/2] holds and we used [ρv+]xHτε(2)=[ρv+]xHτε(Ω)Cρv+H2(Ω)\|[\rho v_{+}]_{x}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}=\|[\rho v_{+}]_{x}\|_{H^{\tau_{\varepsilon}}(\Omega)}\leqslant C_{\rho}\|v_{+}\|_{H^{2}(\Omega)}, where the positive constant CρC_{\rho} only depends on ρ\rho and we used the inequality v+H2(Ω)C0u+H2(Ω)\|v_{+}\|_{H^{2}(\Omega)}\leqslant C_{0}\|u_{+}\|_{H^{2}(\Omega)}.

Next, we estimate FχΩHτε(2)\|F\chi_{\Omega_{-}}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}. Since FH1(2)F\in H^{1}(\mathbb{R}^{2}), by [3, Theorems C.9 and C.10] with r=τε,s=τε+ε,t=1r=\tau_{\varepsilon},s=\tau_{\varepsilon}+\varepsilon,t=1 and d=2d=2, there exists a positive constant C5C_{5} only depending on Ω\Omega_{-} such that

FχΩHτε(2)C5ε1/2FH1(2)χΩHτε+ε(2),\|F\chi_{\Omega_{-}}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}\leqslant C_{5}\varepsilon^{-1/2}\|F\|_{H^{1}(\mathbb{R}^{2})}\|\chi_{\Omega_{-}}\|_{H^{\tau_{\varepsilon}+\varepsilon}(\mathbb{R}^{2})}, (4.22)

where the above factor ε1/2\varepsilon^{-1/2} is from p=022p(s+td/2r)=p=022pε=1122ε1ε2ln2\sum_{p=0}^{\infty}2^{-2p(s+t-d/2-r)}=\sum_{p=0}^{\infty}2^{-2p\varepsilon}=\frac{1}{1-2^{-2\varepsilon}}\leqslant\frac{1}{\varepsilon\sqrt{2}\ln 2} for all 0<ε<1/80<\varepsilon<1/8 in [3, Proof of Theorem C.10] by noting s+td/2r=ε>0s+t-d/2-r=\varepsilon>0. Noting τε+ε=τ1+ε=1/2ε\tau_{\varepsilon}+\varepsilon=\tau-1+\varepsilon=1/2-\varepsilon, we obtain from (4.9) in Lemma 4.3 that χΩHτε+ε(2)CΓε1/2\|\chi_{\Omega}\|_{H^{\tau_{\varepsilon}+\varepsilon}(\mathbb{R}^{2})}\leqslant C_{\Gamma}\varepsilon^{-1/2}. Combining (4.20), (4.21) and (4.22), we obtain

wxHτε(2)CρC0u+H2(Ω)+C4C5CΓ(u+H2(Ω+)+uH2(Ω))ε1.\|w_{x}\|_{H^{\tau_{\varepsilon}}(\mathbb{R}^{2})}\leqslant C_{\rho}C_{0}\|u_{+}\|_{H^{2}(\Omega)}+C_{4}C_{5}C_{\Gamma}(\|u_{+}\|_{H^{2}(\Omega_{+})}+\|u_{-}\|_{H^{2}(\Omega_{-})})\varepsilon^{-1}.

An estimate for wyHτε(Ω)\|w_{y}\|_{H^{\tau_{\varepsilon}}(\Omega)} can be proved similarly. Note that wH3/22ε(Ω)2=wH1(Ω)2+wHτε(Ω)2\|w\|_{H^{3/2-2\varepsilon}(\Omega)}^{2}=\|w\|_{H^{1}(\Omega)}^{2}+\|\nabla w\|_{H^{\tau_{\varepsilon}}(\Omega)}^{2} by τε=τ1=1/22ε\tau_{\varepsilon}=\tau-1=1/2-2\varepsilon. Noting that w=ρuH01(Ω)w=\rho u\in H^{1}_{0}(\Omega) and

wH1(Ω)2=ρu+H1(Ω+)2+ρuH1(Ω)2Cρ[u+H2(Ω+)2+uH2(Ω)2],\|w\|_{H^{1}(\Omega)}^{2}=\|\rho u_{+}\|_{H^{1}(\Omega_{+})}^{2}+\|\rho u_{-}\|_{H^{1}(\Omega_{-})}^{2}\leqslant C_{\rho}[\|u_{+}\|^{2}_{H^{2}(\Omega_{+})}+\|u_{-}\|^{2}_{H^{2}(\Omega_{-})}],

