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

An Immersed Weak Galerkin Method for Elliptic Interface Problems on Polygonal Meshes

Hyeokjoo Park and Do Y. Kwak
Abstract.

In this paper we present an immersed weak Galerkin method for solving second-order elliptic interface problems on polygonal meshes, where the meshes do not need to be aligned with the interface. The discrete space consists of constants on each edge and broken linear polynomials satisfying the interface conditions in each element. For triangular meshes, such broken linear plynomials coincide with the basis functions in immersed finite element methods [26]. We establish some approximation properties of the broken linear polynomials and the discrete weak gradient of a certain projection of the solution on polygonal meshes. We then prove an optimal error estimate of our scheme in the discrete H1H^{1}-seminorm under some assumptions on the exact solution. Numerical experiments are provided to confirm our theoretical analysis.

Key words and phrases:
immersed weak Galerkin method, elliptic interface problem, unfitted mesh, polygonal mesh
2010 Mathematics Subject Classification:
65N12, 65N15, 65N30, 35J15
Department of Mathematical Sciences, Korea Advanced Institute of Science and Technology, Daejeon, 34141, Korea (hjpark235@kaist.ac.kr, kdy@kaist.ac.kr), This work is partially supported by NRF, contract No. 2021R1A2C1003340.

1. Introduction

There are a wide range of physical and engineering problems that are governed by partial differential equations having an interface. For example, a second-order elliptic partial differential equation with a discontinuous coefficient is often used as a model problem in material sciences and porous media involving multiple materials or media. To solve such a problem, one can use some classical numerical schemes with interface-fitted meshes, such as finite element methods (FEMs), discontinuous Galerkin (DG) methods, etc. However, it is difficult and takes a lot of time to generate such fitted meshes when the domain boundary and the interface are geometrically complicated. Even worse, when the interface is moving, one needs to generate a new fitted mesh as time evolves.

To overcome such difficulties, researchers developed and studied some numerical methods using unfitted/structured meshes, such as cut finite element methods (CutFEMs) [19, 12, 20, 3], extended finite element methods (XFEMs) [5, 6, 23, 28, 34], immersed finite element methods (IFEMs) [29, 30, 26, 22, 24], to name just a few. In particular, the IFEMs use basis functions that are modified so that they satisfy the interface conditions. The authors in [29, 30] studied IFEMs using uniform triangular or rectangular grids. In [24, 31], the performance of the IFEMs was improved by adding penalty terms that are commonly used in DG methods. Linear and bilinear nonconforming IFEMs were studied in [26, 32]. The IFEM was also successfully applied to other interface problems: interface elasticity problems [25], elliptic eigenvalue inteface problems [27], Stokes interface problems [1], etc.

On the other hand, several numerical methods using general polygonal or polyhedral meshes have been developed, such as hybrid high-order (HHO) methods [17, 18, 16], virtual element methods (VEMs) [4, 2, 10], weak Galerkin (WG) methods (or weak Galerkin finite element methods) [39, 40, 35], etc. Here we explain the WG methods in some detail. In WG methods, the discrete space consists of polynomials on an element interior and polynomials on its edges, and the differential operators are replaced by the so-called weak differential operators. Compared to the classical FEMs, the WG methods have several advantages. For example, WG methods can handle the general polygonal and polyhedral meshes while the FEMs cannot. In addition, the WG methods can be generalized to higher orders directly. Due to such advantages, the WG methods were successfully applied to various problems: Darcy problems [40], Stokes equations [41], elasticity problems [44], Maxwell equations [36], etc. For more thorough survey, we refer to [39, 35, 43, 33, 42, 21] and references therein.

In [37], an immersed WG method was proposed for the elliptic interface problems for triangular meshes. However, their method cannot be generalized to the polygonal meshes since it is impossible to define the Lagrange-type immersed finite element interpolation on polygonal elements. Besides, the discrete bilinear form formulated in their method is different from the usual WG method; they use the usual gradient and DG-type consistency terms.

In this paper, we develop a new immersed WG method for the elliptic interface problems. Our method uses general polygonal meshes which allow the interface cut through the interior. We generalize the discrete weak gradient to the case when the coefficient is discontinuous, and use it to define the bilinear form. Our weak gradient coincides with the usual one [35] when the coefficient is constant. However, they are different from each other when the coefficient is non-constant.

The rest of the paper is organized as follows. In the next section, we describe the model problem and summarize some preliminaries. In Section 3, we propose our immersed WG method for the model problem, and prove that the discrete problem is well-posed. In Section 4, we prove some technical inequalities and approximation properties of broken linear polynomials on polygonal elements. In Section 5, we derive an optimal error estimate in the discrete H1H^{1}-seminorm under some regularity assumptions on the exact solution. Finally, in Section 6, we present some numerical experiments that confirm our theoretical analysis.

2. Preliminaries

We follow the usual notation of Sobolev spaces, inner product, seminorms, and norms (see, for example, [15]). Let DD be a bounded domain in \mathbb{R} or 2\mathbb{R}^{2}. For σ0\sigma\geq 0, we denote by σ,D\|\cdot\|_{\sigma,D} and ||σ,D|\cdot|_{\sigma,D} the usual norm and seminorm of the Sobolev space Hσ(D)H^{\sigma}(D), respectively. We also denote by (,)0,D(\cdot,\cdot)_{0,D} the usual inner product in L2(D)L^{2}(D). We define H1/2(D)H^{-1/2}(D) as the dual space of H1/2(D)H^{1/2}(D) equipped with the norm given by

u1/2,DsupvH1/2(D)u,vDv1/2,D,\|u\|_{-1/2,D}\coloneqq\sup_{v\in H^{1/2}(D)}\frac{\left\langle u,v\right\rangle_{D}}{\|v\|_{1/2,D}},

where ,D\left\langle\cdot,\cdot\right\rangle_{D} is the duality pairing. For a nonnegative integer kk, we denote by k(D)\mathbb{P}_{k}(D) the space of all polynomials of degree k\leq k on DD.

2.1. Model problem

Let Ω\Omega be a convex polygonal domain in 2\mathbb{R}^{2}, which is separated into two disjoint subdomains Ω+\Omega^{+} and Ω\Omega^{-} by an interface Γ=ΩΩ+\Gamma=\partial\Omega^{-}\cap\partial\Omega^{+} as in Figure 1. Here we assume that Γ\Gamma is a regular C2C^{2}-curve that is not self-intersecting. For any domain DΩD\subset\Omega and any function u:Du:D\to\mathbb{R}, we define its jump across the portion of the interface ΓD\Gamma\cap D as

[u]ΓDu|DΩ+u|DΩ.\left[u\right]_{\Gamma\cap D}\coloneqq u|_{D\cap\Omega^{+}}-u|_{D\cap\Omega^{-}}.

We consider the following elliptic interface problem: Given fL2(Ω)f\in L^{2}(\Omega), find uH01(Ω)u\in H_{0}^{1}(\Omega) such that

{(βu)=fin Ω+Ω,u=0on Ω,\left\{\begin{array}[]{rl}-\nabla\cdot(\beta\nabla u)=f&\textrm{in $\Omega^{+}\cup\Omega^{-}$},\\ u=0&\textrm{on $\partial\Omega$},\end{array}\right. (2.1)

with the jump conditions on the interface

[u]Γ=0,[βu𝒏]Γ=0,\left[u\right]_{\Gamma}=0,\quad\left[\beta\frac{\partial u}{\partial\boldsymbol{n}}\right]_{\Gamma}=0, (2.2)

where β\beta is a positive and piecewise C1C^{1}-function on Ω¯\overline{\Omega} bounded below and above by two positive constants β\beta_{*} and β\beta^{*} with 0<ββ+<0<\beta^{-}\leq\beta^{+}<\infty. That is,

β(𝒙)={β+(𝒙)if 𝒙Ω+,β(𝒙)if 𝒙Ω,\beta(\boldsymbol{x})=\left\{\begin{array}[]{ll}\beta^{+}(\boldsymbol{x})&\textrm{if $\boldsymbol{x}\in\Omega^{+}$},\\ \beta^{-}(\boldsymbol{x})&\textrm{if $\boldsymbol{x}\in\Omega^{-}$},\end{array}\right.

for some functions β+C1(Ω+¯),βC1(Ω¯)\beta^{+}\in C^{1}(\overline{\Omega^{+}}),\beta^{-}\in C^{1}(\overline{\Omega^{-}}) such that ββsβ\beta_{*}\leq\beta^{s}\leq\beta^{*}, s=+,s=+,-. A weak formulation of the model problem (2.1)-(2.2) is written as follows: Find uH01(Ω)u\in H_{0}^{1}(\Omega) such that

Ωβuvd𝒙=Ωfvd𝒙vH01(Ω).\int_{\Omega}\beta\nabla u\cdot\nabla v\mathop{}\!\mathrm{d}\boldsymbol{x}=\int_{\Omega}fv\mathop{}\!\mathrm{d}\boldsymbol{x}\quad\forall v\in H_{0}^{1}(\Omega). (2.3)

For any domain DΩD\subset\Omega, let us introduce the space

H~2(D){uH1(D):u|DΩsH2(DΩs),s=+,}\widetilde{H}^{2}(D)\coloneqq\left\{u\in H^{1}(D):u|_{D\cap\Omega^{s}}\in H^{2}(D\cap\Omega^{s}),\ s=+,-\right\}

equipped with the following norm and seminorm:

uH~2(D)2\displaystyle\|u\|_{\widetilde{H}^{2}(D)}^{2} \displaystyle\coloneqq u1,D2+|u|2,DΩ+2+|u|2,DΩ2,\displaystyle\|u\|_{1,D}^{2}+|u|_{2,D\cap\Omega^{+}}^{2}+|u|_{2,D\cap\Omega^{-}}^{2},
|u|H~2(D)2\displaystyle|u|_{\widetilde{H}^{2}(D)}^{2} \displaystyle\coloneqq |u|2,DΩ+2+|u|2,DΩ2.\displaystyle|u|_{2,D\cap\Omega^{+}}^{2}+|u|_{2,D\cap\Omega^{-}}^{2}.

We also define

H~Γ2(D){uH~2(D):[βu𝒏]ΓD=0}.\widetilde{H}_{\Gamma}^{2}(D)\coloneqq\left\{u\in\widetilde{H}^{2}(D):\left[\beta\frac{\partial u}{\partial\boldsymbol{n}}\right]_{\Gamma\cap D}=0\right\}.

Then we have the following regularity theorem for the solution uu of the variational problem (2.3); see [8, 13].

Theorem 2.1.

Suppose that fL2(Ω)f\in L^{2}(\Omega). Then the variational problem (2.3) has a unique solution uH01(Ω)H~Γ2(Ω)u\in H_{0}^{1}(\Omega)\cap\widetilde{H}_{\Gamma}^{2}(\Omega) satisfying

uH~2(Ω)Cf0,Ω\|u\|_{\widetilde{H}^{2}(\Omega)}\leq C\|f\|_{0,\Omega} (2.4)

for some generic positive constant CC.

Ω\Omega^{-}Ω+\Omega^{+}Γ\Gamma
Figure 1. A domain Ω\Omega with interface Γ\Gamma.

2.2. Mesh assumptions

Let {𝒯h}h\{\mathcal{T}_{h}\}_{h} be a family of decompositions (meshes) of Ω\Omega into polygonal elements TT with maximum diameter hh. Let h\mathcal{E}_{h} be the set of all edges in 𝒯h\mathcal{T}_{h}. Let hi\mathcal{E}_{h}^{i} and hb\mathcal{E}_{h}^{b} denote the set of all interior and boundary edges in 𝒯h\mathcal{T}_{h}, respectively. For each T𝒯hT\in\mathcal{T}_{h}, let T\mathcal{E}_{T} be the set of all edges of TT. For each T𝒯hT\in\mathcal{T}_{h}, we denote by |T||T| the area of TT, by hTh_{T} the diameter of TT, and by 𝒏T\boldsymbol{n}_{T} its exterior unit normal vector along the boundary T\partial T. For each ehe\in\mathcal{E}_{h}, we denote by |e||e| the length of ee. For ehie\in\mathcal{E}_{h}^{i}, we define 𝒏e\boldsymbol{n}_{e} by a unit normal vector of ee with orientation fixed once and for all. For ehbe\in\mathcal{E}_{h}^{b}, we define 𝒏e\boldsymbol{n}_{e} by a unit normal vector on ee in the outward direction with respect to Ω\Omega.

We call an element T𝒯hT\in\mathcal{T}_{h} an interface element if the interface Γ\Gamma passes through the interior of TT; otherwise we call TT a noninterface element. We denote by 𝒯hI\mathcal{T}_{h}^{I} the collection of all interface elements in 𝒯h\mathcal{T}_{h}, and by 𝒯hN\mathcal{T}_{h}^{N} the collection of all non-interface elements in 𝒯h\mathcal{T}_{h}. For an interface element T𝒯hT\in\mathcal{T}_{h}, we denote by ΓhT\Gamma_{h}^{T} the line segment connecting the intersections of Γ\Gamma and the edges of TT. This line segment divides TT into two parts T+T^{+} and TT^{-} with T¯=T+T¯\overline{T}=\overline{T^{+}\cup T^{-}} (see, for example, Figure 2). For any function u:Tu:T\to\mathbb{R}, we define its jump across ΓhTT\Gamma_{h}^{T}\cap T as

[u]ΓhTu|T+u|T.\left[u\right]_{\Gamma_{h}^{T}}\coloneqq u|_{T^{+}}-u|_{T^{-}}.

We assume that {𝒯h}h\{\mathcal{T}_{h}\}_{h} satisfies the following regularity assumptions [4, 40, 26].

Assumption 2.2.

