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

Convergence analysis of a weak Galerkin finite element method on a Shishkin mesh for a singularly perturbed fourth-order problem in 2D

Shicheng Liu [email protected] Xiangyun Meng [email protected] Qilong Zhai [email protected] School of Mathematics, Jilin University, Changchun 130012, China School of Mathematics and Statistics, Beijing Jiaotong University, Beijing 100044, China
Abstract

In this paper, we apply the weak Galerkin (WG) finite element method to solve the singularly perturbed fourth-order boundary value problem in 2D domain. A Shishkin mesh is used to ensure that the method exhibits uniform convergence, regardless of the singular perturbation parameter. Asymptotically optimal order error estimate in a H2H^{2} discrete norm is established for the corresponding WG solutions. Numerical tests are presented to verify the theory of convergence.

keywords:
Weak Galerkin finite element method, fourth-order differential equation, singularly perturbed, Shishkin mesh.
MSC:
[2020] 65N15 , 65N30 , 35B25
journal: Journal of Computational and Applied Mathematics

1 Introduction

This paper is concerned with a singularly perturbed fourth-order boundary value problem in a square region Ω\Omega. The problem is solved by the weak Galerkin (WG) method with a Shishkin mesh, as follows:

ε2Δ2uΔu\displaystyle\varepsilon^{2}\Delta^{2}u-\Delta u =f,inΩ,\displaystyle=f,\quad\text{in}~{}\Omega, (1.1)
u\displaystyle u =0,onΩ,\displaystyle=0,\quad\text{on}~{}\partial\Omega, (1.2)
u𝐧\displaystyle\frac{\partial u}{\partial\mathbf{n}} =0,onΩ,\displaystyle=0,\quad\text{on}~{}\partial\Omega, (1.3)

with a positive parameter ε\varepsilon satisfying 0<ε10<\varepsilon\ll 1 and fL2(Ω)f\in L^{2}(\Omega). This problem is used for thin elastic plates clamped in tension. ff represents the transverse load, ε\varepsilon symbolizes the ratio of bending stiffness to tensile stiffness of the plate, and the function’s solution, denoted as uu, represents the displacement of the plate. This problem emerges in the investigation of the linearization of the fourth-order perturbation associated with the fully nonlinear Monge-Ampère equation[2, 3].

A variational formulation for the fourth-order equation (1.1) with the boundary conditions (1.2) and (1.3) seeks uH2(Ω)u\in H^{2}(\Omega) such that

ε2(Δu,Δv)+(u,v)=(f,v),vH02(Ω),\varepsilon^{2}(\Delta u,\Delta v)+(\nabla u,\nabla v)=(f,v),\quad\forall v\in H^{2}_{0}(\Omega), (1.4)

where H02(Ω)H^{2}_{0}(\Omega) is a subspace of the Sobolev space H2(Ω)H^{2}(\Omega) consisting of functions with vanishing value and normal derivative on Ω\partial\Omega.

Research into numerical methods for singularly perturbed differential equations commenced in the early 1970s, with the frontier of research continuously expanding ever since. Bakhvalov made an important early contribution to the optimization of numerical methods by means of special meshes [1] in 1969. In the early 1990s, G.I. Shishkin proposed piecewise-equidistant meshes [20, 26]. These meshes, characterized by their very simple structure, are usually easy to analyze. Shishkin meshes for various problems and numerical methods have been studied since and they are still popular. Numerous scholars have explored the numerical analysis of analogous problems using various finite element methods, such as the mixed finite element method in [10], the hp finite element method in [7], the continuous interior penalty finite element method in [11], the upwind finite volume element method in [36], the conforming finite element method in [24]. And also some papers consider finite element methods on quasi-uniform meshes: a continuous interior penalty finite element method in [3] and nonconforming finite element method in [5, 13, 15, 19, 23, 31, 34, 38].

In this paper, we employ the WG method with a Shishkin mesh to investigate the convergence behavior of a fourth-order boundary value problem that exhibits singular perturbation. The WG method proves to be an effective numerical technique for the partial differential equations(PDEs). The initial proposal for its application in solving second-order elliptic problems was made by Junping Wang and Xiu Ye in [28]. The core concept involves establishing distinct basis functions for the interior and boundary of each partitioned element, and substituting the traditional differential operator with a discretized weak differential operator. The WG method has been applied to Stokes equations [29, 32, 37], elasticity equations [4, 14, 27, 39], Maxwell’s equations [22], biharmonic equations [8, 44], Navier-Stokes equations [16, 18, 43], Brinkman equations [21, 35, 40], the multigrid approach [6], the incompressible flow [45], the maximum principle [17, 30], the post-processing technique [33] and so on. For singular perturbed value problems, the WG method has also yielded some results, such as the singularly perturbed convection-diffusion problems for WG in 1D [42, 46] and 2D [41], the singularly perturbed biharmonic equation for WG in uniform mesh [9].