we conclude from the above inequality estimating wxHτϵ(2)\|w_{x}\|_{H^{\tau_{\epsilon}}(\mathbb{R}^{2})} and similarly wyHτϵ(2)\|w_{y}\|_{H^{\tau_{\epsilon}}(\mathbb{R}^{2})} that

wH3/22ε(Ω)2C6ε2(u+H2(Ω+)2+uH2(Ω)2), 0<ε<1/8,\|w\|_{H^{3/2-2\varepsilon}(\Omega)}^{2}\leqslant C_{6}\varepsilon^{-2}(\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{-})}^{2}),\qquad\forall\;0<\varepsilon<1/8, (4.23)

where C6:=Cρ+2(CρC0+C4C5CΓ)2+2C42C52CΓ2<C_{6}:=C_{\rho}+2(C_{\rho}C_{0}+C_{4}C_{5}C_{\Gamma})^{2}+2C_{4}^{2}C_{5}^{2}C_{\Gamma}^{2}<\infty. Since τ=3/22ε\tau=3/2-2\varepsilon for 0<ε<1/80<\varepsilon<1/8, we deduce from (4.19) and (4.23) that τ1=1/22ε\tau-1=1/2-2\varepsilon, 2J(14ε)=4J(τ1)-2J(1-4\varepsilon)=-4J(\tau-1) and

j=2J1α~jΨ~j2D,supp(α~j)Γ22j|u,α~j|224J(τ1)C3wHτ(2)2C3C622J(H(ε))2(u+H2(Ω+)2+uH2(Ω+)2),\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\leqslant 2^{-4J(\tau-1)}C_{3}\|w\|_{H^{\tau}(\mathbb{R}^{2})}^{2}\leqslant C_{3}C_{6}2^{-2J}(H(\varepsilon))^{-2}(\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{+})}^{2}), (4.24)

where H(ε):=ε24JεH(\varepsilon):=\varepsilon 2^{-4J\varepsilon}, which can be written as H(ε)=εh4εH(\varepsilon)=\varepsilon h^{4\varepsilon} with h:=2Jh:=2^{-J}. Note that

H(ε)=h4ε+4εh4εlog(h)=h4ε(1+4εlog(h)),H^{\prime}(\varepsilon)=h^{4\varepsilon}+4\varepsilon h^{4\varepsilon}\log(h)=h^{4\varepsilon}(1+4\varepsilon\log(h)),

where log\log is the natural logarithm. Setting H(ε)=0H^{\prime}(\varepsilon)=0 gives ε=14log(h1)>0\varepsilon=\frac{1}{4\log(h^{-1})}>0, i.e., ε=14|log(h)|1\varepsilon=\frac{1}{4}|\log(h)|^{-1} because 0<h<10<h<1. Taking ε=14|log(h)|1\varepsilon=\frac{1}{4}|\log(h)|^{-1} in (4.24), we conclude that H(ε)=14|log(h)|1e1H(\varepsilon)=\frac{1}{4}|\log(h)|^{-1}e^{-1}, (H(ε))2=16e2|log(h)|2(H(\varepsilon))^{-2}=16e^{2}|\log(h)|^{2}, and finally we deduce from (4.24) that

j=2J1α~jΨ~j2D,supp(α~j)Γ22j|u,α~j|2Cv|log(h)|222J(u+H2(Ω+)2+uH2(Ω+)2),\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\tilde{\alpha}_{j}\in\tilde{\Psi}_{j}^{2D},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\leqslant C_{v}|\log(h)|^{2}2^{-2J}(\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{+})}^{2}), (4.25)

where Cv:=16e2C3C6<C_{v}:=16e^{2}C_{3}C_{6}<\infty.