There exists ρ>0\rho>0 independent of hh such that

  1. (i)

    the decomposition 𝒯h\mathcal{T}_{h} consists of a finite number of nonoverlapping polygonal elements;

  2. (ii)

    for any T𝒯hT\in\mathcal{T}_{h} the diameter of any edge of TT is larger than ρhT\rho h_{T};

  3. (iii)

    every element TT of 𝒯h\mathcal{T}_{h} is star-shaped with respect to a ball BTB_{T} with center 𝒙T\boldsymbol{x}_{T} and radius ρhT\rho h_{T};

  4. (iv)

    if ee is an edge of T𝒯hT\in\mathcal{T}_{h} then |e|ρhT|e|\geq\rho h_{T};

  5. (v)

    the interface Γ\Gamma meets the edges of an interface element at no more than two points;

  6. (vi)

    the interface Γ\Gamma meets each edge in h\mathcal{E}_{h} at most once, except possibly it passes through two vertices.

Remark 2.3.

The assumptions (v) and (vi) are resonable if hh is sufficiently small. Note also that the assumptions (i)-(iv) imply that the following properties [10]:

  • Every T𝒯hT\in\mathcal{T}_{h} has at most NN edges and vertices, where NN is independent of hh.

  • Each element T𝒯hT\in\mathcal{T}_{h} can be decomposed as NN triangles, obtained by connecting the vertices of TT to 𝒙T\boldsymbol{x}_{T}, such that the minimum angle of the triangles is controlled by ρ\rho.

Throughout this paper, CC will denote a generic positive constant independent of hh, not necessarily the same in each occurrence.

Γ\GammaTΩT\cap\Omega^{-}TΩ+T\cap\Omega^{+}
Γ\GammaΓhT\Gamma_{h}^{T}TT^{-}T+T^{+}
Figure 2. An interface element TT in 𝒯h\mathcal{T}_{h}.

3. Immersed Weak Galerkin Method

In this section, we describe an immersed WG method for the problem (2.3).

3.1. Broken polynomial space

Let T𝒯hT\in\mathcal{T}_{h} be an interface element. We define the piecewise constant function β¯T\overline{\beta}_{T} on the element TT as follows:

β¯T(𝒙)={β¯+if 𝒙T+,β¯if 𝒙T,\overline{\beta}_{T}(\boldsymbol{x})=\left\{\begin{array}[]{ll}\overline{\beta}^{+}&\textrm{if $\boldsymbol{x}\in T^{+}$},\\ \overline{\beta}^{-}&\textrm{if $\boldsymbol{x}\in T^{-}$},\end{array}\right.

where β¯sβs(𝒙s)\overline{\beta}^{s}\coloneqq\beta^{s}(\boldsymbol{x}^{s}) and 𝒙s\boldsymbol{x}^{s} denotes the barycenter of TsT^{s} for s=+,s=+,-. We also let β¯\overline{\beta} be the piecewise constant function such that β¯|T=β¯T\overline{\beta}|_{T}=\overline{\beta}_{T} on each T𝒯hT\in\mathcal{T}_{h}. The broken polynomial space ^1(T)\widehat{\mathbb{P}}_{1}(T) of degree 1\leq 1 is defined by

^1(T){q:q|T+1(T+),q|T1(T),[q]ΓhT=0,[β¯Tq𝒏]ΓhT=0}.\widehat{\mathbb{P}}_{1}(T)\coloneqq\left\{q:q|_{T^{+}}\in\mathbb{P}_{1}(T^{+}),\ q|_{T^{-}}\in\mathbb{P}_{1}(T^{-}),\ \left[q\right]_{\Gamma_{h}^{T}}=0,\ \left[\overline{\beta}_{T}\frac{\partial q}{\partial\boldsymbol{n}}\right]_{\Gamma_{h}^{T}}=0\right\}.

It is easy to see that dim^1(T)=3\dim\widehat{\mathbb{P}}_{1}(T)=3 (see, for example, [26, Theorem 2.2]), and the following piecewise polynomials form a basis of ^1(T)\widehat{\mathbb{P}}_{1}(T):

φ1(𝒙)=1,φ2(𝒙)=𝒕(𝒙𝒙0),φ3(𝒙)=β¯T1𝒏(𝒙𝒙0),\varphi_{1}(\boldsymbol{x})=1,\quad\varphi_{2}(\boldsymbol{x})=\boldsymbol{t}\cdot(\boldsymbol{x}-\boldsymbol{x}_{0}),\quad\varphi_{3}(\boldsymbol{x})=\overline{\beta}_{T}^{-1}\boldsymbol{n}\cdot(\boldsymbol{x}-\boldsymbol{x}_{0}),

where 𝒙0\boldsymbol{x}_{0} is the midpoint of the line segment ΓhT\Gamma_{h}^{T}, 𝒏=(n1,n2)\boldsymbol{n}=(n_{1},n_{2}) is a unit vector normal to ΓhT\Gamma_{h}^{T} pointing from T+T^{+} to TT^{-}, and 𝒕=(n2,n1)\boldsymbol{t}=(-n_{2},n_{1}). Note that, since ^1(T)H1(T)\widehat{\mathbb{P}}_{1}(T)\subset H^{1}(T), the space ^1(T)\nabla\widehat{\mathbb{P}}_{1}(T) is well-defined, and the vector-valued functions φ2\nabla\varphi_{2} and φ3\nabla\varphi_{3} form a basis of ^1(T)\nabla\widehat{\mathbb{P}}_{1}(T).

For convenience, we set ^1(T)1(T)\widehat{\mathbb{P}}_{1}(T)\coloneqq\mathbb{P}_{1}(T) for any non-interface element T𝒯hT\in\mathcal{T}_{h}. Let

^1(Ω):={qL2(Ω):q|T^1(T)T𝒯h}.\widehat{\mathbb{P}}_{1}(\Omega):=\big{\{}q\in L^{2}(\Omega):q|_{T}\in\widehat{\mathbb{P}}_{1}(T)\ \forall T\in\mathcal{T}_{h}\big{\}}.

3.2. Weak Galerkin finite element space

We define the weak Galerkin finite element space VhV_{h} associated to 𝒯h\mathcal{T}_{h} and its subspace Vh,0V_{h,0} as follows:

Vh\displaystyle V_{h} \displaystyle\coloneqq {v={v0,v}:v0|T^1(T)T𝒯h,v|e0(e)eh},\displaystyle\left\{v=\{v_{0},v_{\partial}\}:v_{0}|_{T}\in\widehat{\mathbb{P}}_{1}(T)\ \forall T\in\mathcal{T}_{h},\ v_{\partial}|_{e}\in\mathbb{P}_{0}(e)\ \forall e\in\mathcal{E}_{h}\right\},
Vh,0\displaystyle V_{h,0} \displaystyle\coloneqq {vVh:v=0onΩ}.\displaystyle\left\{v\in V_{h}:v_{\partial}=0\ \textrm{on}\ \partial\Omega\right\}.

Here we note that, for any v={v0,v}Vhv=\{v_{0},v_{\partial}\}\in V_{h}, its second component vv_{\partial} is a single-valued function on each edge ehe\in\mathcal{E}_{h}. Thus, the space VhV_{h} has 33 degrees of freedom on the interior of each element T𝒯hT\in\mathcal{T}_{h} and 11 degree of freedom on each edge ehe\in\mathcal{E}_{h}.

For each element T𝒯hT\in\mathcal{T}_{h}, let Q0Q_{0} be the L2L^{2}-projection from L2(T)L^{2}(T) onto ^1(T)\widehat{\mathbb{P}}_{1}(T). Similarly, for each edge ehe\in\mathcal{E}_{h}, let QQ_{\partial} the L2L^{2}-projection from L2(e)L^{2}(e) onto 0(e)\mathbb{P}_{0}(e). We then define a projection operator Qh:H1(Ω)VhQ_{h}:H^{1}(\Omega)\to V_{h} by

Qhv={Q0v,Qv},vH1(Ω).Q_{h}v=\{Q_{0}v,Q_{\partial}v\},\quad v\in H^{1}(\Omega). (3.1)

3.3. Discrete problem and well-posedness

For each vh={v0,v}Vhv_{h}=\{v_{0},v_{\partial}\}\in V_{h}, we define a discrete weak gradient wvh\nabla_{w}v_{h} of vhv_{h} as a vector-valued function satisfying wv|T^1(T)\nabla_{w}v|_{T}\in\nabla\widehat{\mathbb{P}}_{1}(T) and

Tβ¯Twvhqd𝒙=Tβ¯Tv0qd𝒙T(Qv0v)(β¯Tq𝒏T)dsq^1(T),\int_{T}\overline{\beta}_{T}\nabla_{w}v_{h}\cdot\nabla q\mathop{}\!\mathrm{d}\boldsymbol{x}=\int_{T}\overline{\beta}_{T}\nabla v_{0}\cdot\nabla q\mathop{}\!\mathrm{d}\boldsymbol{x}-\int_{\partial T}(Q_{\partial}v_{0}-v_{\partial})\left(\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{T}\right)\mathop{}\!\mathrm{d}s\quad\forall q\in\widehat{\mathbb{P}}_{1}(T), (3.2)

for each element T𝒯hT\in\mathcal{T}_{h}.

We next introduce two bilinear forms on Vh×VhV_{h}\times V_{h} as follows:

a(uh,vh)\displaystyle a(u_{h},v_{h}) \displaystyle\coloneqq T𝒯hTβ¯Twuhwvhd𝒙,\displaystyle\sum_{T\in\mathcal{T}_{h}}\int_{T}\overline{\beta}_{T}\nabla_{w}u_{h}\cdot\nabla_{w}v_{h}\mathop{}\!\mathrm{d}\boldsymbol{x},
s(uh,vh)\displaystyle s(u_{h},v_{h}) \displaystyle\coloneqq λT𝒯hhT1T(Qu0u)(Qv0v)ds,\displaystyle\lambda\sum_{T\in\mathcal{T}_{h}}h_{T}^{-1}\int_{\partial T}(Q_{\partial}u_{0}-u_{\partial})(Q_{\partial}v_{0}-v_{\partial})\mathop{}\!\mathrm{d}s,

for any uh={u0,u}Vhu_{h}=\{u_{0},u_{\partial}\}\in V_{h} and vh={v0,v}Vhv_{h}=\{v_{0},v_{\partial}\}\in V_{h}, where λ\lambda is an arbitrary positive constant. The stabilization as(,)a_{s}(\cdot,\cdot) of a(,)a(\cdot,\cdot) is defined by

as(uh,vh)=a(uh,vh)+s(uh,vh)uh,vhVh.a_{s}(u_{h},v_{h})=a(u_{h},v_{h})+s(u_{h},v_{h})\quad\forall u_{h},v_{h}\in V_{h}.

We are now ready to formulate the immersed WG method for solving (2.3) as follows: Find uhVh,0u_{h}\in V_{h,0} such that

as(uh,vh)=(f,v0)0,Ω,vh={v0,v}Vh,0.a_{s}(u_{h},v_{h})=(f,v_{0})_{0,\Omega},\quad\forall v_{h}=\{v_{0},v_{\partial}\}\in V_{h,0}. (3.3)

We next analyze the well-posedness of the discrete problem (3.3). Define the energy-norm ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} by

|vh|as(vh,vh)vhVh.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\coloneqq\sqrt{a_{s}(v_{h},v_{h})}\quad\forall v_{h}\in V_{h}.

Clearly ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} is a seminorm on VhV_{h}. Moreover, ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} is a norm on Vh,0V_{h,0}, as shown in the following lemma.

Lemma 3.1.

||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} is a norm on Vh,0V_{h,0}.

Proof.

It suffices to show that |vh|=0{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=0 \Rightarrow vh0v_{h}\equiv 0 for any vhVh,0v_{h}\in V_{h,0}. Suppose that vh={v0,v}Vh,0v_{h}=\{v_{0},v_{\partial}\}\in V_{h,0} satisfies |vh|=0{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=0. Since

0=|vh|2=T𝒯hTβ¯T|wvh|2d𝒙+λT𝒯heThT1e|Qv0v|2ds0={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}^{2}=\sum_{T\in\mathcal{T}_{h}}\int_{T}\overline{\beta}_{T}|\nabla_{w}v_{h}|^{2}\mathop{}\!\mathrm{d}\boldsymbol{x}+\lambda\sum_{T\in\mathcal{T}_{h}}\sum_{e\subset\partial T}h_{T}^{-1}\int_{e}|Q_{\partial}v_{0}-v_{\partial}|^{2}\mathop{}\!\mathrm{d}s

and since 0<β<β¯T0<\beta_{*}<\overline{\beta}_{T} for any T𝒯hT\in\mathcal{T}_{h}, we obtain wvh0\nabla_{w}v_{h}\equiv 0 and Qv0=vQ_{\partial}v_{0}=v_{\partial} on each edge ehe\in\mathcal{E}_{h}. Then

0\displaystyle 0 =\displaystyle= Tβ¯Twvhv0d𝒙=Tβ¯Tv0v0d𝒙+eTe(vQv0)(β¯Tv0𝒏)ds\displaystyle\int_{T}\overline{\beta}_{T}\nabla_{w}v_{h}\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}=\int_{T}\overline{\beta}_{T}\nabla v_{0}\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}+\sum_{e\subset\partial T}\int_{e}(v_{\partial}-Q_{\partial}v_{0})\left(\overline{\beta}_{T}\frac{\partial v_{0}}{\partial\boldsymbol{n}}\right)\mathop{}\!\mathrm{d}s
=\displaystyle= Tβ¯T|v0|2d𝒙Tβ|v0|2d𝒙\displaystyle\int_{T}\overline{\beta}_{T}|\nabla v_{0}|^{2}\mathop{}\!\mathrm{d}\boldsymbol{x}\geq\int_{T}\beta_{*}|\nabla v_{0}|^{2}\mathop{}\!\mathrm{d}\boldsymbol{x}

for any T𝒯hT\in\mathcal{T}_{h}. This shows that v0=0\nabla v_{0}=0 on each T𝒯hT\in\mathcal{T}_{h}. Note that, for each T𝒯hT\in\mathcal{T}_{h}, q=0\nabla q=0 implies q=constantq=\textrm{constant} for any q^1(T)q\in\widehat{\mathbb{P}}_{1}(T). Since v0^1(T)v_{0}\in\widehat{\mathbb{P}}_{1}(T) on each T𝒯hT\in\mathcal{T}_{h}, we obtain that v0v_{0} is constant on each T𝒯hT\in\mathcal{T}_{h}. Since Qv0=vQ_{\partial}v_{0}=v_{\partial} on each ehe\in\mathcal{E}_{h} and v=0v_{\partial}=0 on Ω\partial\Omega, we conclude that v0=v=0v_{0}=v_{\partial}=0. ∎

The well-posedness of the discrete problem (3.3) directly follows from the lemma.