This paper is organized as follows. In Section 2, we introduce the Shishkin mesh and the assumptions associated. In Section 3, we give the definitions of the weak Laplacian operator and weak gradient operator. We also present WG finite element schemes for the singularly perturbed value problem. In Section 4, we introduce some local L2L^{2} projection operators and give some approximation properties. In Section 5, we establish error estimates for the WG scheme in a H2H^{2}-equivalent discrete norm. And in Section 6, we report the results of two numerical experiments.

2 Preliminaries and notations

To solve problem (1.1)-(1.3), we suppose the following assumption holds in [19], which involves structuring the solution and decomposing uu into smooth and layered components.

Assumption 2.1.

The solution uu to the singularly perturbed fourth-order boundary value problem (1.1)-(1.3) can be expressed as the sum of its smooth and layered components, as follows:

u=S+l=14El+E12+E23+E34+E41.\displaystyle u=S+\sum_{l=1}^{4}E_{l}+E_{12}+E_{23}+E_{34}+E_{41}.

In this decomposition, SS represents a smooth function, each ElE_{l} corresponds to a boundary layer component along the sides of Ω¯\bar{\Omega} in anti-clockwise order, and the remaining components are corner layer parts along the corners of Ω¯\bar{\Omega} in anti-clockwise order. Furthermore, there exists a constant CC for all points (x,y)Ω¯(x,y)\in\bar{\Omega}, which is independent of xx and yy. This constant satisfies the following conditions for 0i+jk+10\leq i+j\leq k+1,

|i+jSxiyj|C,\displaystyle\left|\frac{\partial^{i+j}S}{\partial x^{i}\partial y^{j}}\right|\leq C,
|i+jE1xiyj|Cε1jeyε,\displaystyle\left|\frac{\partial^{i+j}E_{1}}{\partial x^{i}\partial y^{j}}\right|\leq C\varepsilon^{1-j}e^{-\frac{y}{\varepsilon}},
|i+jE4xiyj|Cε1iexε,\displaystyle\left|\frac{\partial^{i+j}E_{4}}{\partial x^{i}\partial y^{j}}\right|\leq C\varepsilon^{1-i}e^{-\frac{x}{\varepsilon}},
|i+jE41xiyj|Cε1ijexεeyε.\displaystyle\left|\frac{\partial^{i+j}E_{41}}{\partial x^{i}\partial y^{j}}\right|\leq C\varepsilon^{1-i-j}e^{-\frac{x}{\varepsilon}}e^{-\frac{y}{\varepsilon}}.

Moreover, the other components of the decomposition are bounded in a similar manner.

In order to solve the layer structure in the solution of problems (1.1)-(1.3), a well-suited layer-adapted Shishkin mesh be considered. This mesh is refined in the layers. For a comprehensive discussion on the construction of Shishkin meshes, please refer to [25].

Consider a positive integer N4N\geq 4 that is divisible by 44. We introduce a mesh transition parameter λ\lambda to determine the location at which the mesh switches from coarse to fine. This parameter is defined by

λ=min{αεlnN,14},\lambda=min\left\{\alpha\varepsilon lnN,\frac{1}{4}\right\}, (2.1)

where α\alpha is a positive constant, selected to be k+1k+1 for the subsequent analysis.

Create a piecewise equidistant mesh for the interval [0,1][0,1] by dividing it as follows: divide [0,λ][0,\lambda] into N/4N/4 subintervals, [λ,1λ][\lambda,1-\lambda] into N/2N/2 subintervals, and [1λ,1][1-\lambda,1] into N/4N/4 subintervals. The Shishkin mesh for the problem (1.1)-(1.3) is formed by taking the tensor product of two such one-dimensional meshes, as illustrated in Figure 1. The fine meshwidth denoted as hh and the coarse meshwidth denoted as HH in the Shishkin mesh represented by 𝒯N\mathcal{T}_{N} exhibit the following characteristics:

h=4λNCεN1lnNandH=212λNCN1h=\frac{4\lambda}{N}\leq C\varepsilon N^{-1}lnN\quad\text{and}\quad H=2\frac{1-2\lambda}{N}\leq CN^{-1} (2.2)

for some constant CC. The domain Ω\Omega is divided into some subdomains, see Figure 1. Denote Ωl1ΩsΩl3\Omega_{l1}\cup\Omega_{s}\cup\Omega_{l3} by Ωr1\Omega_{r}^{1} and Ωl2ΩsΩl4\Omega_{l2}\cup\Omega_{s}\cup\Omega_{l4} by Ωr2\Omega_{r}^{2}.