Since uH01(Ω)u\in H^{1}_{0}(\Omega), we have the following wavelet expansion

u=αΦJ02Du,α~α+j=J0αjΨj2Du,α~jαj=αΦJ02Du,[2jα~]2jα+j=J0αjΨj2Du,[2jα~j]2jαj.\displaystyle u=\sum_{\alpha\in\Phi^{2D}_{J_{0}}}\langle u,\tilde{\alpha}\rangle\alpha+\sum_{j=J_{0}}^{\infty}\sum_{\alpha_{j}\in\Psi^{2D}_{j}}\langle u,\tilde{\alpha}_{j}\rangle\alpha_{j}=\sum_{\alpha\in\Phi^{2D}_{J_{0}}}\langle u,[2^{j}\tilde{\alpha}]\rangle 2^{-j}\alpha+\sum_{j=J_{0}}^{\infty}\sum_{\alpha_{j}\in\Psi^{2D}_{j}}\langle u,[2^{j}\tilde{\alpha}_{j}]\rangle 2^{-j}\alpha_{j}.

We define

I1\displaystyle I_{1} :=j=JαjΨj2D,supp(α~j)Ω+u,[2jα~j][2jαj],I2:=j=JαjΨj2D,supp(α~j)Ωu,[2jα~j][2jαj],\displaystyle:=\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{+}\end{subarray}}\langle u,[2^{j}\tilde{\alpha}_{j}]\rangle[2^{-j}\alpha_{j}],\qquad I_{2}:=\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{-}\end{subarray}}\langle u,[2^{j}\tilde{\alpha}_{j}]\rangle[2^{-j}\alpha_{j}],
I3\displaystyle I_{3} :=j=2J1αjΨj,supp(α~j)Γu,[2jα~j][2jαj],andůh:=αJ02Du,α~αI1I2I3.\displaystyle:=\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}\langle u,[2^{j}\tilde{\alpha}_{j}]\rangle[2^{-j}\alpha_{j}],\quad\text{and}\quad\mathring{u}_{h}:=\sum_{\alpha\in\mathcal{B}_{J_{0}}^{2D}}\langle u,\tilde{\alpha}\rangle\alpha-I_{1}-I_{2}-I_{3}.

Recall that Vhwav:=span(J0,JS,H1(Ω))V_{h}^{wav}:=\mbox{span}(\mathcal{B}_{J_{0},J}^{S,H^{1}(\Omega)}). Then obviously,

uůJ=I1+I2+I3andůhVhwav.u-\mathring{u}_{J}=I_{1}+I_{2}+I_{3}\quad\mbox{and}\quad\mathring{u}_{h}\in V_{h}^{wav}.

Because J0S,H1(Ω)\mathcal{B}_{J_{0}}^{S,H^{1}(\Omega)} is a Riesz basis of H01(Ω)H^{1}_{0}(\Omega), we deduce from uůh=I1+I2+I3u-\mathring{u}_{h}=I_{1}+I_{2}+I_{3} that there must exist a positive constant C7C_{7}, only depending on the wavelet basis J0S,H1(Ω)\mathcal{B}_{J_{0}}^{S,H^{1}(\Omega)}, such that

uůhH1(Ω)2=I1+I2+I3H1(Ω)2\displaystyle\|u-\mathring{u}_{h}\|^{2}_{H^{1}(\Omega)}=\|I_{1}+I_{2}+I_{3}\|^{2}_{H^{1}(\Omega)}
C7(j=JαjΨj2D,supp(α~j)Ω+|u,[2jα~j]|2+j=JαjΨj2D,supp(α~j)Ω|u,[2jα~j]|2+j=2J1αjΨj2D,supp(α~j)Γ|u,[2jα~j]|2)\displaystyle\leqslant C_{7}\left(\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{+}\end{subarray}}|\langle u,[2^{j}\tilde{\alpha}_{j}]\rangle|^{2}+\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{-}\end{subarray}}|\langle u,[2^{j}\tilde{\alpha}_{j}]\rangle|^{2}+\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}|\langle u,[2^{j}\tilde{\alpha}_{j}]\rangle|^{2}\right)
=C7(j=JαjΨj2D,supp(α~j)Ω+22j|u,α~j|2+j=JαjΨj2D,supp(α~j)Ω22j|u,α~j|2+j=2J1αjΨj2D,supp(α~j)Γ22j|u,α~j|2).\displaystyle=C_{7}\left(\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{+}\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}+\sum_{j=J}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\subseteq\Omega_{-}\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}+\sum_{j=2J-1}^{\infty}\sum_{\begin{subarray}{c}\alpha_{j}\in\Psi^{2D}_{j},\\ \operatorname{supp}(\tilde{\alpha}_{j})\cap\Gamma\neq\emptyset\end{subarray}}2^{2j}|\langle u,\tilde{\alpha}_{j}\rangle|^{2}\right).