Corollary 3.2.

The discrete problem (3.3) is well-posed.

Proof.

From Lemma 3.1, the bilinear form as(,)a_{s}(\cdot,\cdot) on Vh,0V_{h,0} is coercive and continuous with respect to the norm ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} on Vh,0V_{h,0}. The conclusion follows from the Lax-Milgram Lemma. ∎

4. Some Estimates on Interface Elements

In this section, we present some inequalities for the function spaces on the interface elements, which are needed for the error analysis of the immersed WG method.

4.1. Geometric assumptions on interface elements

Let T𝒯hT\in\mathcal{T}_{h} be an interface element. Recall that ΓhT\Gamma_{h}^{T} denotes the line segment connecting two intersection points of Γ\Gamma and the edges of TT. Although the analysis works for C2C^{2}-interface, we assume for the simplicity of presentation, that on each mesh element TT, the portion ΓT\Gamma\cap T is a line segment so that ΓT=ΓhT\Gamma\cap T=\Gamma_{h}^{T} and Ts=TΩsT^{s}=T\cap\Omega^{s} for s=+,s=+,-. In addition, we assume that ΓT\Gamma\cap T aligns with the xx-axis and the origin of the xyxy-plane is contained in TT, so that

T+=T{(x1,x2)2:x20},T=T{(x1,x2)2:x20}T^{+}=T\cap\{(x_{1},x_{2})\in\mathbb{R}^{2}:x_{2}\geq 0\},\quad T^{-}=T\cap\{(x_{1},x_{2})\in\mathbb{R}^{2}:x_{2}\leq 0\} (4.1)

(see Figure 3). Since hT=diam(T)h_{T}=\operatorname{diam}(T), we have T[hT,hT]2T\subset[-h_{T},h_{T}]^{2}. Since βsC1(Ωs¯)\beta^{s}\in C^{1}(\overline{\Omega^{s}}) and β¯T=β¯s\overline{\beta}_{T}=\overline{\beta}^{s} on TsT^{s} for s=+,s=+,-, we have

max𝒙Ts|β(𝒙)β¯T(𝒙)|ChT,max𝒙eΩs|β(𝒙)β¯T(𝒙)|ChT,s=+,,\max_{\boldsymbol{x}\in T^{s}}|\beta(\boldsymbol{x})-\overline{\beta}_{T}(\boldsymbol{x})|\leq Ch_{T},\quad\max_{\boldsymbol{x}\in e\cap\Omega^{s}}|\beta(\boldsymbol{x})-\overline{\beta}_{T}(\boldsymbol{x})|\leq Ch_{T},\quad s=+,-, (4.2)

where eTe\subset\partial T. Let 𝒏Γ=(n1,h,n2,h)\boldsymbol{n}_{\Gamma}=(n_{1,h},n_{2,h}) be the unit vector normal to Γ\Gamma pointing from T+T^{+} to TT^{-}, and let 𝒕Γ=(n2,h,n1,h)\boldsymbol{t}_{\Gamma}=(-n_{2,h},n_{1,h}).

Remark 4.1.

We briefly discuss the case when the interface is not piecewise linear, that is, ΓTΓhT\Gamma\cap T\neq\Gamma_{h}^{T}. Without loss of generality we assume that ΓhT\Gamma_{h}^{T} aligns with the xx-axis and TT is contained in the box Ix×IyI_{x}\times I_{y}, where IxI_{x} and IyI_{y} are intervals with length not greater than 2hT2h_{T}. Since Γ\Gamma is a regular C2C^{2}-curve, there exists a parametrization t(t,γ(t))t\mapsto(t,\gamma(t)) of the curve ΓT\Gamma\cap T for some γC2(Ix)\gamma\in C^{2}(I_{x}), when hh is sufficiently small. Then the unit normal vector 𝒏Γ\boldsymbol{n}_{\Gamma} along ΓT\Gamma\cap T pointing from Ω+\Omega^{+} to Ω\Omega^{-} is

𝒏Γ(t,γ(t))=(γ(t)(1+|γ(t)|2)1/2,1(1+|γ(t)|2)1/2),tIx.\boldsymbol{n}_{\Gamma}(t,\gamma(t))=\left(\frac{\gamma^{\prime}(t)}{(1+|\gamma^{\prime}(t)|^{2})^{1/2}},\frac{-1}{(1+|\gamma^{\prime}(t)|^{2})^{1/2}}\right),\quad t\in I_{x}.

Let us extend the vector-valued function 𝒏Γ\boldsymbol{n}_{\Gamma} to the box Ix×IyI_{x}\times I_{y} by setting (t,y)𝒏Γ(t,γ(t))(t,y)\mapsto\boldsymbol{n}_{\Gamma}(t,\gamma(t)). Then, since γ\gamma is C2C^{2}, we have

sup𝒙T|𝒏Γ(𝒙)𝒏Γh|ChT,\sup_{\boldsymbol{x}\in T}\big{|}\boldsymbol{n}_{\Gamma}(\boldsymbol{x})-\boldsymbol{n}_{\Gamma}^{h}\big{|}\leq Ch_{T}, (4.3)

where 𝒏Γh\boldsymbol{n}_{\Gamma}^{h} is the unit normal vector along ΓhT\Gamma_{h}^{T} pointing from T+T^{+} to TT^{-}. In addition, one can obtain a similar result for the tangential vector of ΓT\Gamma\cap T. Next, according to Lemma 2 in [7],

u0,Tr2ChT2s=+,((u)|Ωs0,ΓT2+hT2|u|1,TΩs2),uH~2(T),\|\nabla u\|_{0,T_{r}}^{2}\leq Ch_{T}^{2}\sum_{s=+,-}\left(\|(\nabla u)|_{\Omega_{s}}\|_{0,\Gamma\cap T}^{2}+h_{T}^{2}|\nabla u|_{1,T\cap\Omega^{s}}^{2}\right),\quad\forall u\in\widetilde{H}^{2}(T), (4.4)

where TrT_{r} is a subset of TT given by

Tr=T(Ω+T+)(ΩT);T_{r}=T-(\Omega^{+}\cap T^{+})-(\Omega^{-}\cap T^{-});

see Figure 2. Note also that the first estimate in (4.2) is modified as follows:

sup𝒙Ts(TΩs)|β(𝒙)β¯T(𝒙)|ChT,s=+,.\sup_{\boldsymbol{x}\in T^{s}\cap(T\cap\Omega^{s})}|\beta(\boldsymbol{x})-\overline{\beta}_{T}(\boldsymbol{x})|\leq Ch_{T},\quad s=+,-. (4.5)

Using the estimates (4.3)-(4.5) and the standard trace inequality, all the results below can be derived with only minor modification. We leave the detailed analysis for a future investigation.

Lemma 4.2.

If hh is sufficiently small, then either T+T^{+} or TT^{-} contains a ball with radius ρhT/8\rho h_{T}/8.

Proof.

Recall that TT is star-shaped with respect to a ball BB centered at 𝒙T=(xT,yT)\boldsymbol{x}_{T}=(x_{T},y_{T}) with radius ρhT\rho h_{T}. First, assume that |yT|ρhT/8|y_{T}|\leq\rho h_{T}/8. Consider the ball B+B^{+} centered at (xT,yT+ρhT/2)(x_{T},y_{T}+\rho h_{T}/2) with radius ρhT/8\rho h_{T}/8. Then B+BT+B^{+}\subset B\cap T^{+}.

One can show that, by the same argument, for the case yTρhT/8y_{T}\geq\rho h_{T}/8 the set T+T^{+} contains the ball centered at (xT,yT+ρhT/2)(x_{T},y_{T}+\rho h_{T}/2) with radius ρhT/8\rho h_{T}/8, and for the case yTρhT/8y_{T}\leq-\rho h_{T}/8 the set TT^{-} contains the ball centered at (xT,yTρhT/2)(x_{T},y_{T}-\rho h_{T}/2) with radius ρhT/8\rho h_{T}/8. ∎

ΓhT\Gamma_{h}^{T} (=ΓT=\Gamma\cap T)T+T^{+}TT^{-}𝟎\boldsymbol{0}xxyy𝒏Γ\boldsymbol{n}_{\Gamma}
Figure 3. Geometric assumptions on an interface element TT.

4.2. Some inequalities for the broken polynomial space ^1\widehat{\mathbb{P}}_{1}

Recall that, on each element T𝒯hT\in\mathcal{T}_{h}, the standard trace inequality holds:

hT1/2v0,TC(v0,T+hTv0,T)vH1(T).h_{T}^{1/2}\|v\|_{0,\partial T}\leq C\left(\|v\|_{0,T}+h_{T}\|\nabla v\|_{0,T}\right)\quad\forall v\in H^{1}(T). (4.6)

The following lemma provides a trace inequality for the space ^1\nabla\widehat{\mathbb{P}}_{1}.

Lemma 4.3.

Let T𝒯hT\in\mathcal{T}_{h} be an interface element. Then there exists a positive constant CC depending only on ρ\rho and β\beta such that for any q^1(T)q\in\widehat{\mathbb{P}}_{1}(T) and any edge ee of TT,

β¯Tq0,eChT1/2β¯T1/2q0,T,\left\|\overline{\beta}_{T}\nabla q\right\|_{0,e}\leq Ch_{T}^{-1/2}\big{\|}\overline{\beta}_{T}^{1/2}\nabla q\big{\|}_{0,T}, (4.7)
Proof.

Recall that the following piecewise polynomials form a basis of the space ^1(T)\widehat{\mathbb{P}}_{1}(T):

φ1(𝒙)=1,φ2(𝒙)=𝒕Γ(𝒙𝒙0),φ3(𝒙)=β¯T1𝒏Γ(𝒙𝒙0),𝒙T,\varphi_{1}(\boldsymbol{x})=1,\quad\varphi_{2}(\boldsymbol{x})=\boldsymbol{t}_{\Gamma}\cdot(\boldsymbol{x}-\boldsymbol{x}_{0}),\quad\varphi_{3}(\boldsymbol{x})=\overline{\beta}_{T}^{-1}\boldsymbol{n}_{\Gamma}\cdot(\boldsymbol{x}-\boldsymbol{x}_{0}),\quad\forall\boldsymbol{x}\in T,

where 𝒙0\boldsymbol{x}_{0} is the midpoint of ΓhT\Gamma_{h}^{T}. Let q=aφ1+bφ2+cφ3q=a\varphi_{1}+b\varphi_{2}+c\varphi_{3} for a,b,ca,b,c\in\mathbb{R}. Then

q=b𝒕Γ+cβ¯T1𝒏Γ,qq=b2+c2β¯T2.\nabla q=b\boldsymbol{t}_{\Gamma}+c\overline{\beta}_{T}^{-1}\boldsymbol{n}_{\Gamma},\quad\nabla q\cdot\nabla q=b^{2}+c^{2}\overline{\beta}_{T}^{-2}.

By 2.2 (iii), we have

β¯Tq0,e2=\displaystyle\left\|\overline{\beta}_{T}\nabla q\right\|_{0,e}^{2}= e|β¯Tq|2ds((β)2b2+c2)|e|C(β,β,ρ)(b2+c2)hT,\displaystyle\ \int_{e}\left|\overline{\beta}_{T}\nabla q\right|^{2}\mathop{}\!\mathrm{d}s\leq((\beta^{*})^{2}b^{2}+c^{2})|e|\leq C(\beta_{*},\beta^{*},\rho)(b^{2}+c^{2})h_{T},
β¯T1/2q0,T2=\displaystyle\big{\|}\overline{\beta}_{T}^{1/2}\nabla q\big{\|}_{0,T}^{2}= Tβ¯T|q|2d𝒙β(b2+c2(β)2)|T|C(β,β,ρ)(b2+c2)hT2.\displaystyle\ \int_{T}\overline{\beta}_{T}|\nabla q|^{2}\mathop{}\!\mathrm{d}\boldsymbol{x}\geq\beta_{*}(b^{2}+c^{2}(\beta^{*})^{-2})|T|\geq C(\beta_{*},\beta^{*},\rho)(b^{2}+c^{2})h_{T}^{2}.

Thus there exists a positive constant CC depending only on ρ\rho and β\beta such that the inequality (4.7) holds. ∎

Note that we have the following inverse inequality holds (see, for example, (2.6) of [10]):

|q|1,TChT1q0,Tq1(T),|q|1,BChT1q0,Bq1(B),|q|_{1,T}\leq Ch_{T}^{-1}\|q\|_{0,T}\quad\forall q\in\mathbb{P}_{1}(T),\qquad|q|_{1,B}\leq Ch_{T}^{-1}\|q\|_{0,B}\quad\forall q\in\mathbb{P}_{1}(B), (4.8)

where BB is a ball in 2\mathbb{R}^{2} with radius ρhT\rho h_{T} and CC is a positive constant depending only on ρ\rho. The following lemma shows that the inverse inequality also holds for the space ^1\widehat{\mathbb{P}}_{1}.

Lemma 4.4.

Let T𝒯hT\in\mathcal{T}_{h} be an interface element. There exists a positive constant CC depending only on ρ\rho and β\beta such that

|q|1,TChT1q0,Tq^1(T).|q|_{1,T}\leq Ch_{T}^{-1}\|q\|_{0,T}\quad\forall q\in\widehat{\mathbb{P}}_{1}(T).
Proof.