Refer to caption
Figure 1: A rectangular Shishkin mesh with N=8N=8 and dissection of Ω\Omega.

3 Weak differential operator and WG scheme

To propose the weak Galer-kin method, we introduce some key concepts. Consider a element TT belonging to the partition 𝒯N\mathcal{T}_{N} with a boundary denoted by T\partial T. Let the set of all edges in 𝒯N\mathcal{T}_{N} be represented as N\mathcal{E}_{N}. We define a weak function v={v0,vb,𝐯g}v=\{v_{0},v_{b},\mathbf{v}_{g}\} on the element TT, where v0L2(T)v_{0}\in L^{2}(T), vbH12(T)v_{b}\in H^{\frac{1}{2}}(\partial T), and 𝐯g𝐧H12(T)\mathbf{v}_{g}\cdot\mathbf{n}\in H^{\frac{1}{2}}(\partial T), with 𝐧\mathbf{n} representing the outward normal direction on T\partial T. Furthermore, the first and second components, v0v_{0} and vbv_{b}, correspond to values of vv in the interior and on the boundary of TT. The third component 𝐯g\mathbf{v}_{g} is employed to approximate the gradient of vv along the boundary of TT. It’s important to note that each edge eNe\in\mathcal{E}_{N} has a unique value for vbv_{b} and 𝐯g\mathbf{v}_{g}. Additionally, it’s worth mentioning that vbv_{b} and 𝐯g\mathbf{v}_{g} may not necessarily be associated with the trace of v0v_{0} and v0\nabla v_{0} on T\partial T.

For any integer k3k\geq 3, we establish a local discrete weak function space for any element T𝒯NT\in\mathcal{T}_{N} denoted as Wk(T)W_{k}(T):

Wk(T)={v={v0,vb,𝐯g}:v0k(T),vbk(e),𝐯g[k(e)]2,eT},W_{k}(T)=\left\{v=\{v_{0},v_{b},\mathbf{v}_{g}\}:v_{0}\in\mathbb{Q}_{k}(T),v_{b}\in\mathbb{P}_{k}(e),\mathbf{v}_{g}\in[\mathbb{P}_{k}(e)]^{2},e\in\partial T\right\},

where ee is the edge of T\partial T, and k\mathbb{Q}_{k} is the space of polynomials which are of degree not exceeding kk with respect to each one of the variables xx and yy. By extending Wk(T)W_{k}(T) to encompass all element T𝒯NT\in\mathcal{T}_{N}, we introduce the definition of a weak Galerkin space:

VN={v={v0,vb,𝐯g}:v|TWk(T),T𝒯N}.V_{N}=\left\{v=\{v_{0},v_{b},\mathbf{v}_{g}\}:v|_{T}\in W_{k}(T),\forall T\in\mathcal{T}_{N}\right\}.

Let VN0V_{N}^{0} represent the subspace of VhV_{h} where the traces vanish:

VN0={v={v0,vb,𝐯g}Vh,vb|e=0,𝐯g𝐧|e=0,eTΩ}.V_{N}^{0}=\left\{v=\{v_{0},v_{b},\mathbf{v}_{g}\}\in V_{h},v_{b}|_{e}=0,\mathbf{v}_{g}\cdot\mathbf{n}|_{e}=0,e\subset\partial T\cap\partial\Omega\right\}.

For any v={v0,vb,𝐯g}v=\{v_{0},v_{b},\mathbf{v}_{g}\} and a fixed integer k3k\geq 3, we define a discrete weak Laplacian operator Δw,k\Delta_{w,k} as a unique polynomial Δw,kvk(T)\Delta_{w,k}v\in\mathbb{Q}_{k}(T) on TT satisfying the following equation:

(Δw,kv,φ)T=(v0,Δφ)Tvb,φ𝐧+𝐯g𝐧,φ,φk(T),\displaystyle\left(\Delta_{w,k}v,\varphi\right)_{T}=\left(v_{0},\Delta\varphi\right)_{T}-\left\langle v_{b},\nabla\varphi\cdot\mathbf{n}\right\rangle+\left\langle\mathbf{v}_{g}\cdot\mathbf{n},\varphi\right\rangle,\quad\forall\varphi\in\mathbb{Q}_{k}(T), (3.1)

where 𝐧\mathbf{n} is the unit outward normal vector to T\partial T. Likewise, a discrete weak gradient w,k\nabla_{w,k} is defined on TT as a unique polynomial w,kv[k(T)]2\nabla_{w,k}v\in[\mathbb{Q}_{k}(T)]^{2} satisfying:

(w,kv,𝐪)T=(v0,𝐪)T+vb,𝐪𝐧T,𝐪[k(T)]2.\displaystyle\left(\nabla_{w,k}v,\mathbf{q}\right)_{T}=-\left(v_{0},\nabla\cdot\mathbf{q}\right)_{T}+\left\langle v_{b},\mathbf{q}\cdot\mathbf{n}\right\rangle_{\partial T},\quad\forall\mathbf{q}\in\left[\mathbb{Q}_{k}(T)\right]^{2}. (3.2)

For simplicity, when there is no confusion, we drop the subscript kk in the notations Δw,k\Delta_{w,k} and w,k\nabla_{w,k} for the discrete weak Laplacian and the discrete weak gradient. Additionally, we introduce the following notations:

(Δwv,Δww)𝒯N:\displaystyle\left(\Delta_{w}v,\Delta_{w}w\right)_{\mathcal{T}_{N}}: =T𝒯N(Δwv,Δww)T,\displaystyle=\sum_{T\in\mathcal{T}_{N}}\left(\Delta_{w}v,\Delta_{w}w\right)_{T},
(wv,ww)𝒯N:\displaystyle\left(\nabla_{w}v,\nabla_{w}w\right)_{\mathcal{T}_{N}}: =T𝒯N(wv,ww)T.\displaystyle=\sum_{T\in\mathcal{T}_{N}}\left(\nabla_{w}v,\nabla_{w}w\right)_{T}.

Let us introduce a stabilizer, which is a bilinear form for any uN={u0,ub,𝐮g}u_{N}=\{u_{0},u_{b},\mathbf{u}_{g}\} and v={v0,vb,𝐯g}v=\{v_{0},v_{b},\mathbf{v}_{g}\} in the space VhV_{h}. It is defined as follows:

s(uh,v)=T𝒯N(ε2h1u0𝐮g,v0𝐯gT+(ε2h2H1+H1)u0ub,v0vbT),\displaystyle\begin{aligned} s\left(u_{h},v\right)=\sum_{T\in\mathcal{T}_{N}}\Big{(}\varepsilon^{2}h^{-1}&\left\langle\nabla u_{0}-\mathbf{u}_{g},\nabla v_{0}-\mathbf{v}_{g}\right\rangle_{\partial T}\\ &+\left(\varepsilon^{2}h^{-2}H^{-1}+H^{-1}\right)\left\langle u_{0}-u_{b},v_{0}-v_{b}\right\rangle_{\partial T}\Big{)},\end{aligned} (3.3)

where hh and HH is defined as (2.2).

Algorithm 1 WG Algorithm

To obtain a numerical approximation for (1.1)-(1.3), we seek uN={u0,ub,𝐮g}VNu_{N}=\{u_{0},u_{b},\mathbf{u}_{g}\}\in V_{N} that satisfies the following equation:

ε2(ΔwuN,Δwv)N+(wuN,wv)N+s(uN,v)=(f,v),\displaystyle\varepsilon^{2}(\Delta_{w}u_{N},\Delta_{w}v)_{N}+(\nabla_{w}u_{N},\nabla_{w}v)_{N}+s(u_{N},v)=(f,v), (3.4)

for any v={v0,vb,𝐯g}VN0v=\{v_{0},v_{b},\mathbf{v}_{g}\}\in V_{N}^{0}.

The following lemma provides a valuable result regarding the finite element space VN0V^{0}_{N}.

Lemma 3.1.

For any vVN0v\in V^{0}_{N}, let |v|{|||}v{|||} be given as follows:

|v|2=ε2(Δwv,Δwv)𝒯N+(wv,wv)𝒯N+s(v,v),\displaystyle{|||}v{|||}^{2}=\varepsilon^{2}(\Delta_{w}v,\Delta_{w}v)_{\mathcal{T}_{N}}+(\nabla_{w}v,\nabla_{w}v)_{\mathcal{T}_{N}}+s(v,v), (3.5)

Then, ||||||{|||}\cdot{|||} defines a norm in VN0V^{0}_{N}.

Proof.