By (4.14), (4.15), and (4.25), we have

uůhH1(Ω)2\displaystyle\|u-\mathring{u}_{h}\|^{2}_{H^{1}(\Omega)} C7[u+H2(Ω+)2+uH2(Ω+)2](22JC1+22JC2+22J|log(h)|2Cv)\displaystyle\leqslant C_{7}[\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{+})}^{2}](2^{-2J}C_{1}+2^{-2J}C_{2}+2^{-2J}|\log(h)|^{2}C_{v})
C82[u+H2(Ω+)2+uH2(Ω+)2]22J|log(h)|2\displaystyle\leqslant C^{2}_{8}\left[\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{+})}^{2}\right]2^{-2J}|\log(h)|^{2}
=C82[u+H2(Ω+)2+uH2(Ω+)2](h|log(h)|)2,\displaystyle=C_{8}^{2}\left[\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{+})}^{2}\right](h|\log(h)|)^{2},

where C82:=C7(C1+C2+Cv)<C_{8}^{2}:=C_{7}(C_{1}+C_{2}+C_{v})<\infty. This proves

uůhH1(Ω)C8h|log(h)|u+H2(Ω+)2+uH2(Ω+)2.\|u-\mathring{u}_{h}\|_{H^{1}(\Omega)}\leqslant C_{8}h|\log(h)|\sqrt{\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{+})}^{2}}. (4.26)

By the Cea’s lemma, there exists a positive constant CaC_{a}, only depends on the diffusion coefficient aa and Ω\Omega, such that

uuhH1(Ω)CainfvVhwavuvH1(Ω).\|u-u_{h}\|_{H^{1}(\Omega)}\leqslant C_{a}\inf_{v\in V_{h}^{wav}}\|u-v\|_{H^{1}(\Omega)}.

Because ůhVhwav\mathring{u}_{h}\in V_{h}^{wav}, we conclude that

uuhH1(Ω)CainfvVhwavuvH1(Ω)CauůhH1(Ω)\|u-u_{h}\|_{H^{1}(\Omega)}\leqslant C_{a}\inf_{v\in V_{h}^{wav}}\|u-v\|_{H^{1}(\Omega)}\leqslant C_{a}\|u-\mathring{u}_{h}\|_{H^{1}(\Omega)}

and consequently,

uuhH1(Ω)2C9h2|log(h)|2(u+H2(Ω+)2+uH2(Ω+)2),\|u-u_{h}\|^{2}_{H^{1}(\Omega)}\leqslant C_{9}h^{2}|\log(h)|^{2}(\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{+})}^{2}), (4.27)

where C9:=C82Ca2C_{9}:=C_{8}^{2}C_{a}^{2}. This proves the first inequality in (2.17) for convergence in H01(Ω)H_{0}^{1}(\Omega). Because NJ=𝒪(h2)N_{J}=\mathscr{O}(h^{-2}), the second inequality in (2.17) follows.

Proving the L2(Ω)L_{2}(\Omega) convergence. We now use the Aubin-Nitsche’s technique to prove (2.18) for L2(Ω)L_{2}(\Omega) convergence. Note that the bilinear form a(,)a(\cdot,\cdot) is symmetric. Suppose that wH01(Ω)w\in H_{0}^{1}(\Omega) satisfies

a(w,v)=uuh,v,vH01(Ω),a(w,v)=\langle u-u_{h},v\rangle,\qquad v\in H_{0}^{1}(\Omega), (4.28)

and its wavelet approximated solution whVhwavw_{h}\in V_{h}^{wav} satisfies

a(wh,vh)=uuh,vh,vhVhwav.a(w_{h},v_{h})=\langle u-u_{h},v_{h}\rangle,\qquad v_{h}\in V_{h}^{wav}.