By Lemma 4.2, we may assume that T+T^{+} contains a ball B+B^{+} with radius ρhT/8\rho h_{T}/8. As in the proof of the previous lemma, consider the basis {φ1,φ2,φ3}\{\varphi_{1},\varphi_{2},\varphi_{3}\} of ^1(T)\widehat{\mathbb{P}}_{1}(T) and let q=aφ1+bφ2+cφ3q=a\varphi_{1}+b\varphi_{2}+c\varphi_{3} for a,b,ca,b,c\in\mathbb{R}, and define

q+a+b𝒕Γ(𝒙𝒙0)+c(β¯+)1𝒏Γ(𝒙𝒙0).q_{+}\coloneqq a+b\boldsymbol{t}_{\Gamma}\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})+c\big{(}\overline{\beta}^{+}\big{)}^{-1}\boldsymbol{n}_{\Gamma}\cdot(\boldsymbol{x}-\boldsymbol{x}_{0}).

Then q=q+q=q_{+} on T+T^{+}. By (4.8),

|q+|1,B+ChT1q+0,B+=ChT1q0,B+ChT1q0,T.|q_{+}|_{1,B^{+}}\leq Ch_{T}^{-1}\|q_{+}\|_{0,B^{+}}=Ch_{T}^{-1}\|q\|_{0,B^{+}}\leq Ch_{T}^{-1}\|q\|_{0,T}. (4.9)

Since 𝒕Γ𝒏Γ=0\boldsymbol{t}_{\Gamma}\cdot\boldsymbol{n}_{\Gamma}=0,

|q+|1,B+2\displaystyle|q_{+}|_{1,B^{+}}^{2} =\displaystyle= B+|b𝒕Γ+c(β¯+)1𝒏Γ|2d𝒙=B+(b2+(β¯+)2c2)d𝒙\displaystyle\int_{B^{+}}\big{|}b\boldsymbol{t}_{\Gamma}+c\big{(}\overline{\beta}^{+}\big{)}^{-1}\boldsymbol{n}_{\Gamma}\big{|}^{2}\mathop{}\!\mathrm{d}\boldsymbol{x}=\int_{B^{+}}\big{(}b^{2}+\big{(}\overline{\beta}^{+}\big{)}^{-2}c^{2}\big{)}\mathop{}\!\mathrm{d}\boldsymbol{x} (4.10)
\displaystyle\geq πρ2hT264C(β,β)(b2+c2),\displaystyle\frac{\pi\rho^{2}h_{T}^{2}}{64}C(\beta_{*},\beta^{*})(b^{2}+c^{2}),
|q|1,T2\displaystyle|q|_{1,T}^{2} =\displaystyle= T+|b𝒕Γ+c(β¯+)1𝒏Γ|2d𝒙+T|b𝒕Γ+c(β¯)1𝒏Γ|2d𝒙\displaystyle\int_{T^{+}}\big{|}b\boldsymbol{t}_{\Gamma}+c\big{(}\overline{\beta}^{+}\big{)}^{-1}\boldsymbol{n}_{\Gamma}\big{|}^{2}\mathop{}\!\mathrm{d}\boldsymbol{x}+\int_{T^{-}}\big{|}b\boldsymbol{t}_{\Gamma}+c\big{(}\overline{\beta}^{-}\big{)}^{-1}\boldsymbol{n}_{\Gamma}\big{|}^{2}\mathop{}\!\mathrm{d}\boldsymbol{x} (4.11)
=\displaystyle= T+(b2+(β¯+)2c2)d𝒙+T(b2+(β¯)2c2)d𝒙\displaystyle\int_{T^{+}}\big{(}b^{2}+\big{(}\overline{\beta}^{+}\big{)}^{-2}c^{2}\big{)}\mathop{}\!\mathrm{d}\boldsymbol{x}+\int_{T^{-}}\big{(}b^{2}+\big{(}\overline{\beta}^{-}\big{)}^{-2}c^{2}\big{)}\mathop{}\!\mathrm{d}\boldsymbol{x}
\displaystyle\leq C(β,β)hT2(b2+c2).\displaystyle C(\beta_{*},\beta^{*})h_{T}^{2}(b^{2}+c^{2}).

Combining the inequalities (4.9)-(4.11), we obtain

|q|1,T8πρC(β,β)|q+|1,B+C(β,β,ρ)hT1q0,T.|q|_{1,T}\leq\frac{8}{\sqrt{\pi}\rho}C(\beta_{*},\beta^{*})|q_{+}|_{1,B^{+}}\leq C(\beta_{*},\beta^{*},\rho)h_{T}^{-1}\|q\|_{0,T}.

This completes the proof of the lemma. ∎

4.3. Approximation properties of the broken polynomial space ^1\widehat{\mathbb{P}}_{1}

In this subsection, we derive some approximation properties of the broken linear polynomial space ^1(T)\widehat{\mathbb{P}}_{1}(T).

It is well-known that, on each non-interface element T𝒯hT\in\mathcal{T}_{h}, for any uH2(T)u\in H^{2}(T) there exists q1q\in\mathbb{P}_{1} such that

uq0,T+hT|uq|1,TCρhT2u2,T,\|u-q\|_{0,T}+h_{T}|u-q|_{1,T}\leq C_{\rho}h_{T}^{2}\|u\|_{2,T}, (4.12)

where CρC_{\rho} is a positive constant depending only on ρ\rho [11, Lemma 4.3.8].

Theorem 4.5.

Let uH~Γ2(Ω)u\in\widetilde{H}_{\Gamma}^{2}(\Omega). Then there exists q^1(Ω)q\in\widehat{\mathbb{P}}_{1}(\Omega) such that

uq0,Ω+h|uq|1,ΩCh2uH~2(Ω),\|u-q\|_{0,\Omega}+h|u-q|_{1,\Omega}\leq Ch^{2}\|u\|_{\widetilde{H}^{2}(\Omega)},

where CC is a positive constant depending only on ρ\rho and β\beta.

Proof.

Let T𝒯hT\in\mathcal{T}_{h} be an interface element. Then we have

u=(u𝒕Γ)𝒕Γ+(u𝒏Γ)𝒏Γ\nabla u=(\nabla u\cdot\boldsymbol{t}_{\Gamma})\boldsymbol{t}_{\Gamma}+(\nabla u\cdot\boldsymbol{n}_{\Gamma})\boldsymbol{n}_{\Gamma} (4.13)

on TT. We note that u𝒕ΓH1(T)\nabla u\cdot\boldsymbol{t}_{\Gamma}\in H^{1}(T) and βu𝒏ΓH1(T)\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}\in H^{1}(T). Thus, from (4.12), there exist ct,cnc_{t},c_{n}\in\mathbb{R} such that

u𝒕Γct0,TCρhT|u𝒕Γ|1,T,βu𝒏Γcn0,TCρhT|βu𝒏Γ|1,T.\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-c_{t}\|_{0,T}\leq C_{\rho}h_{T}|\nabla u\cdot\boldsymbol{t}_{\Gamma}|_{1,T},\quad\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-c_{n}\|_{0,T}\leq C_{\rho}h_{T}|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}|_{1,T}.

Note that

|u𝒕Γ|1,TCuH~2(T),|u𝒏Γ|1,TCuH~2(T).|\nabla u\cdot\boldsymbol{t}_{\Gamma}|_{1,T}\leq C\|u\|_{\widetilde{H}^{2}(T)},\quad|\nabla u\cdot\boldsymbol{n}_{\Gamma}|_{1,T}\leq C\|u\|_{\widetilde{H}^{2}(T)}. (4.14)

Thus

u𝒕Γct0,TChTuH~2(T),βu𝒏Γcn0,TChTuH~2(T).\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-c_{t}\|_{0,T}\leq Ch_{T}\|u\|_{\widetilde{H}^{2}(T)},\quad\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-c_{n}\|_{0,T}\leq Ch_{T}\|u\|_{\widetilde{H}^{2}(T)}. (4.15)

Let

𝒓ct𝒕Γ+β¯T1cn𝒏Γ.\boldsymbol{r}\coloneqq c_{t}\boldsymbol{t}_{\Gamma}+\overline{\beta}_{T}^{-1}c_{n}\boldsymbol{n}_{\Gamma}.

Then 𝒓^1(T)\boldsymbol{r}\in\nabla\widehat{\mathbb{P}}_{1}(T). By (4.13), (4.15), and (4.2),

u𝒓0,T\displaystyle\|\nabla u-\boldsymbol{r}\|_{0,T} \displaystyle\leq u𝒕Γct0,T+β1cnβ¯Tu𝒏Γ0,T\displaystyle\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-c_{t}\|_{0,T}+\beta_{*}^{-1}\|c_{n}-\overline{\beta}_{T}\nabla u\cdot\boldsymbol{n}_{\Gamma}\|_{0,T} (4.16)
\displaystyle\leq ChTuH~2(T)+β1cnβu𝒏Γ0,T+β1(ββ¯T)u0,T\displaystyle Ch_{T}\|u\|_{\widetilde{H}^{2}(T)}+\beta_{*}^{-1}\|c_{n}-\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}\|_{0,T}+\beta_{*}^{-1}\|(\beta-\overline{\beta}_{T})\nabla u\|_{0,T}
\displaystyle\leq ChTuH~2(T).\displaystyle Ch_{T}\|u\|_{\widetilde{H}^{2}(T)}.

Since 𝒓^1(T)\boldsymbol{r}\in\nabla\widehat{\mathbb{P}}_{1}(T), there exists q^1(T)q\in\widehat{\mathbb{P}}_{1}(T) such that q=𝒓\nabla q=\boldsymbol{r} and Tqd𝒙=Tud𝒙\int_{T}q\mathop{}\!\mathrm{d}\boldsymbol{x}=\int_{T}u\mathop{}\!\mathrm{d}\boldsymbol{x}. Then (4.16) and Poincaré-Friedrichs inequality (cf. [9]) imply that

uq0,TChT|uq|1,TChT2uH~2(T).\|u-q\|_{0,T}\leq Ch_{T}|u-q|_{1,T}\leq Ch_{T}^{2}\|u\|_{\widetilde{H}^{2}(T)}.

This completes the proof of the theorem. ∎

As a corollary, we obtain the estimate for the L2L^{2}-projection Q0Q_{0} onto the space ^1\widehat{\mathbb{P}}_{1} as follows.

Corollary 4.6.

There exists a positive constant CC, depending only on ρ\rho and β\beta, such that

uQ0u0,Ω+h|uQ0u|1,ΩCh2uH~2(Ω)uH~Γ2(Ω).\|u-Q_{0}u\|_{0,\Omega}+h|u-Q_{0}u|_{1,\Omega}\leq Ch^{2}\|u\|_{\widetilde{H}^{2}(\Omega)}\quad\forall u\in\widetilde{H}_{\Gamma}^{2}(\Omega).
Proof.

Let T𝒯hT\in\mathcal{T}_{h} be an interface element. By Theorem 4.5, there exists q^1(T)q^{\prime}\in\widehat{\mathbb{P}}_{1}(T) such that

uq0,T+hT|uq|1,TChT2uH~2(T),\|u-q^{\prime}\|_{0,T}+h_{T}|u-q^{\prime}|_{1,T}\leq Ch_{T}^{2}\|u\|_{\widetilde{H}^{2}(T)}, (4.17)

where CC is a positive constant depending only on ρ\rho and β\beta. Since Q0v0,Tv0,T\|Q_{0}v\|_{0,T}\leq\|v\|_{0,T} for any vH1(T)v\in H^{1}(T) and Q0q=qQ_{0}q=q for any q^1(T)q\in\widehat{\mathbb{P}}_{1}(T), we obtain

uQ0u0,Tuq0,T+Q0qQ0u0,TChT2uH~2(T).\|u-Q_{0}u\|_{0,T}\leq\|u-q^{\prime}\|_{0,T}+\|Q_{0}q^{\prime}-Q_{0}u\|_{0,T}\leq Ch_{T}^{2}\|u\|_{\widetilde{H}^{2}(T)}.

By Lemma 4.4,

|uQ0u|1,T\displaystyle|u-Q_{0}u|_{1,T} \displaystyle\leq |uq|1,T+|Q0qQ0u|1,T|uq|1,T+hT1Q0qQ0u0,T\displaystyle|u-q^{\prime}|_{1,T}+|Q_{0}q^{\prime}-Q_{0}u|_{1,T}\leq|u-q^{\prime}|_{1,T}+h_{T}^{-1}\|Q_{0}q^{\prime}-Q_{0}u\|_{0,T}
\displaystyle\leq |uq|1,T+hT1qu0,TChTuH~2(T).\displaystyle|u-q^{\prime}|_{1,T}+h_{T}^{-1}\|q^{\prime}-u\|_{0,T}\leq Ch_{T}\|u\|_{\widetilde{H}^{2}(T)}.

This completes the proof. ∎

The following lemma gives the L2L^{2}-norm estimate of βuβ¯T(Q0u)\beta\nabla u-\overline{\beta}_{T}\nabla(Q_{0}u) on each mesh edge (see Proposition 5.2 in [24]).

Lemma 4.7.

There exists a positive constant CC independent of hh such that

T𝒯hβuβ¯T(Q0u)0,T2ChuH~2(Ω)2uH~Γ2(Ω).\sum_{T\in\mathcal{T}_{h}}\left\|\beta\nabla u-\overline{\beta}_{T}\nabla(Q_{0}u)\right\|_{0,\partial T}^{2}\leq Ch\|u\|_{\widetilde{H}^{2}(\Omega)}^{2}\quad\forall u\in\widetilde{H}_{\Gamma}^{2}(\Omega).
Proof.