We shall only confirm the positivity property for ||||||{|||}\cdot{|||}. Consider v={v0,vb,𝐯g}VN0v=\{v_{0},v_{b},\mathbf{v}_{g}\}\in V^{0}_{N} with the assumption that |v|=0{|||}v{|||}=0. This implies Δwv=0\Delta_{w}v=0 and wv=0\nabla_{w}v=0 in each element TT, while also satisfying v0=vbv_{0}=v_{b} and v0=𝐯g\nabla v_{0}=\mathbf{v}_{g} on T\partial T. Next, we will proof that Δv0=0\Delta v_{0}=0 in each element TT. To this end, for any φk(T)\varphi\in\mathbb{Q}_{k}(T), employing the definition (3.1) and the fact that Δwv=0\Delta_{w}v=0, we obtain

0=(Δwv,φ)T=(v0,Δφ)Tvb,φ𝐧T+𝐯g𝐧,φT=(Δv0,φ)T+v0vb,φ𝐧T+𝐯g𝐧v0𝐧,φT=(Δv0,φ)T,\begin{split}0&=\left(\Delta_{w}v,\varphi\right)_{T}\\ &=\left(v_{0},\Delta\varphi\right)_{T}-\left\langle v_{b},\nabla\varphi\cdot\mathbf{n}\right\rangle_{\partial T}+\left\langle\mathbf{v}_{g}\cdot\mathbf{n},\varphi\right\rangle_{\partial T}\\ &=\left(\Delta v_{0},\varphi\right)_{T}+\left\langle v_{0}-v_{b},\nabla\varphi\cdot\mathbf{n}\right\rangle_{\partial T}+\left\langle\mathbf{v}_{g}\cdot\mathbf{n}-\nabla v_{0}\cdot\mathbf{n},\varphi\right\rangle_{\partial T}\\ &=\left(\Delta v_{0},\varphi\right)_{T},\end{split} (3.6)

where we have applied the fact that v0vb=0v_{0}-v_{b}=0 and v0𝐯g=0\nabla v_{0}-\mathbf{v}_{g}=0 in the final equality. The identity (3.6) indicates that Δv0=0\Delta v_{0}=0 in each element TT. Together with the conditions v0=vbv_{0}=v_{b} and v0=𝐯g\nabla v_{0}=\mathbf{v}_{g} on T\partial T, we conclude that vv is a globally smooth harmonic function on Ω\Omega. Considering the boundary conditions vb=0v_{b}=0 and 𝐯g𝐧=0\mathbf{v}_{g}\cdot\mathbf{n}=0, we infer that the unique solution is v0v\equiv 0 on Ω\Omega. This concludes the proof. ∎

4 Local L2L^{2} projection operators and approximation properties

In this section, we introduce some projection operators for each element T𝒯NT\in\mathcal{T}_{N}. Consider the L2L^{2} projection operator 𝒬0\mathcal{Q}_{0}, which projects onto k(T)\mathbb{Q}_{k}(T). Additionally, for each edge eTe\in\partial T, we consider the L2L^{2} projection operators 𝒬b\mathcal{Q}_{b} and 𝐐g\mathbf{Q}_{g} onto local polynomial spaces k(e)\mathbb{P}_{k}(e) and [k(e)]2[\mathbb{P}_{k}(e)]^{2}, respectively. We define a projection 𝒬N\mathcal{Q}_{N} of uu into the finite element space VNV_{N} such that on each element TT

𝒬Nu={𝒬0u,𝒬bu,𝐐g(u)}.\displaystyle\mathcal{Q}_{N}u=\left\{\mathcal{Q}_{0}u,\mathcal{Q}_{b}u,\mathbf{Q}_{g}(\nabla u)\right\}.

Furthermore, let 𝐐N\mathbf{Q}_{N} represent the local L2L^{2} projection onto [k(T)]2[\mathbb{Q}_{k}(T)]^{2}. The following lemma demonstrates that the weak Laplacian Δw\Delta_{w} is the polynomial projection of the classical Laplacian Δ\Delta.

Lemma 4.1.

On each element T𝒯NT\in\mathcal{T}_{N}, for any vH2(T)v\in H^{2}(T),

Δw(𝒬Nv)=𝒬0(Δv).\displaystyle\Delta_{w}(\mathcal{Q}_{N}v)=\mathcal{Q}_{0}(\Delta v). (4.1)
Proof.

For any τk(T)\tau\in\mathbb{Q}_{k}(T), we obtain that

(Δw𝒬Nu,τ)T\displaystyle\left(\Delta_{w}\mathcal{Q}_{N}u,\tau\right)_{T} =(𝒬0u,Δτ)T𝒬bu,τ𝐧T+𝐐g(u)𝐧,τT\displaystyle=\left(\mathcal{Q}_{0}u,\Delta\tau\right)_{T}-\left\langle\mathcal{Q}_{b}u,\nabla\tau\cdot\mathbf{n}\right\rangle_{\partial T}+\left\langle\mathbf{Q}_{g}(\nabla u)\cdot\mathbf{n},\tau\right\rangle_{\partial T}
=(u,Δτ)Tu,τ𝐧T+u𝐧,τT\displaystyle=\left(u,\Delta\tau\right)_{T}-\left\langle u,\nabla\tau\cdot\mathbf{n}\right\rangle_{\partial T}+\left\langle\nabla u\cdot\mathbf{n},\tau\right\rangle_{\partial T}
=(Δu,τ)T\displaystyle=\left(\Delta u,\tau\right)_{T}
=(𝒬0Δu,τ)T.\displaystyle=\left(\mathcal{Q}_{0}\Delta u,\tau\right)_{T}.