By the same proof of the inequality of (4.27), we have

(wwh)L2(Ω)2C9h2|log(h)|2(w+H2(Ω+)2+wH2(Ω)2)\|\nabla(w-w_{h})\|^{2}_{L_{2}(\Omega)}\leqslant C_{9}h^{2}|\log(h)|^{2}(\|w_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|w_{-}\|_{H^{2}(\Omega_{-})}^{2})

for some positive constant C9C_{9}. Because g=0g=0 and gΓ=0g_{\Gamma}=0 in the weak formulation (4.28), [33] (also see [12, Theorem 2.1]) guarantees the existence of a positive constant C10C_{10} such that

w+H2(Ω+)2+wH2(Ω)2C10uuhL2(Ω)2,\|w_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|w_{-}\|_{H^{2}(\Omega_{-})}^{2}\leqslant C_{10}\|u-u_{h}\|^{2}_{L_{2}(\Omega)},

where uuhu-u_{h} is treated as the source term for the solution ww in (4.28). Therefore,

(wwh)L2(Ω)2C9h2|log(h)|2(w+H2(Ω+)2+wH2(Ω)2)C9C10h2|log(h)|2uuhL2(Ω)2.\|\nabla(w-w_{h})\|^{2}_{L_{2}(\Omega)}\leqslant C_{9}h^{2}|\log(h)|^{2}(\|w_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|w_{-}\|_{H^{2}(\Omega_{-})}^{2})\leqslant C_{9}C_{10}h^{2}|\log(h)|^{2}\|u-u_{h}\|^{2}_{L_{2}(\Omega)}. (4.29)

Since v=uuhH01(Ω)v=u-u_{h}\in H^{1}_{0}(\Omega), we deduce from a(w,v)=uuh,va(w,v)=\langle u-u_{h},v\rangle that

uuhL2(Ω)2=a(w,uuh)=a(wwh,uuh),\|u-u_{h}\|^{2}_{L_{2}(\Omega)}=a(w,u-u_{h})=a(w-w_{h},u-u_{h}),

where we used the Galerkin orthogonality a(wh,uuh)=a(uuh,wh)=0a(w_{h},u-u_{h})=a(u-u_{h},w_{h})=0 for whVhwavw_{h}\in V_{h}^{wav}. Consequently, we deduce from (4.27) and (4.29) that

uuhL2(Ω)2\displaystyle\|u-u_{h}\|^{2}_{L_{2}(\Omega)} =a(wwh,uuh)C11(wwh)L2(Ω)(uuh)L2(Ω)\displaystyle=a(w-w_{h},u-u_{h})\leqslant C_{11}\|\nabla(w-w_{h})\|_{L_{2}(\Omega)}\|\nabla(u-u_{h})\|_{L_{2}(\Omega)}
h|log(h)|C11C9C10uuhL2(Ω)C9h|log(h)|(u+H2(Ω+)2+uH2(Ω)2)1/2\displaystyle\leqslant h|\log(h)|C_{11}\sqrt{C_{9}C_{10}}\|u-u_{h}\|_{L_{2}(\Omega)}\sqrt{C_{9}}h|\log(h)|(\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{-})}^{2})^{1/2}
=Ch2|log(h)|2(u+H2(Ω+)2+uH2(Ω)2)1/2uuhL2(Ω),\displaystyle=Ch^{2}|\log(h)|^{2}(\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{-})}^{2})^{1/2}\|u-u_{h}\|_{L_{2}(\Omega)},

where C:=C11C9C10C:=C_{11}C_{9}\sqrt{C_{10}}, from which we conclude that the first inequality of (2.18) holds, i.e.,

uuhL2(Ω)Ch2|log(h)|2(u+H2(Ω+)2+uH2(Ω)2)1/2.\|u-u_{h}\|_{L_{2}(\Omega)}\leqslant Ch^{2}|\log(h)|^{2}(\|u_{+}\|_{H^{2}(\Omega_{+})}^{2}+\|u_{-}\|_{H^{2}(\Omega_{-})}^{2})^{1/2}.

The second inequality of (2.18) follows trivially by noting #NJ=𝒪(h2)\#N_{J}=\mathscr{O}(h^{-2}).