Let T𝒯hT\in\mathcal{T}_{h} be an interface element. Let q=Q0uq=Q_{0}u, and let eTe\subset\partial T. As in (4.13), we have

u=(u𝒕Γ)𝒕Γ+(u𝒏Γ)𝒏Γ,q=(q𝒕Γ)𝒕Γ+(q𝒏Γ)𝒏Γ\nabla u=(\nabla u\cdot\boldsymbol{t}_{\Gamma})\boldsymbol{t}_{\Gamma}+(\nabla u\cdot\boldsymbol{n}_{\Gamma})\boldsymbol{n}_{\Gamma},\quad\nabla q=(\nabla q\cdot\boldsymbol{t}_{\Gamma})\boldsymbol{t}_{\Gamma}+(\nabla q\cdot\boldsymbol{n}_{\Gamma})\boldsymbol{n}_{\Gamma} (4.18)

on TT. Since uH~2(T)u\in\widetilde{H}^{2}(T), we have u𝒕ΓH1(T)\nabla u\cdot\boldsymbol{t}_{\Gamma}\in H^{1}(T) and βu𝒏ΓH1(T)\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}\in H^{1}(T). Note also that β¯Tq𝒏Γ\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma} and q𝒕Γ\nabla q\cdot\boldsymbol{t}_{\Gamma} are constants on TT. Then, by (4.2),

βuβ¯Tq0,e\displaystyle\left\|\beta\nabla u-\overline{\beta}_{T}\nabla q\right\|_{0,e} \displaystyle\leq βuβ¯Tu0,e+β¯Tuβ¯Tq0,e\displaystyle\left\|\beta\nabla u-\overline{\beta}_{T}\nabla u\right\|_{0,e}+\left\|\overline{\beta}_{T}\nabla u-\overline{\beta}_{T}\nabla q\right\|_{0,e} (4.19)
\displaystyle\leq ChTu0,e+Cuq0,e.\displaystyle Ch_{T}\|\nabla u\|_{0,e}+C\|\nabla u-\nabla q\|_{0,e}.

By the trace inequality (4.6) and (4.14),

u0,e\displaystyle\|\nabla u\|_{0,e} \displaystyle\leq u𝒕Γ0,e+β1βu𝒏Γ0,e\displaystyle\|\nabla u\cdot\boldsymbol{t}_{\Gamma}\|_{0,e}+\beta_{*}^{-1}\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}\|_{0,e} (4.20)
\displaystyle\leq ChT1/2(u𝒕Γ0,T+hT|u𝒕Γ|1,T)+ChT1/2(βu𝒏Γ0,T+hT|βu𝒏Γ|1,T)\displaystyle Ch_{T}^{-1/2}\left(\|\nabla u\cdot\boldsymbol{t}_{\Gamma}\|_{0,T}+h_{T}|\nabla u\cdot\boldsymbol{t}_{\Gamma}|_{1,T}\right)+Ch_{T}^{-1/2}\left(\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}\|_{0,T}+h_{T}|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}|_{1,T}\right)
\displaystyle\leq ChT1/2|u|1,T+ChT1/2uH~2(T)\displaystyle Ch_{T}^{-1/2}|u|_{1,T}+Ch_{T}^{1/2}\|u\|_{\widetilde{H}^{2}(T)}
\displaystyle\leq ChT1/2uH~2(T).\displaystyle Ch_{T}^{-1/2}\|u\|_{\widetilde{H}^{2}(T)}.

By (4.18) and (4.2),

uq0,e\displaystyle\|\nabla u-\nabla q\|_{0,e} \displaystyle\leq u𝒕Γq𝒕Γ0,e+β1β¯Tu𝒏Γβ¯Tq𝒏Γ0,e\displaystyle\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-\nabla q\cdot\boldsymbol{t}_{\Gamma}\|_{0,e}+\beta_{*}^{-1}\|\overline{\beta}_{T}\nabla u\cdot\boldsymbol{n}_{\Gamma}-\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma}\|_{0,e} (4.21)
\displaystyle\leq u𝒕Γq𝒕Γ0,e+β1(β¯Tβ)u𝒏Γ0,e\displaystyle\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-\nabla q\cdot\boldsymbol{t}_{\Gamma}\|_{0,e}+\beta_{*}^{-1}\|(\overline{\beta}_{T}-\beta)\nabla u\cdot\boldsymbol{n}_{\Gamma}\|_{0,e}
+β1βu𝒏Γβ¯Tq𝒏Γ0,e\displaystyle+\beta_{*}^{-1}\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma}\|_{0,e}
\displaystyle\leq u𝒕Γq𝒕Γ0,e+Cβ1hTu0,e\displaystyle\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-\nabla q\cdot\boldsymbol{t}_{\Gamma}\|_{0,e}+C\beta_{*}^{-1}h_{T}\|\nabla u\|_{0,e}
+β1βu𝒏Γβ¯Tq𝒏Γ0,e.\displaystyle+\beta_{*}^{-1}\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma}\|_{0,e}.

By the trace inequality (4.6), Corollary 4.6, and (4.2),

u𝒕Γq𝒕Γ0,e\displaystyle\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-\nabla q\cdot\boldsymbol{t}_{\Gamma}\|_{0,e} \displaystyle\leq ChT1/2u𝒕Γq𝒕Γ0,T+ChT1/2|u𝒕Γq𝒕Γ|1,T\displaystyle Ch_{T}^{-1/2}\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-\nabla q\cdot\boldsymbol{t}_{\Gamma}\|_{0,T}+Ch_{T}^{1/2}|\nabla u\cdot\boldsymbol{t}_{\Gamma}-\nabla q\cdot\boldsymbol{t}_{\Gamma}|_{1,T} (4.22)
\displaystyle\leq ChT1/2u𝒕Γq𝒕Γ0,T+ChT1/2|u𝒕Γ|1,T\displaystyle Ch_{T}^{-1/2}\|\nabla u\cdot\boldsymbol{t}_{\Gamma}-\nabla q\cdot\boldsymbol{t}_{\Gamma}\|_{0,T}+Ch_{T}^{1/2}|\nabla u\cdot\boldsymbol{t}_{\Gamma}|_{1,T}
\displaystyle\leq ChT1/2uH~2(T),\displaystyle Ch_{T}^{1/2}\|u\|_{\widetilde{H}^{2}(T)},
βu𝒏Γβ¯Tq𝒏Γ0,e\displaystyle\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma}\|_{0,e} \displaystyle\leq ChT1/2βu𝒏Γβ¯Tq𝒏Γ0,T+ChT1/2|βu𝒏Γβ¯Tq𝒏Γ|1,T\displaystyle Ch_{T}^{-1/2}\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma}\|_{0,T}+Ch_{T}^{1/2}|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma}|_{1,T} (4.23)
\displaystyle\leq ChT1/2βu𝒏Γβ¯Tq𝒏Γ0,T+ChT1/2|βu𝒏Γ|1,T\displaystyle Ch_{T}^{-1/2}\|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}-\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{\Gamma}\|_{0,T}+Ch_{T}^{1/2}|\beta\nabla u\cdot\boldsymbol{n}_{\Gamma}|_{1,T}
\displaystyle\leq ChT1/2(ββ¯T)u0,T+ChT1/2uq0,T+ChT1/2uH~2(T)\displaystyle Ch_{T}^{-1/2}\|(\beta-\overline{\beta}_{T})\nabla u\|_{0,T}+Ch_{T}^{-1/2}\|\nabla u-\nabla q\|_{0,T}+Ch_{T}^{1/2}\|u\|_{\widetilde{H}^{2}(T)}
\displaystyle\leq ChT1/2uH~2(T).\displaystyle Ch_{T}^{1/2}\|u\|_{\widetilde{H}^{2}(T)}.

Now the conclusion follows from the inequalities (4.19)-(4.23). ∎

The following lemma gives the L2L^{2}-norm estimate of w(Qhu)(Q0u)\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u) on each element in 𝒯h\mathcal{T}_{h}.

Lemma 4.8.

There exists a positive constant CC independent of hh such that

w(Qhu)(Q0u)0,ΩChuH~2(Ω)uH~Γ2(Ω).\|\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u)\|_{0,\Omega}\leq Ch\|u\|_{\widetilde{H}^{2}(\Omega)}\quad u\in\widetilde{H}_{\Gamma}^{2}(\Omega).
Proof.

Let TT be an interface element. By the definition of the discrete weak gradient (3.2), we have

Tβ¯T(w(Qhu)(Q0u))qd𝒙=T(Q(Q0u)Qu)(β¯Tq𝒏)dsq^1(T).\int_{T}\overline{\beta}_{T}\left(\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u)\right)\cdot\nabla q\mathop{}\!\mathrm{d}\boldsymbol{x}=-\int_{\partial T}(Q_{\partial}(Q_{0}u)-Q_{\partial}u)\left(\overline{\beta}_{T}\frac{\partial q}{\partial\boldsymbol{n}}\right)\mathop{}\!\mathrm{d}s\quad\forall q\in\widehat{\mathbb{P}}_{1}(T).

Let q^1(T)q\in\widehat{\mathbb{P}}_{1}(T) satisfy q=w(Qhu)(Q0u)\nabla q=\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u). By the trace inequality (4.6), Lemma 4.3, Poincaré-Friedrichs inequality, and Corollary 4.6, we obtain

w(Qhu)(Q0u)0,Ω2CT𝒯huQ0u0,Tβ¯Tq0,T\displaystyle\|\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u)\|_{0,\Omega}^{2}\leq C\sum_{T\in\mathcal{T}_{h}}\|u-Q_{0}u\|_{0,\partial T}\left\|\overline{\beta}_{T}\nabla q\right\|_{0,\partial T}
CT𝒯h(hT1uQ0u0,T+|uQ0u|1,T)β¯T1/2q0,T\displaystyle\qquad\leq C\sum_{T\in\mathcal{T}_{h}}\left(h_{T}^{-1}\|u-Q_{0}u\|_{0,T}+|u-Q_{0}u|_{1,T}\right)\big{\|}\overline{\beta}_{T}^{1/2}\nabla q\big{\|}_{0,T}
CT𝒯h|uQ0u|1,Tβ¯T1/2q0,T\displaystyle\qquad\leq C\sum_{T\in\mathcal{T}_{h}}|u-Q_{0}u|_{1,T}\big{\|}\overline{\beta}_{T}^{1/2}\nabla q\big{\|}_{0,T}
ChuH~2(Ω)w(Qhu)(Q0u)0,Ω,\displaystyle\qquad\leq Ch\|u\|_{\widetilde{H}^{2}(\Omega)}\|\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u)\|_{0,\Omega},

and this completes the proof. ∎

5. Error Analysis

In this section, we present the error estimate in the discrete H1H^{1}-seminorm for the scheme (3.3).

5.1. Discrete H1H^{1}-seminorm

We introduce a discrete H1H^{1}-seminorm as follows: For vh={v0,v}Vhv_{h}=\{v_{0},v_{\partial}\}\in V_{h},

|vh|1,h(T𝒯hv00,T2+λhT1Qv0v0,T2)1/2.|v_{h}|_{1,h}\coloneqq\left(\sum_{T\in\mathcal{T}_{h}}\|\nabla v_{0}\|_{0,T}^{2}+\lambda h_{T}^{-1}\|Q_{\partial}v_{0}-v_{\partial}\|_{0,\partial T}^{2}\right)^{1/2}.

The following lemma shows that two seminorms ||||||{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\cdot\right|\kern-1.07639pt\right|\kern-1.07639pt\right|} and ||1,h|\cdot|_{1,h} on VhV_{h} are equivalent.

Lemma 5.1.

There exist two positive constants C1C_{1} and C2C_{2} independent of hh such that

C1|vh|1,h|vh|C2|vh|1,hvhVh.C_{1}|v_{h}|_{1,h}\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\leq C_{2}|v_{h}|_{1,h}\quad\forall v_{h}\in V_{h}.
Proof.

The proof is similar to the proof of Lemma 5.3 in [35]. Let vh={v0,v}Vhv_{h}=\{v_{0},v_{\partial}\}\in V_{h}. By the definition of the discrete weak gradient (3.2), we have

Tβ¯Twvhqd𝒙=Tβ¯Tv0qd𝒙+T(vQv0)(β¯Tq𝒏T)dsq^1(T).\int_{T}\overline{\beta}_{T}\nabla_{w}v_{h}\cdot\nabla q\mathop{}\!\mathrm{d}\boldsymbol{x}=\int_{T}\overline{\beta}_{T}\nabla v_{0}\cdot\nabla q\mathop{}\!\mathrm{d}\boldsymbol{x}+\int_{\partial T}\left(v_{\partial}-Q_{\partial}v_{0}\right)\left(\overline{\beta}_{T}\nabla q\cdot\boldsymbol{n}_{T}\right)\mathop{}\!\mathrm{d}s\quad\forall q\in\widehat{\mathbb{P}}_{1}(T). (5.1)

Let q^1(Ω)q\in\widehat{\mathbb{P}}_{1}(\Omega) satisfy q=wvh\nabla q=\nabla_{w}v_{h} on each T𝒯hT\in\mathcal{T}_{h}. Then, by Lemma 4.3,

β¯1/2wvh0,Ω2=T𝒯h(Tβ¯Tv0wvhd𝒙+T(vQv0)(β¯Twvh𝒏T)ds)\displaystyle\big{\|}\overline{\beta}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,\Omega}^{2}=\sum_{T\in\mathcal{T}_{h}}\left(\int_{T}\overline{\beta}_{T}\nabla v_{0}\cdot\nabla_{w}v_{h}\mathop{}\!\mathrm{d}\boldsymbol{x}+\int_{\partial T}\left(v_{\partial}-Q_{\partial}v_{0}\right)\left(\overline{\beta}_{T}\nabla_{w}v_{h}\cdot\boldsymbol{n}_{T}\right)\mathop{}\!\mathrm{d}s\right)
CT𝒯h(v00,Tβ¯T1/2wvh0,T+Qv0v0,Tβ¯Twvh0,T)\displaystyle\qquad\leq C\sum_{T\in\mathcal{T}_{h}}\left(\|\nabla v_{0}\|_{0,T}\big{\|}\overline{\beta}_{T}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,T}+\|Q_{\partial}v_{0}-v_{\partial}\|_{0,\partial T}\|\overline{\beta}_{T}\nabla_{w}v_{h}\|_{0,\partial T}\right)
CT𝒯h(v00,Tβ¯T1/2wvh0,T+Ch1/2Qv0v0,Tβ¯T1/2wvh0,T)\displaystyle\qquad\leq C\sum_{T\in\mathcal{T}_{h}}\left(\|\nabla v_{0}\|_{0,T}\big{\|}\overline{\beta}_{T}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,T}+Ch^{-1/2}\|Q_{\partial}v_{0}-v_{\partial}\|_{0,\partial T}\big{\|}\overline{\beta}_{T}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,T}\right)
C|vh|1,hβ¯1/2wvh0,Ω.\displaystyle\qquad\leq C|v_{h}|_{1,h}\big{\|}\overline{\beta}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,\Omega}.