This concludes the proof. ∎

A similar lemma holds for the weak gradient w\nabla_{w}, as indicated in the subsequent lemma.

Lemma 4.2.

On each element T𝒯NT\in\mathcal{T}_{N}, for any vH2(T)v\in H^{2}(T),

w(𝒬Nu)=𝐐N(u).\displaystyle\nabla_{w}(\mathcal{Q}_{N}u)=\mathbf{Q}_{N}(\nabla u). (4.2)
Proof.

For any 𝐪[k(T)]2\mathbf{q}\in[\mathbb{Q}_{k}(T)]^{2}, from the weak gradient definition and integration by parts, we get

(w𝒬Nu,𝐪)T\displaystyle\left(\nabla_{w}\mathcal{Q}_{N}u,\mathbf{q}\right)_{T} =(𝒬0u,𝐪)T+𝒬bu,𝐪𝐧T\displaystyle=-\left(\mathcal{Q}_{0}u,\nabla\cdot\mathbf{q}\right)_{T}+\left\langle\mathcal{Q}_{b}u,\mathbf{q}\cdot\mathbf{n}\right\rangle_{\partial T}
=(u,𝐪)T+u,𝐪𝐧T\displaystyle=-\left(u,\nabla\cdot\mathbf{q}\right)_{T}+\left\langle u,\mathbf{q}\cdot\mathbf{n}\right\rangle_{\partial T}
=(u,𝐪)T\displaystyle=\left(\nabla u,\mathbf{q}\right)_{T}
=(𝐐Nu,𝐪)T.\displaystyle=\left(\mathbf{Q}_{N}\nabla u,\mathbf{q}\right)_{T}.

This corresponds to identity (4.2). ∎

Now, we introduce notation that will be employed in the following lemmas. Consider any element TT in the partition 𝒯N\mathcal{T}_{N}. We define T1\partial T_{1} and T2\partial T_{2} as the sets of element edges that are parallel to the xx and yy axes, respectively. The following lemmas are employed in the convergence analysis, and readers are directed to [12, chapter 3.1] for a detailed proof process.

Lemma 4.3.

Consider ϕHk+1(T)\phi\in H^{k+1}(T) with k3k\geq 3. Let 𝒬0ϕ\mathcal{Q}_{0}\phi denote the L2L^{2}-projection of ϕ\phi onto k(T)\mathbb{Q}_{k}(T). Then the following inequality estimate holds,

ϕ𝒬0ϕTiC(hjk+12jk+1ϕT+hik+1hj12ik+1ϕT+hj12hikikjϕT),\displaystyle\|\phi-\mathcal{Q}_{0}\phi\|_{\partial T_{i}}\leq C\left(h_{j}^{k+\frac{1}{2}}\|\partial_{j}^{k+1}\phi\|_{T}+h_{i}^{k+1}h_{j}^{-\frac{1}{2}}\|\partial_{i}^{k+1}\phi\|_{T}+h_{j}^{\frac{1}{2}}h_{i}^{k}\|\partial_{i}^{k}\partial_{j}\phi\|_{T}\right), (4.3)
i(ϕ𝒬0ϕ)TiC(hikhj12ik+1ϕT+hik1hj12ikjϕT+hjk12jkiϕT),\displaystyle\|\partial_{i}(\phi-\mathcal{Q}_{0}\phi)\|_{\partial T_{i}}\leq C\left(h_{i}^{k}h_{j}^{-\frac{1}{2}}\|\partial_{i}^{k+1}\phi\|_{T}+h_{i}^{k-1}h_{j}^{\frac{1}{2}}\|\partial_{i}^{k}\partial_{j}\phi\|_{T}+h_{j}^{k-\frac{1}{2}}\|\partial_{j}^{k}\partial_{i}\phi\|_{T}\right), (4.4)
j(ϕ𝒬0ϕ)TiC(hjk12jk+1ϕT+hikhj12ikjϕT),\displaystyle\|\partial_{j}(\phi-\mathcal{Q}_{0}\phi)\|_{\partial T_{i}}\leq C\left(h_{j}^{k-\frac{1}{2}}\|\partial_{j}^{k+1}\phi\|_{T}+h_{i}^{k}h_{j}^{-\frac{1}{2}}\|\partial_{i}^{k}\partial_{j}\phi\|_{T}\right), (4.5)

where i,j{1,2}i,j\in\{1,2\}, and iji\neq j.