Proving that the condition number is uniformly bounded. Take vhVhwavv_{h}\in V^{wav}_{h}. Then, vh=ηJ0,JS,H01(Ω)cηηv_{h}=\sum_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}c_{\eta}\eta. We want to find an upper bound for a(vh,vh)a(v_{h},v_{h}). Note that

a(vh,vh)aL(Ω)vhL2(Ω)2aL(Ω)(vhL2(Ω)2+vhL2(Ω)2)C,2aL(Ω)ηJ0,JS,H1(Ω)|cη|2,a(v_{h},v_{h})\leqslant\|a\|_{L_{\infty}(\Omega)}\|\nabla v_{h}\|_{L_{2}(\Omega)}^{2}\leqslant\|a\|_{L_{\infty}(\Omega)}\left(\|v_{h}\|_{L_{2}(\Omega)}^{2}+\|\nabla v_{h}\|_{L_{2}(\Omega)}^{2}\right)\leqslant C_{\mathcal{B},2}\|a\|_{L_{\infty}(\Omega)}\sum_{\eta\in\mathcal{B}^{S,H_{1}(\Omega)}_{J_{0},J}}|c_{\eta}|^{2},

where we used the fact that J0H1(Ω)\mathcal{B}^{H_{1}(\Omega)}_{J_{0}} is a Riesz basis of the Sobolev space H01(Ω)H^{1}_{0}(\Omega) to arrive at the final inequality. Since vhv_{h} satisfies the zero Dirichlet boundary condition, by the Poincaré inequality, we have vhL2(Ω)CPvhL2(Ω)\|v_{h}\|_{L_{2}(\Omega)}\leqslant C_{P}\|\nabla v_{h}\|_{L_{2}(\Omega)} with CPC_{P} being a positive constant that depends only on Ω\Omega, which implies that

vhL2(Ω)2+vhL2(Ω)2(1+CP2)vhL2(Ω)2.\|v_{h}\|_{L_{2}(\Omega)}^{2}+\|\nabla v_{h}\|_{L_{2}(\Omega)}^{2}\leqslant(1+C_{P}^{2})\|\nabla v_{h}\|_{L_{2}(\Omega)}^{2}.

Moreover, we have

a(vh,vh)\displaystyle a(v_{h},v_{h}) a1L(Ω)1vhL2(Ω)2a1L(Ω)1(1+CP2)1(vhL2(Ω)2+vhL2(Ω)2)\displaystyle\geqslant\|a^{-1}\|_{L_{\infty}(\Omega)}^{-1}\|\nabla v_{h}\|_{L_{2}(\Omega)}^{2}\geqslant\|a^{-1}\|_{L_{\infty}(\Omega)}^{-1}(1+C_{P}^{2})^{-1}(\|v_{h}\|_{L_{2}(\Omega)}^{2}+\|\nabla v_{h}\|_{L_{2}(\Omega)}^{2})
C,1a1L(Ω)1(1+CP2)1ηJ0,JS,H01(Ω)|cη|2,\displaystyle\geqslant C_{\mathcal{B},1}\|a^{-1}\|_{L_{\infty}(\Omega)}^{-1}(1+C_{P}^{2})^{-1}\sum_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}|c_{\eta}|^{2},

where we used the fact that J0H01(Ω)\mathcal{B}^{H^{1}_{0}(\Omega)}_{J_{0}} is a Riesz basis of the Sobolev space H01(Ω)H^{1}_{0}(\Omega) to arrive at the final inequality. Combining the lower and upper bounds of a(vh,vh)a(v_{h},v_{h}), we have

C,1a1L(Ω)1(1+CP2)1ηJ0,JS,H01(Ω)|cη|2a(vh,vh)C,2aL(Ω)ηJ0,JS,H01(Ω)|cη|2,C_{\mathcal{B},1}\|a^{-1}\|_{L_{\infty}(\Omega)}^{-1}(1+C_{P}^{2})^{-1}\sum_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}|c_{\eta}|^{2}\leqslant a(v_{h},v_{h})\leqslant C_{\mathcal{B},2}\|a\|_{L_{\infty}(\Omega)}\sum_{\eta\in\mathcal{B}^{S,H^{1}_{0}(\Omega)}_{J_{0},J}}|c_{\eta}|^{2},

which gives an upper bound of the condition number in the form of CwaL(Ω)a1L(Ω)C_{w}\|a\|_{L_{\infty}(\Omega)}\|a^{-1}\|_{L_{\infty}(\Omega)}, where Cw:=(1+CP2)C,2C,11<C_{w}:=(1+C_{P}^{2})C_{\mathcal{B},2}C_{\mathcal{B},1}^{-1}<\infty. ∎