Thus we have β¯1/2wvh0,ΩC|vh|1,h\|\overline{\beta}^{1/2}\nabla_{w}v_{h}\|_{0,\Omega}\leq C|v_{h}|_{1,h}. Since s(vh,vh)|vh|1,h2s(v_{h},v_{h})\leq|v_{h}|_{1,h}^{2}, we have

|vh|2=β¯1/2wvh0,Ω2+s(vh,vh)C|vh|1,h2.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}^{2}=\big{\|}\overline{\beta}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,\Omega}^{2}+s(v_{h},v_{h})\leq C|v_{h}|_{1,h}^{2}.

On the other hand, let q^1(Ω)q\in\widehat{\mathbb{P}}_{1}(\Omega) satisfy q=v0\nabla q=\nabla v_{0} on each T𝒯hT\in\mathcal{T}_{h}. Then, by (5.1) and Lemma 4.3 we have

v00,Ω2\displaystyle\|\nabla v_{0}\|_{0,\Omega}^{2} \displaystyle\leq CT𝒯hTβ¯Tv0v0d𝒙\displaystyle C\sum_{T\in\mathcal{T}_{h}}\int_{T}\overline{\beta}_{T}\nabla v_{0}\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}
=\displaystyle= CT𝒯h(Tβ¯Twvhv0d𝒙T(vQv0)(β¯Tv0𝒏T)ds)\displaystyle C\sum_{T\in\mathcal{T}_{h}}\left(\int_{T}\overline{\beta}_{T}\nabla_{w}v_{h}\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}-\int_{\partial T}(v_{\partial}-Q_{\partial}v_{0})\left(\overline{\beta}_{T}\nabla v_{0}\cdot\boldsymbol{n}_{T}\right)\mathop{}\!\mathrm{d}s\right)
\displaystyle\leq CT𝒯h(β¯T1/2wvh0,Tv00,T+vQv00,Tβ¯Tv0𝒏T0,T)\displaystyle C\sum_{T\in\mathcal{T}_{h}}\left(\big{\|}\overline{\beta}_{T}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,T}\|\nabla v_{0}\|_{0,T}+\|v_{\partial}-Q_{\partial}v_{0}\|_{0,\partial T}\left\|\overline{\beta}_{T}\nabla v_{0}\cdot\boldsymbol{n}_{T}\right\|_{0,\partial T}\right)
\displaystyle\leq CT𝒯h(β¯T1/2wvh0,Tv00,T+hT1/2vQv00,Tv00,T)\displaystyle C\sum_{T\in\mathcal{T}_{h}}\left(\big{\|}\overline{\beta}_{T}^{1/2}\nabla_{w}v_{h}\big{\|}_{0,T}\|\nabla v_{0}\|_{0,T}+h_{T}^{-1/2}\|v_{\partial}-Q_{\partial}v_{0}\|_{0,\partial T}\left\|\nabla v_{0}\right\|_{0,T}\right)
\displaystyle\leq C|||vh|||v00,Ω.\displaystyle C{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\|\nabla v_{0}\|_{0,\Omega}.

Thus v00,ΩC|||vh|||\|\nabla v_{0}\|_{0,\Omega}\leq C{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}. Since s(vh,vh)|||vh|||2s(v_{h},v_{h})\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}^{2}, we obtain

|vh|1,h2=v00,Ω2+s(vh,vh)C|||vh|||2.|v_{h}|_{1,h}^{2}=\|\nabla v_{0}\|_{0,\Omega}^{2}+s(v_{h},v_{h})\leq C{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}^{2}.

Hence we have proved the lemma. ∎

5.2. Error equation

The error equation presented in the following lemma will be used to derive the error estimate.

Lemma 5.2.

Let uH01(Ω)u\in H_{0}^{1}(\Omega) be the solution of (2.3) with fL2(Ω)f\in L^{2}(\Omega), and let uhVh,0u_{h}\in V_{h,0} be the solution of (3.3). Then we have

as(Qhuuh,vh)\displaystyle a_{s}(Q_{h}u-u_{h},v_{h}) =\displaystyle= s(Qhu,vh)+T𝒯hT(β¯Tw(Qhu)βu)v0d𝒙\displaystyle s(Q_{h}u,v_{h})+\sum_{T\in\mathcal{T}_{h}}\int_{T}(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\beta\nabla u)\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x} (5.2)
+T𝒯hT(Qv0v)(β¯Tw(Qhu)βu)𝒏Tds\displaystyle+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\left(Q_{\partial}v_{0}-v_{\partial}\right)\left(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\beta\nabla u\right)\cdot\boldsymbol{n}_{T}\mathop{}\!\mathrm{d}s
+T𝒯hT(v0Qv0)βu𝒏ds,vhVh,0.\displaystyle+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\left(v_{0}-Q_{\partial}v_{0}\right)\beta\frac{\partial u}{\partial\boldsymbol{n}}\mathop{}\!\mathrm{d}s,\quad\forall v_{h}\in V_{h,0}.
Proof.

Note that, for any vhVh,0v_{h}\in V_{h,0},

as(Qhuuh,vh)=\displaystyle a_{s}(Q_{h}u-u_{h},v_{h})= as(Qhu,vh)(f,v0)0,Ω\displaystyle\ a_{s}(Q_{h}u,v_{h})-(f,v_{0})_{0,\Omega}
=\displaystyle= as(Qhu,vh)T𝒯h(Tβuv0d𝒙Tβu𝒏v0ds)\displaystyle\ a_{s}(Q_{h}u,v_{h})-\sum_{T\in\mathcal{T}_{h}}\left(\int_{T}\beta\nabla u\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}-\int_{\partial T}\beta\frac{\partial u}{\partial\boldsymbol{n}}v_{0}\mathop{}\!\mathrm{d}s\right)
=\displaystyle= T𝒯hTβ¯Tw(Qhu)wvhd𝒙\displaystyle\ \sum_{T\in\mathcal{T}_{h}}\int_{T}\overline{\beta}_{T}\nabla_{w}(Q_{h}u)\cdot\nabla_{w}v_{h}\mathop{}\!\mathrm{d}\boldsymbol{x}
+s(Qhu,vh)T𝒯h(Tβuv0d𝒙Tβu𝒏v0ds)\displaystyle\ +s(Q_{h}u,v_{h})-\sum_{T\in\mathcal{T}_{h}}\left(\int_{T}\beta\nabla u\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}-\int_{\partial T}\beta\frac{\partial u}{\partial\boldsymbol{n}}v_{0}\mathop{}\!\mathrm{d}s\right)
=\displaystyle= T𝒯hTβ¯Tw(Qhu)v0d𝒙T𝒯hT(v0v)(Q(β¯Tw(Qhu)𝒏T))ds\displaystyle\ \sum_{T\in\mathcal{T}_{h}}\int_{T}\overline{\beta}_{T}\nabla_{w}(Q_{h}u)\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}-\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(v_{0}-v_{\partial})\left(Q_{\partial}\left(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)\cdot\boldsymbol{n}_{T}\right)\right)\mathop{}\!\mathrm{d}s
+s(Qhu,vh)T𝒯h(Tβuv0d𝒙Tβu𝒏v0ds)\displaystyle\ +s(Q_{h}u,v_{h})-\sum_{T\in\mathcal{T}_{h}}\left(\int_{T}\beta\nabla u\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}-\int_{\partial T}\beta\frac{\partial u}{\partial\boldsymbol{n}}v_{0}\mathop{}\!\mathrm{d}s\right)
=\displaystyle= s(Qhu,vh)+T𝒯hT(β¯Tw(Qhu)βu)v0d𝒙\displaystyle\ s(Q_{h}u,v_{h})+\sum_{T\in\mathcal{T}_{h}}\int_{T}(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\beta\nabla u)\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}
T𝒯hT(v0v)(Q(β¯Tw(Qhu)𝒏T))ds+T𝒯hTβu𝒏v0ds.\displaystyle\ -\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(v_{0}-v_{\partial})\left(Q_{\partial}\left(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)\cdot\boldsymbol{n}_{T}\right)\right)\mathop{}\!\mathrm{d}s+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\beta\frac{\partial u}{\partial\boldsymbol{n}}v_{0}\mathop{}\!\mathrm{d}s.

Since [βu𝒏]e=0\left[\beta\frac{\partial u}{\partial\boldsymbol{n}}\right]_{e}=0 for each interior edge ee and v|e=0v_{\partial}|_{e}=0 for each boundary edge ee, we obtain

T𝒯hTβu𝒏v0ds\displaystyle\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\beta\frac{\partial u}{\partial\boldsymbol{n}}v_{0}\mathop{}\!\mathrm{d}s =\displaystyle= T𝒯hTβu𝒏(v0v)ds\displaystyle\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\beta\frac{\partial u}{\partial\boldsymbol{n}}(v_{0}-v_{\partial})\mathop{}\!\mathrm{d}s
=\displaystyle= T𝒯hT(v0v)βu𝒏dsT𝒯hT(v0v)(Q(βu𝒏T))ds\displaystyle\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(v_{0}-v_{\partial})\beta\frac{\partial u}{\partial\boldsymbol{n}}\mathop{}\!\mathrm{d}s-\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(v_{0}-v_{\partial})(Q_{\partial}(\beta\nabla u\cdot\boldsymbol{n}_{T}))\mathop{}\!\mathrm{d}s
+T𝒯hT(v0v)(Q(βu𝒏T))ds\displaystyle+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(v_{0}-v_{\partial})(Q_{\partial}(\beta\nabla u\cdot\boldsymbol{n}_{T}))\mathop{}\!\mathrm{d}s
=\displaystyle= T𝒯hT(v0Qv0)βu𝒏ds+T𝒯hT(v0v)(Q(βu𝒏T))ds.\displaystyle\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(v_{0}-Q_{\partial}v_{0})\beta\frac{\partial u}{\partial\boldsymbol{n}}\mathop{}\!\mathrm{d}s+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(v_{0}-v_{\partial})(Q_{\partial}(\beta\nabla u\cdot\boldsymbol{n}_{T}))\mathop{}\!\mathrm{d}s.

Using the above equation we obtain

as(Qhuuh,vh)=\displaystyle a_{s}(Q_{h}u-u_{h},v_{h})= s(Qhu,vh)+T𝒯hT(β¯Tw(Qhu)βu)v0d𝒙\displaystyle\ s(Q_{h}u,v_{h})+\sum_{T\in\mathcal{T}_{h}}\int_{T}(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\beta\nabla u)\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}
+T𝒯hT(Qv0v)(β¯Tw(Qhu)βu)𝒏Tds\displaystyle\ +\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\left(Q_{\partial}v_{0}-v_{\partial}\right)\left(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\beta\nabla u\right)\cdot\boldsymbol{n}_{T}\mathop{}\!\mathrm{d}s
+T𝒯hT(v0Qv0)βu𝒏ds.\displaystyle\ +\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\left(v_{0}-Q_{\partial}v_{0}\right)\beta\frac{\partial u}{\partial\boldsymbol{n}}\mathop{}\!\mathrm{d}s.

This completes the proof of the lemma. ∎

The following lemma can be found in [14].

Lemma 5.3.

Let T𝒯hT\in\mathcal{T}_{h} and let eTe\subset\partial T. Then there exists a positive constant CC independent of hh such that

uu¯e1/2,eCh|u|1,TuH1(T).\|u-\overline{u}_{e}\|_{-1/2,e}\leq Ch|u|_{1,T}\quad\forall u\in H^{1}(T).

where u¯e=1|e|euds\overline{u}_{e}=\frac{1}{|e|}\int_{e}u\mathop{}\!\mathrm{d}s.

5.3. Error estimate

Now we prove the error estimates in the energy norm and the discrete H1H^{1}-seminorm.

Theorem 5.4.

Suppose that uH~2(Ω)H01(Ω)u\in\widetilde{H}^{2}(\Omega)\cap H_{0}^{1}(\Omega) is the solution of (2.3) with fL2(Ω)f\in L^{2}(\Omega). Suppose further that βuH1(Ω)\beta\nabla u\in H^{1}(\Omega). Let uhVh,0u_{h}\in V_{h,0} be the solution of (3.3). Then there exists a positive constant CC independent of hh such that

|||Qhuuh|||ChuH~2(Ω).{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|Q_{h}u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\leq Ch\|u\|_{\widetilde{H}^{2}(\Omega)}.
Proof.

Let vh=Qhuuhv_{h}=Q_{h}u-u_{h}. From the error equation (5.2), we have

|||Qhuuh|||2\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|Q_{h}u-u_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}^{2} =\displaystyle= as(Qhuuh,vh)\displaystyle\ a_{s}(Q_{h}u-u_{h},v_{h}) (5.3)
=\displaystyle= s(Qhu,vh)+T𝒯hT(β¯Tw(Qhu)βu)v0d𝒙\displaystyle\ s(Q_{h}u,v_{h})+\sum_{T\in\mathcal{T}_{h}}\int_{T}(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\beta\nabla u)\cdot\nabla v_{0}\mathop{}\!\mathrm{d}\boldsymbol{x}
+T𝒯hT(Qv0v)(β¯Tw(Qhu)βu)𝒏Tds\displaystyle+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\left(Q_{\partial}v_{0}-v_{\partial}\right)\left(\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\beta\nabla u\right)\cdot\boldsymbol{n}_{T}\mathop{}\!\mathrm{d}s
+T𝒯hT(v0Qv0)βu𝒏ds\displaystyle+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\left(v_{0}-Q_{\partial}v_{0}\right)\beta\frac{\partial u}{\partial\boldsymbol{n}}\mathop{}\!\mathrm{d}s
\displaystyle\eqqcolon I1+I2+I3+I4.\displaystyle\ I_{1}+I_{2}+I_{3}+I_{4}.