Lemma 4.4.

Let vk(T)v\in\mathbb{Q}_{k}(T) with k3k\geq 3 such that the following inequalities holds,

vTi\displaystyle\|v\|_{\partial T_{i}}\leq Chj12vT,\displaystyle Ch_{j}^{-\frac{1}{2}}\|v\|_{T}, (4.6)
ivT\displaystyle\|\partial_{i}v\|_{T}\leq Chi1vT,\displaystyle Ch_{i}^{-1}\|v\|_{T}, (4.7)

where CC is a constant only depends on kk and i{1,2}i\in\{1,2\}, iji\neq j.

By applying Lemma 4.3 and Lemma 4.4, we can deduce the following estimates which are valuable for the convergence analysis of the WG finite element schemes (3.4).

Lemma 4.5.

Let k3k\geq 3, uHk+1(Ω)u\in H^{k+1}(\Omega). There exists a constant CC such that the following estimates hold true,

i=1,2(T𝒯NΔu𝒬0ΔuTi2)12C(N(k32)+ε1N(k32)lnk32N+ε1N12α),\displaystyle\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\Delta u-\mathcal{Q}_{0}\Delta u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq C\left(N^{-(k-\frac{3}{2})}+\varepsilon^{-1}N^{-(k-\frac{3}{2})}\ln^{k-\frac{3}{2}}N+\varepsilon^{-1}N^{\frac{1}{2}-\alpha}\right), (4.8)
i=1,2(T𝒯N(Δu𝒬0Δu)Ti2)12C(N(k52)+ε2N(k52)lnk52N+ε2N32α),\displaystyle\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\nabla\left(\Delta u-\mathcal{Q}_{0}\Delta u\right)\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq C\left(N^{-(k-\frac{5}{2})}+\varepsilon^{-2}N^{-(k-\frac{5}{2})}\ln^{k-\frac{5}{2}}N+\varepsilon^{-2}N^{\frac{3}{2}-\alpha}\right), (4.9)
i=1,2(T𝒯Nu𝐐NuTi2)12C(N(k12)+N(k12)lnk12N+N12α),\displaystyle\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\nabla u-\mathbf{Q}_{N}\nabla u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq C\left(N^{-(k-\frac{1}{2})}+N^{-(k-\frac{1}{2})}\ln^{k-\frac{1}{2}}N+N^{\frac{1}{2}-\alpha}\right), (4.10)
i=1,2(T𝒯N(𝒬0uu)Ti2)12C(N(k12)+N(k12)lnk12N+N32α),\displaystyle\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\nabla(\mathcal{Q}_{0}u-u)\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq C\left(N^{-(k-\frac{1}{2})}+N^{-(k-\frac{1}{2})}\ln^{k-\frac{1}{2}}N+N^{\frac{3}{2}-\alpha}\right), (4.11)
i=1,2(T𝒯N𝒬0uuTi2)12C(N(k+12)+εN(k+12)lnk+12N+εN12α).\displaystyle\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\mathcal{Q}_{0}u-u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq C\left(N^{-(k+\frac{1}{2})}+\varepsilon N^{-(k+\frac{1}{2})}\ln^{k+\frac{1}{2}}N+\varepsilon N^{\frac{1}{2}-\alpha}\right). (4.12)
Proof.

To derive (4.8), we estimate T𝒯NΔu𝒬0ΔuTi\sum_{T\in\mathcal{T}{N}}\|\Delta u-\mathcal{Q}_{0}\Delta u\|_{\partial T_{i}} by breaking down the function uu in Assumption 2.1. Each term in the decomposition will be considered individually. To begin, we can apply inequality (4.3) to obtain

(TΩcΔS𝒬0ΔSTi2)12\displaystyle\left(\sum_{T\in\Omega_{c}}\|\Delta S-\mathcal{Q}_{0}\Delta S\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq Chk32NhCN(k32),\displaystyle Ch^{k-\frac{3}{2}}\cdot N\cdot h\leq CN^{-(k-\frac{3}{2})},
(TΩsΔS𝒬0ΔSTi2)12\displaystyle\left(\sum_{T\in\Omega_{s}}\|\Delta S-\mathcal{Q}_{0}\Delta S\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq CHk32NHCN(k32),\displaystyle CH^{k-\frac{3}{2}}\cdot N\cdot H\leq CN^{-(k-\frac{3}{2})},
(TΩlΔS𝒬0ΔSTi2)12\displaystyle\left(\sum_{T\in\Omega_{l}}\|\Delta S-\mathcal{Q}_{0}\Delta S\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq CNh112h212(hjk32+hj12hik1+hj12hik2)\displaystyle CN\cdot h_{1}^{\frac{1}{2}}h_{2}^{\frac{1}{2}}\left(h_{j}^{k-\frac{3}{2}}+h_{j}^{-\frac{1}{2}}h_{i}^{k-1}+h_{j}^{\frac{1}{2}}h_{i}^{k-2}\right)
\displaystyle\leq CN(k32),\displaystyle CN^{-(k-\frac{3}{2})},