References

  • [1] S. Adjerid, I. Babuska, R. Guo, and T. Lin, An enriched immersed finite element method for interface problems with nonhomogeneous jump conditions. Comput. Methods Appl. Mech. Engrg. 404 (2023), 115770.
  • [2] I. Babuška, U. Banerjee, and K. Kergrene, Strongly stable generalized finite element method: Application to interface problems. Comput. Methods Appl. Mech. Engrg. 327 (2017), 58-92.
  • [3] S. Benzoni-Gavage and D. Serre, Multidimensional hyperbolic partial differential equations. Oxford Math. Monogr. Oxford University Press, Oxford, 2007, xxvi+508 pp.
  • [4] S. C. Brenner and L. R. Scott, The mathematical theory of finite element methods. Texts in Applied Mathematics. Springer, New York, 2008. xvii + 397 pp.
  • [5] E. Burman and A. Ern, An unfitted hybrid high-order method for elliptic interface problems. SIAM J. Numer. Anal. 56 (2018), no. 3, 1525–1546.
  • [6] E. Burman, S. Claus, P. Hansbo, M. G. Larson, and André Massing, CutFEM: Discretizing geometry and partial differential equations. Int. J. Numer. Meth. Engng. 104 (2015), 472-501.
  • [7] Z. Cai, X. Ye, and S. Zhang, Discontinuous Galerkin finite element methods for interface problems: a priori and a posteriori error estimations. SIAM J. Numer. Anal. 49 (2011), no. 5, 1761-1787.
  • [8] D. Černá, Wavelets on the interval and their applications, Habilitation thesis at Masaryk University, (2019).
  • [9] L. Chen, H. Wei, and M. Wen, An interface-fitted mesh generator and virtual element methods for elliptic interface problems. J. Comput. Phys. 334 (2017), 327-348.
  • [10] Z. Chen and Y. Liu, An arbitrarily high order unfitted finite element method for elliptic interface problems with automatic mesh generation. J. Comput. Phys. 491 (2023), 112384.
  • [11] Z. Chen, K. Li, and X. Xiang, An adaptive high-order unfitted finite element method for elliptic interface problems, Numer. Math. 149 (2021), 507-548.
  • [12] Z. Chen and J. Zou, Finite element methods and their convergence for elliptic and parabolic interface problems. Numer. Math. 79 (1998), 175-202.
  • [13] A. Cohen, Numerical analysis of wavelet methods, Stud. Math. Appl. 32 (2003), Amsterdam.
  • [14] W. Dahmen, Wavelet and multiscale methods for operator equations. Acta Numer., 6, Cambridge University Press, Cambridge, 1997, 55–228.
  • [15] W. Dahmen, A. Kunoth and K. Urban, Biorthogonal spline wavelets on the interval—stability and moment conditions. Appl. Comput. Harmon. Anal. 6 (1999), 132–196.
  • [16] Q. Feng, B. Han, and M. Michelle, Sixth-order compact finite difference method for 2D Helmholtz equations with singular sources and reduced pollution effect. Commun. Comput. Phys. 34 (2023), 672-712.
  • [17] Q. Feng, B. Han, and P. Minev, A high order compact finite difference scheme for elliptic interface problems with discontinuous and high-contrast coefficients. Appl. Math. Comput. 431 (2022), 127314.
  • [18] Q. Feng, B. Han, and P. Minev, Sixth-order hybrid finite difference methods for elliptic interface problems with mixed boundary conditions. J. Comput. Phys. 497 (2024), 112635.
  • [19] H. Feng and S. Zhao, A fourth order finite difference method for solving elliptic interface problems with the FFT acceleration, J. Comput. Phys. 419 (2020), 109677.
  • [20] Y. Gong, B. Li, and Z. Li, Immersed-interface finite-element methods for elliptic interface problems with nonhomogeneous jump conditions. SIAM J. Numer. Anal. 46 (2008), no. 1, 472-495.
  • [21] R. Guo and T. Lin, A group of immersed finite-element spaces for elliptic interface problems. IMA J. Numer. Anal. 39 (2019), 482-511.