By the trace inequality (4.6), Poincaré-Friedrichs inequality, and Corollary 4.6,

|I1|\displaystyle|I_{1}| \displaystyle\leq CT𝒯hhT1/2uQ0u0,ThT1/2vQv00,T\displaystyle C\sum_{T\in\mathcal{T}_{h}}h_{T}^{-1/2}\|u-Q_{0}u\|_{0,\partial T}h_{T}^{-1/2}\|v_{\partial}-Q_{\partial}v_{0}\|_{0,\partial T} (5.4)
\displaystyle\leq CT𝒯h(hT1uQ0u0,T+|uQ0u|1,T)hT1/2vQv00,T\displaystyle C\sum_{T\in\mathcal{T}_{h}}\left(h_{T}^{-1}\|u-Q_{0}u\|_{0,T}+|u-Q_{0}u|_{1,T}\right)h_{T}^{-1/2}\|v_{\partial}-Q_{\partial}v_{0}\|_{0,\partial T}
\displaystyle\leq CT𝒯h|uQ0u|1,ThT1/2vQv00,T\displaystyle C\sum_{T\in\mathcal{T}_{h}}|u-Q_{0}u|_{1,T}h_{T}^{-1/2}\|v_{\partial}-Q_{\partial}v_{0}\|_{0,\partial T}
\displaystyle\leq ChuH~2(Ω)|||vh|||.\displaystyle Ch\|u\|_{\widetilde{H}^{2}(\Omega)}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}.

From (4.2), Lemma 4.8, and Lemma 5.1,

|I2|\displaystyle|I_{2}| \displaystyle\leq T𝒯h(β¯Tw(Qhu)β¯Tu0,T+(β¯Tβ)u0,T)v00,T\displaystyle\sum_{T\in\mathcal{T}_{h}}\left(\|\overline{\beta}_{T}\nabla_{w}(Q_{h}u)-\overline{\beta}_{T}\nabla u\|_{0,T}+\|(\overline{\beta}_{T}-\beta)\nabla u\|_{0,T}\right)\|\nabla v_{0}\|_{0,T} (5.5)
\displaystyle\leq ChuH~2(Ω)|||vh|||.\displaystyle Ch\|u\|_{\widetilde{H}^{2}(\Omega)}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}.

Since w(Qhu),(Q0u)^1(T)\nabla_{w}(Q_{h}u),\nabla(Q_{0}u)\in\widehat{\mathbb{P}}_{1}(T) on each T𝒯hT\in\mathcal{T}_{h}, using Lemma 4.3, Lemma 4.7, and Lemma 4.8, we have

|I3|\displaystyle|I_{3}| \displaystyle\leq T𝒯hhT1/2vQv00,ThT1/2(β¯T(w(Qhu)(Q0u))0,T+β¯T(Q0u)βu0,T)\displaystyle\sum_{T\in\mathcal{T}_{h}}h_{T}^{-1/2}\|v_{\partial}-Q_{\partial}v_{0}\|_{0,\partial T}h_{T}^{1/2}\left(\left\|\overline{\beta}_{T}\left(\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u)\right)\right\|_{0,\partial T}+\left\|\overline{\beta}_{T}\nabla(Q_{0}u)-\beta\nabla u\right\|_{0,\partial T}\right) (5.6)
\displaystyle\leq C|||vh|||(T𝒯hhT(β¯T(w(Qhu)(Q0u))0,T2+β¯T(Q0u)βu0,T2))1/2\displaystyle C{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\left(\sum_{T\in\mathcal{T}_{h}}h_{T}\left(\left\|\overline{\beta}_{T}\left(\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u)\right)\right\|_{0,\partial T}^{2}+\left\|\overline{\beta}_{T}\nabla(Q_{0}u)-\beta\nabla u\right\|_{0,\partial T}^{2}\right)\right)^{1/2}
\displaystyle\leq C|||vh|||(T𝒯h(β¯T1/2(w(Qhu)(Q0u))0,T2+hTβ¯T(Q0u)βu0,T2))1/2\displaystyle C{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\left(\sum_{T\in\mathcal{T}_{h}}\left(\big{\|}\overline{\beta}_{T}^{1/2}\left(\nabla_{w}(Q_{h}u)-\nabla(Q_{0}u)\right)\big{\|}_{0,T}^{2}+h_{T}\left\|\overline{\beta}_{T}\nabla(Q_{0}u)-\beta\nabla u\right\|_{0,\partial T}^{2}\right)\right)^{1/2}
\displaystyle\leq ChuH~2(Ω)|||vh|||.\displaystyle Ch\|u\|_{\widetilde{H}^{2}(\Omega)}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}.

Let T𝒯hT\in\mathcal{T}_{h} and eTe\subset\partial T. Since βuH1(Ω)\beta\nabla u\in H^{1}(\Omega), by Lemma 5.3 and the trace theorem, we have

|e(v0Qv0)βu𝒏ds|\displaystyle\left|\int_{e}(v_{0}-Q_{\partial}v_{0})\beta\frac{\partial u}{\partial\boldsymbol{n}}\mathop{}\!\mathrm{d}s\right| \displaystyle\leq βu1/2,ev0Qv01/2,e\displaystyle\|\beta\nabla u\|_{1/2,e}\|v_{0}-Q_{\partial}v_{0}\|_{-1/2,e}
\displaystyle\leq ChuH~2(T)|v0|1,T\displaystyle Ch\|u\|_{\widetilde{H}^{2}(T)}|\nabla v_{0}|_{1,T}

Thus we obtain from Remark 2.3 that

|I4|\displaystyle|I_{4}| \displaystyle\leq T𝒯heT|e(v0Qv0)βu𝒏ds|\displaystyle\sum_{T\in\mathcal{T}_{h}}\sum_{e\subset\partial T}\left|\int_{e}(v_{0}-Q_{\partial}v_{0})\beta\frac{\partial u}{\partial\boldsymbol{n}}\mathop{}\!\mathrm{d}s\right| (5.7)
\displaystyle\leq ChT𝒯huH~2(T)|v0|1,T\displaystyle Ch\sum_{T\in\mathcal{T}_{h}}\|u\|_{\widetilde{H}^{2}(T)}|\nabla v_{0}|_{1,T}
\displaystyle\leq Ch|||vh|||uH~2(Ω).\displaystyle Ch{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}\|u\|_{\widetilde{H}^{2}(\Omega)}.

Now combining the inequalities (5.3)-(5.7) we have

|||uhQhu|||2ChuH~2(Ω)|||vh|||=ChuH~2(Ω)|||uhQhu|||.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u_{h}-Q_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}^{2}\leq Ch\|u\|_{\widetilde{H}^{2}(\Omega)}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|v_{h}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}=Ch\|u\|_{\widetilde{H}^{2}(\Omega)}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|u_{h}-Q_{h}u\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}.

This concludes the proof of the theorem. ∎

Using Lemma 5.1 and Theorem 5.4, we immediately obtain the following discrete H1H^{1}-seminorm error estimate.

Corollary 5.5.

Suppose that uH~2(Ω)H01(Ω)u\in\widetilde{H}^{2}(\Omega)\cap H_{0}^{1}(\Omega) is the solution of (2.3) with fL2(Ω)f\in L^{2}(\Omega). Suppose further that βuH1(Ω)\beta\nabla u\in H^{1}(\Omega). Let uhVh,0u_{h}\in V_{h,0} be the solution of (3.3). Then there exists a positive constant CC independent of hh such that

|uhQhu|1,hChuH~2(Ω).|u_{h}-Q_{h}u|_{1,h}\leq Ch\|u\|_{\widetilde{H}^{2}(\Omega)}.

6. Numerical Examples

In this section, we report several numerical results. We solve the problem (2.1)-(2.2) with Ω=[0,1]2\Omega=[0,1]^{2} partitioned into two different families of meshes as follows:

  1. (i)

    M1: uniform square meshes with h=1/23,1/24,,1/28h=1/2^{3},1/2^{4},\cdots,1/2^{8},

  2. (ii)

    M2: unstructured polygonal meshes with h=1/23,1/24,,1/28h=1/2^{3},1/2^{4},\cdots,1/2^{8}.

Some examples of the meshes are shown in Figure 4. The unstructured polygonal meshes are generated from PolyMesher [38]. Let uu be the exact solution and let uh={u0,u}u_{h}=\{u_{0},u_{\partial}\} be the solution of our immersed WG method. We compute errors in the discrete H1H^{1}-seminorm and L2L^{2}-norm, which are given by

|uhQhu|1,h,u0Q0u0,Ω,|u_{h}-Q_{h}u|_{1,h},\quad\|u_{0}-Q_{0}u\|_{0,\Omega},

respectively. For the examples below, the discrete H1H^{1}-seminorm error converges with order O(h)O(h), which agrees with our theoretical result. Moreover, the results show the O(h2)O(h^{2}) error in discrete L2L^{2}-norm, which is optimal.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4. The meshes M1 (left) and M2 (right).
Example 6.1 (Circular interface).

Take a circle centered at (0.5,0.5)(0.5,0.5) with radius r0=0.4r_{0}=0.4 as an interface, and choose the following exact solution

u(x,y)={1β+(r2r02)3if (x,y)Ω+,1β(r2r02)3if (x,y)Ω,u(x,y)=\left\{\begin{array}[]{ll}\frac{1}{\beta^{+}}(r^{2}-r_{0}^{2})^{3}&\textrm{if $(x,y)\in\Omega^{+}$},\\ \frac{1}{\beta^{-}}(r^{2}-r_{0}^{2})^{3}&\textrm{if $(x,y)\in\Omega^{-}$},\end{array}\right.

where β+\beta^{+} and β\beta^{-} are constants and r=(x0.5)2+(y0.5)2r=\sqrt{(x-0.5)^{2}+(y-0.5)^{2}}. In this example, we consider two cases when (β+,β)=(1,10)(\beta^{+},\beta^{-})=(1,10), (10,1)(10,1), (1,1000)(1,1000), and (1000,1)(1000,1). The results are reported in Tables 1 to 4.

Example 6.2 (Sharp edge).

In this example, we consider an interface with sharp edge. Let L(x,y)=(2y1)2+((2x2)tanθ)2(2x1)L(x,y)=-(2y-1)^{2}+((2x-2)\tan\theta)^{2}(2x-1) be the level-set function, with θ=10\theta=10^{\circ}, and

Γ={(x,y)Ω:L(x,y)=0},Ω+={(x,y)Ω:L(x,y)>0},Ω={(x,y)Ω:L(x,y)<0}.\Gamma=\{(x,y)\in\Omega:L(x,y)=0\},\quad\Omega^{+}=\{(x,y)\in\Omega:L(x,y)>0\},\quad\Omega^{-}=\{(x,y)\in\Omega:L(x,y)<0\}.

Then the interface Γ\Gamma has a sharp corner at (1,0.5)(1,0.5) (see Figure 5). The exact solution is chosen as u=L/βu=L/\beta, where β=1000\beta=1000 on Ω+\Omega^{+} and β=1\beta=1 on Ω\Omega^{-}. The results are reported in Table 5.

Example 6.3 (Variable coefficient).

In this example, we take the level set of L(x,y)=(x0.5)2/r12+(y0.5)2/r221L(x,y)=(x-0.5)^{2}/r_{1}^{2}+(y-0.5)^{2}/r_{2}^{2}-1 with r1=0.25r_{1}=0.25 and r2=0.125r_{2}=0.125 as an interface, that is, we set

Γ={(x,y)Ω:L(x,y)=0},Ω+={(x,y)Ω:L(x,y)>0},Ω={(x,y)Ω:L(x,y)<0}.\Gamma=\{(x,y)\in\Omega:L(x,y)=0\},\quad\Omega^{+}=\{(x,y)\in\Omega:L(x,y)>0\},\quad\Omega^{-}=\{(x,y)\in\Omega:L(x,y)<0\}.

The exact solution is chosen as u=L/βu=L/\beta, where

β(x,y)={1if (x,y)Ω+,1+0.5(2x1)2(2x1)(2y1)+(2y1)2if (x,y)Ω.\beta(x,y)=\left\{\begin{array}[]{ll}1&\textrm{if $(x,y)\in\Omega^{+}$},\\ 1+0.5(2x-1)^{2}-(2x-1)(2y-1)+(2y-1)^{2}&\textrm{if $(x,y)\in\Omega^{-}$}.\end{array}\right.

The results are reported in Table 6.

Table 1. The errors for Example 6.1 with (β+,β)=(1,10)(\beta^{+},\beta^{-})=(1,10).
1/h1/h M1 M2
|uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order |uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order
232^{3} 5.4587e-02 3.3027e-03 5.5979e-02 3.3820e-03
242^{4} 2.6417e-02 1.0471 8.7554e-04 1.9154 2.7216e-02 1.0404 8.8878e-04 1.9280
252^{5} 1.3137e-02 1.0079 2.2319e-04 1.9719 1.3755e-02 0.9845 2.2742e-04 1.9665
262^{6} 6.5602e-03 1.0018 5.6112e-05 1.9919 6.9372e-03 0.9876 5.6936e-05 1.9979
272^{7} 3.2792e-03 1.0004 1.4049e-05 1.9979 3.4921e-03 0.9903 1.4224e-05 2.0011
282^{8} 1.6395e-03 1.0001 3.5135e-06 1.9995 1.7500e-03 0.9967 3.5427e-06 2.0054
Table 2. The errors for Example 6.1 with (β+,β)=(10,1)(\beta^{+},\beta^{-})=(10,1).
1/h1/h M1 M2
|uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order |uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order
232^{3} 5.8617e-02 3.4872e-03 6.0526e-02 3.6934e-03
242^{4} 2.8290e-02 1.0510 9.3409e-04 1.9004 2.9512e-02 1.0362 9.7713e-04 1.9183
252^{5} 1.4189e-02 0.9955 2.4360e-04 1.9391 1.5092e-02 0.9676 2.4927e-04 1.9708
262^{6} 7.1176e-03 0.9953 6.2122e-05 1.9713 7.7258e-03 0.9660 6.2438e-05 1.9972
272^{7} 3.5635e-03 0.9981 1.5643e-05 1.9896 3.9414e-03 0.9710 1.5536e-05 2.0068
282^{8} 1.7825e-03 0.9994 3.9192e-06 1.9969 1.9947e-03 0.9825 3.8632e-06 2.0078
Table 3. The errors for Example 6.1 with (β+,β)=(1,1000)(\beta^{+},\beta^{-})=(1,1000).
1/h1/h M1 M2
|uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order |uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order
232^{3} 5.5356e-02 3.2533e-03 5.8561e-02 3.3911e-03
242^{4} 2.6461e-02 1.0649 8.7488e-04 1.8947 2.7524e-02 1.0893 8.8967e-04 1.9304
252^{5} 1.3148e-02 1.0090 2.2320e-04 1.9707 1.3777e-02 0.9984 2.2758e-04 1.9669
262^{6} 6.5645e-03 1.0021 5.6113e-05 1.9919 6.9396e-03 0.9893 5.6970e-05 1.9981
272^{7} 3.2800e-03 1.0010 1.4049e-05 1.9978 3.4929e-03 0.9904 1.4231e-05 2.0011
282^{8} 1.6398e-03 1.0002 3.5137e-06 1.9994 1.7504e-03 0.9968 3.5445e-06 2.0054
Table 4. The errors for Example 6.1 with (β+,β)=(1000,1)(\beta^{+},\beta^{-})=(1000,1).
1/h1/h M1 M2
|uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order |uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order
232^{3} 1.6219e-01 3.4747e-03 7.4038e-02 3.7966e-03
242^{4} 2.9224e-02 2.4725 8.7313e-04 1.9926 3.1237e-02 1.2450 9.9624e-04 1.9301
252^{5} 1.4104e-02 1.0511 2.2151e-04 1.9788 1.5075e-02 1.0512 2.5707e-04 1.9543
262^{6} 7.0375e-03 1.0029 5.6761e-05 1.9644 7.5627e-03 0.9951 6.4711e-05 1.9901
272^{7} 3.5397e-03 0.9914 1.4787e-05 1.9406 3.7960e-03 0.9944 1.6118e-05 2.0053
282^{8} 1.7847e-03 0.9879 3.8351e-06 1.9470 1.9126e-03 0.9890 3.9855e-06 2.0159
Table 5. The errors for Example 6.2.
1/h1/h M1 M2
|uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order |uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order
232^{3} 6.0360e-01 3.0913e-02 5.6648e-01 3.2621e-02
242^{4} 3.0771e-01 0.9720 7.1992e-03 2.1023 2.8813e-01 0.9753 8.0180e-03 2.0245
252^{5} 1.6788e-01 0.8742 1.7187e-03 2.0665 1.6058e-01 0.8434 1.9645e-03 2.0291
262^{6} 8.3267e-02 1.0116 4.2052e-04 2.0310 7.6565e-02 1.0686 4.7390e-04 2.0515
272^{7} 3.9446e-02 1.0779 1.0327e-04 2.0258 3.5919e-02 1.0919 1.0916e-04 2.1181
282^{8} 1.9396e-02 1.0241 2.5529e-05 2.0162 1.7581e-02 1.0308 2.6011e-05 2.0693
Table 6. The errors for Example 6.3.
1/h1/h M1 M2
|uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order |uhQhu|1,h|u_{h}-Q_{h}u|_{1,h} Order u0Q0u0,Ω\|u_{0}-Q_{0}u\|_{0,\Omega} Order
232^{3} 1.0067e+01 5.8821e-01 9.7421e+00 6.1556e-01
242^{4} 5.0872e+00 0.9847 1.4782e-01 1.9925 4.8530e+00 1.0054 1.5612e-01 1.9793
252^{5} 2.5514e+00 0.9956 3.7083e-02 1.9950 2.4251e+00 1.0009 3.9536e-02 1.9814
262^{6} 1.2772e+00 0.9983 9.2748e-03 1.9994 1.2119e+00 1.0008 9.9313e-03 1.9931
272^{7} 6.3883e-01 0.9995 2.3199e-03 1.9992 6.0612e-01 0.9995 2.4873e-03 1.9974
282^{8} 3.1948e-01 0.9997 5.8000e-04 2.0000 3.0290e-01 1.0007 6.2143e-04 2.0009
Ω\Omega^{-}Ω+\Omega^{+}Γ\Gamma2θ2\theta
Figure 5. The level set of (2y1)2=((2x2)tanθ)2(2x1)(2y-1)^{2}=((2x-2)\tan\theta)^{2}(2x-1).

7. Conclusion

We introduce an immersed WG method for the elliptic interface problems on general unfitted polygonal meshes. The discrete space consists of constant functions on the mesh edges and piecewise linear functions in the mesh elements, satisfying the interface conditions. We prove an optimal-order convergence in the discrete H1H^{1}-seminorm under some assumptions on the exact solution.

References

  • [1] S. Adjerid, N. Chaabane, and T. Lin, An immersed discontinuous finite element method for Stokes interface problems, Comput. Methods Appl. Mech. Engrg., 293 (2015), pp. 170–190.
  • [2] B. Ayuso de Dios, K. Lipnikov, and G. Manzini, The nonconforming virtual element method, ESAIM Math. Model. Numer. Anal., 50 (2016), pp. 879–904.
  • [3] R. Becker, E. Burman, and P. Hansbo, A Nitsche extended finite element method for incompressible elasticity with discontinuous modulus of elasticity, Comput. Methods Appl. Mech. Engrg., 198 (2009), pp. 3352–3360.
  • [4] L. Beirão da Veiga, F. Brezzi, A. Cangiani, G. Manzini, L. D. Marini, and A. Russo, Basic principles of virtual element methods, Math. Models Methods Appl. Sci., 23 (2013), pp. 199–214.
  • [5] T. Belytschko and T. Black, Elastic crack growth in finite elements with minimal remeshing, International journal for numerical methods in engineering, 45 (1999), pp. 601–620.
  • [6] T. Belytschko, C. Parimi, N. Moës, N. Sukumar, and S. Usui, Structured extended finite element methods for solids defined by implicit surfaces, International journal for numerical methods in engineering, 56 (2003), pp. 609–635.
  • [7] J. H. Bramble and J. T. King, A robust finite element method for nonhomogeneous Dirichlet problems in domains with curved boundaries, Math. Comp., 63 (1994), pp. 1–17.
  • [8]  , A finite element method for interface problems in domains with smooth boundaries and interfaces, Adv. Comput. Math., 6 (1996), pp. 109–138.
  • [9] S. C. Brenner, Poincaré-Friedrichs inequalities for piecewise H1H^{1} functions, SIAM J. Numer. Anal., 41 (2003), pp. 306–324.
  • [10] S. C. Brenner, Q. Guan, and L.-Y. Sung, Some estimates for virtual element methods, Comput. Methods Appl. Math., 17 (2017), pp. 553–574.
  • [11] S. C. Brenner and L. R. Scott, The mathematical theory of finite element methods, vol. 15 of Texts in Applied Mathematics, Springer, New York, third ed., 2008.
  • [12] E. Burman, S. Claus, P. Hansbo, M. G. Larson, and A. Massing, CutFEM: discretizing geometry and partial differential equations, Internat. J. Numer. Methods Engrg., 104 (2015), pp. 472–501.
  • [13] Z. Chen and J. Zou, Finite element methods and their convergence for elliptic and parabolic interface problems, Numer. Math., 79 (1998), pp. 175–202.
  • [14] S.-H. Chou, D. Y. Kwak, and K. T. Wee, Optimal convergence analysis of an immersed interface finite element method, Adv. Comput. Math., 33 (2010), pp. 149–168.
  • [15] P. G. Ciarlet, The finite element method for elliptic problems, vol. 40 of Classics in Applied Mathematics, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2002. Reprint of the 1978 original [North-Holland, Amsterdam; MR0520174 (58 #25001)].
  • [16] B. Cockburn, D. A. Di Pietro, and A. Ern, Bridging the hybrid high-order and hybridizable discontinuous Galerkin methods, ESAIM Math. Model. Numer. Anal., 50 (2016), pp. 635–650.
  • [17] D. A. Di Pietro and J. Droniou, The hybrid high-order method for polytopal meshes, vol. 19 of MS&A. Modeling, Simulation and Applications, Springer, Cham, [2020] ©2020. Design, analysis, and applications.
  • [18] D. A. Di Pietro, A. Ern, and S. Lemaire, An arbitrary-order and compact-stencil discretization of diffusion on general meshes based on local reconstruction operators, Comput. Methods Appl. Math., 14 (2014), pp. 461–472.
  • [19] 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), pp. 5537–5552.
  • [20]  , A finite element method for the simulation of strong and weak discontinuities in solid mechanics, Comput. Methods Appl. Mech. Engrg., 193 (2004), pp. 3523–3540.
  • [21] X. Hu, L. Mu, and X. Ye, Weak Galerkin method for the Biot’s consolidation model, Comput. Math. Appl., 75 (2018), pp. 2017–2030.
  • [22] G. Jo, D. Y. Kwak, and Y.-J. Lee, Locally conservative immersed finite element method for elliptic interface problems, J. Sci. Comput., 87 (2021), pp. Paper No. 60, 27.
  • [23] P. Krysl and T. Belytschko, An efficient linear-precision partition of unity basis for unstructured meshless methods, Comm. Numer. Methods Engrg., 16 (2000), pp. 239–255.
  • [24] D. Kwak and J. Lee, A modified P1{P}_{1}-immersed finite element method, International Journal of Pure and Applied Mathematics, 104 (2015), pp. 471–494.
  • [25] D. Y. Kwak, S. Jin, and D. Kyeong, A stabilized P1P_{1}-nonconforming immersed finite element method for the interface elasticity problems, ESAIM Math. Model. Numer. Anal., 51 (2017), pp. 187–207.
  • [26] D. Y. Kwak, K. T. Wee, and K. S. Chang, An analysis of a broken P1P_{1}-nonconforming finite element method for interface problems, SIAM J. Numer. Anal., 48 (2010), pp. 2117–2134.
  • [27] S. Lee, D. Y. Kwak, and I. Sim, Immersed finite element method for eigenvalue problem, J. Comput. Appl. Math., 313 (2017), pp. 410–426.
  • [28] G. Legrain, N. Moës, and E. Verron, Stress analysis around crack tips in finite strain problems using the eXtended finite element method, Internat. J. Numer. Methods Engrg., 63 (2005), pp. 290–314.
  • [29] Z. Li, T. Lin, and X. Wu, New Cartesian grid methods for interface problems using the finite element formulation, Numer. Math., 96 (2003), pp. 61–98.
  • [30] T. Lin, Y. Lin, R. Rogers, and M. L. Ryan, A rectangular immersed finite element space for interface problems, in Scientific computing and applications (Kananaskis, AB, 2000), vol. 7 of Adv. Comput. Theory Pract., Nova Sci. Publ., Huntington, NY, 2001, pp. 107–114.
  • [31] T. Lin, Y. Lin, and X. Zhang, Partially penalized immersed finite element methods for elliptic interface problems, SIAM J. Numer. Anal., 53 (2015), pp. 1121–1144.
  • [32] T. Lin, D. Sheen, and X. Zhang, A nonconforming immersed finite element method for elliptic interface problems, J. Sci. Comput., 79 (2019), pp. 442–463.
  • [33] J. Liu, S. Tavener, and Z. Wang, Lowest-order weak Galerkin finite element method for Darcy flow on convex polygonal meshes, SIAM J. Sci. Comput., 40 (2018), pp. B1229–B1252.
  • [34] N. Moës, J. Dolbow, and T. Belytschko, A finite element method for crack growth without remeshing, Internat. J. Numer. Methods Engrg., 46 (1999), pp. 131–150.
  • [35] L. Mu, J. Wang, and X. Ye, A weak Galerkin finite element method with polynomial reduction, J. Comput. Appl. Math., 285 (2015), pp. 45–58.
  • [36] L. Mu, J. Wang, X. Ye, and S. Zhang, A weak Galerkin finite element method for the Maxwell equations, J. Sci. Comput., 65 (2015), pp. 363–386.
  • [37] L. Mu and X. Zhang, An immersed weak Galerkin method for elliptic interface problems, J. Comput. Appl. Math., 362 (2019), pp. 471–483.
  • [38] C. Talischi, G. H. Paulino, A. Pereira, and I. F. M. Menezes, PolyMesher: a general-purpose mesh generator for polygonal elements written in Matlab, Struct. Multidiscip. Optim., 45 (2012), pp. 309–328.
  • [39] J. Wang and X. Ye, A weak Galerkin finite element method for second-order elliptic problems, J. Comput. Appl. Math., 241 (2013), pp. 103–115.
  • [40]  , A weak Galerkin mixed finite element method for second order elliptic problems, Math. Comp., 83 (2014), pp. 2101–2126.
  • [41]  , A weak Galerkin finite element method for the stokes equations, Adv. Comput. Math., 42 (2016), pp. 155–174.
  • [42] X. Wang, Q. Zhai, R. Wang, and R. Jari, An absolutely stable weak Galerkin finite element method for the Darcy-Stokes problem, Appl. Math. Comput., 331 (2018), pp. 20–32.
  • [43] X. Ye and S. Zhang, A stabilizer free weak Galerkin finite element method on polytopal mesh: Part II, J. Comput. Appl. Math., 394 (2021), pp. Paper No. 113525, 11.
  • [44] S.-Y. Yi, A lowest-order weak Galerkin method for linear elasticity, J. Comput. Appl. Math., 350 (2019), pp. 286–298.