  • [22] B. Han, Compactly supported tight wavelet frames and orthonormal wavelets of exponential decay with a general dilation matrix. J. Comput. Appl. Math. 155 (2003), 43–67.
  • [23] B. Han, Nonhomogeneous wavelet systems in high dimensions. Appl. Comput. Harmon. Anal. 32 (2012), 169–196.
  • [24] B. Han, Framelets and wavelets: Algorithms, analysis, and applications. Applied and Numerical Harmonic Analysis. Birkhäuser/Springer, Cham, 2017. xxxiii + 724 pp.
  • [25] B. Han, Interpolating refinable functions and nsn_{s}-step interpolatory subdivision schemes. Adv. Comput. Math. (2024), 50–98.
  • [26] B. Han and M. Michelle, Derivative-orthogonal Riesz wavelets in Sobolev spaces with applications to differential equations. Appl. Comp. Harmon. Anal. 47 (2019), no. 3, 759-794.
  • [27] B. Han and M. Michelle, Wavelets on intervals derived from arbitrary compactly supported biorthogonal multiwavelets. Appl. Comp. Harmon. Anal. 53 (2021), 270-331.
  • [28] B. Han and M. Michelle, Wavelet Galerkin method for an electromagnetic scattering problem. https://arxiv.org/abs/2303.06770 (2023), 24 pages.
  • [29] B. Han and Z. Shen, Dual wavelet frames and Riesz bases in Sobolev spaces. Constr. Approx. 29 (2009), 369-406.
  • [30] A. Hansbo and P. Hansbo, An unfitted finite element method, based on Nitsche’s method, for elliptic interface problems. Comput. Methods Appl. Mech. Engrg. 191 (2002), 5537–5552.
  • [31] H. Ji, F. Wang, J. Chen, and Z. Li, A new parameter free partially penalized immersed finite element and optimal convergence analysis. Numer. Math. 150 (2022), 1035-1086.
  • [32] K. Kergrene, I. Babuška, and U. Banerjee, Stable generalized finite element method and associated iterative schemes; application to interface problems. Comput. Methods Appl. Mech. Engrg. 305 (2016), 1-36.
  • [33] O. A. Ladyženskaja and N. N. Ural’tseva, Linear and quasilinear elliptic equations. Academic Press, New York-London, 1968, xviii+495 pp.
  • [34] R. J. Leveque and Z. Li, The immersed interface method for elliptic equations with discontinuous coefficients and singular sources. SIAM J. Numer. Anal. 31 (1994), no. 4, 1019–1044.
  • [35] Z. Li, T. Lin, and X. Wu, New Cartesian grid methods for interface problems using the finite element formulation. Numer. Math. 96 (2003), 61-98.
  • [36] Z. Li and K. Ito, The immersed interface method: Numerical solutions of PDEs involving interfaces and irregular domains. Society for Industrial and Applied Mathematics. Philadelphia, 2006. xvi + 332 pp.
  • [37] T. Lin, Y. Lin, and X. Zhang, Partially penalized immersed finite element methods for elliptic interface problems, SIAM J. Numer. Anal. 53 (2015), 1121–1144.
  • [38] R. Massjung, An unfitted discontinuous Galerkin method applied to elliptic interface problems. SIAM J. Numer. Anal. 50 (2012), no. 6, 3134-3162.
  • [39] B. L. Vaughan, B. G. Smith, and D. L. Chopp, A comparison of the extended finite element method with the immersed interface method for elliptic equations with discontinuous coefficients and singular sources. Comm. App. Math. and Comp. Sci. 1 (2006), no. 1, 207-228.
  • [40] Q. Zhang and I. Babuška, A stable generalized finite element method (SGFEM) of degree two for interface problems. Comput. Methods Appl. Mech. Engrg. 363 (2020), 112889.
  • [41] Q. Zhang, C. Cui, U. Banerjee, and I. Babuška, A condensed generalized finite element methods (CGFEM) for interface problems. Comput. Methods Appl. Mech. Engrg. 391 (2022), 114537.
  • [42] Y. C. Zhou, S. Zhao, M. Feig, and G. W. Wei, High order matched interface and boundary method for elliptic equations with discontinuous coefficients and singular sources. J. Comput. Phys. 213 (2006), no. 213, 1–30.