where we have used the fact that hi=hh_{i}=h in the domain Ωc\Omega_{c} and hi=Hh_{i}=H in the domain Ωs\Omega_{s}, for i=1,2i=1,2. Next, we provide estimates only for the sets of element edges parallel to the xx axis, as the orther part follows a similar way. Considering the boundary layer E1E_{1}, we obtain

(TΩcΔE1𝒬0ΔE1T12)12\displaystyle\left(\sum_{T\in\Omega_{c}}\|\Delta E_{1}-\mathcal{Q}_{0}\Delta E_{1}\|_{\partial T_{1}}^{2}\right)^{\frac{1}{2}}\leq Chk32εk(Ωce2yε𝑑x𝑑y)12\displaystyle Ch^{k-\frac{3}{2}}\cdot\varepsilon^{-k}\left(\int_{\Omega_{c}}e^{-\frac{2y}{\varepsilon}}\,dxdy\right)^{\frac{1}{2}}
\displaystyle\leq Cεkhk32εln12N\displaystyle C\varepsilon^{-k}\cdot h^{k-\frac{3}{2}}\cdot\varepsilon\ln^{\frac{1}{2}}N
\displaystyle\leq Cε12N(k32)lnk1N,\displaystyle C\varepsilon^{-\frac{1}{2}}N^{-(k-\frac{3}{2})}\ln^{k-1}N,
(TΩlΔE1𝒬0ΔE1T12)12\displaystyle\left(\sum_{T\in\Omega_{l}}\|\Delta E_{1}-\mathcal{Q}_{0}\Delta E_{1}\|_{\partial T_{1}}^{2}\right)^{\frac{1}{2}}\leq C(ε2kh2k32+εh1k1h212+h1k2h212+εkh2k32\displaystyle C\left(\right.\varepsilon^{2-k}h_{2}^{k-\frac{3}{2}}+\varepsilon h_{1}^{k-1}h_{2}^{-\frac{1}{2}}+h_{1}^{k-2}h_{2}^{\frac{1}{2}}+\varepsilon^{-k}h_{2}^{k-\frac{3}{2}}
+ε1h1k1h212+ε2h1k2h212)(Ωle2yεdxdy)12\displaystyle+\varepsilon^{-1}h_{1}^{k-1}h_{2}^{-\frac{1}{2}}+\varepsilon^{-2}h_{1}^{k-2}h_{2}^{\frac{1}{2}}\left.\right)\left(\int_{\Omega_{l}}e^{-\frac{2y}{\varepsilon}}\,dxdy\right)^{\frac{1}{2}}
\displaystyle\leq Cε32N(k32)lnk32Nε12\displaystyle C\varepsilon^{-\frac{3}{2}}N^{-(k-\frac{3}{2})}\ln^{k-\frac{3}{2}}N\cdot\varepsilon^{\frac{1}{2}}
\displaystyle\leq Cε1N(k32)lnk32N,\displaystyle C\varepsilon^{-1}N^{-(k-\frac{3}{2})}\ln^{k-\frac{3}{2}}N,

where we have used the inequality (4.3). As for the region Ωr2\Omega_{r}^{2} that remains in the partition, by applying inequality (4.6), we derive

(TΩr2ΔE1𝒬0ΔE1T12)12\displaystyle\left(\sum_{T\in\Omega_{r}^{2}}\|\Delta E_{1}-\mathcal{Q}_{0}\Delta E_{1}\|_{\partial T_{1}}^{2}\right)^{\frac{1}{2}}\leq CTΩr2ΔE1T1+TΩr2h212𝒬0ΔE1T\displaystyle C\sum_{T\in\Omega_{r}^{2}}\|\Delta E_{1}\|_{\partial T_{1}}+\sum_{T\in\Omega_{r}^{2}}h_{2}^{-\frac{1}{2}}\|\mathcal{Q}_{0}\Delta E_{1}\|_{T}
\displaystyle\leq C(ε+ε1)(N01e2yε𝑑x)12+H12ΔE1Ωr2\displaystyle C(\varepsilon+\varepsilon^{-1})\left(N\int_{0}^{1}e^{-\frac{2y}{\varepsilon}}\,dx\right)^{\frac{1}{2}}+H^{-\frac{1}{2}}\left\|\Delta E_{1}\right\|_{\Omega_{r}^{2}}
\displaystyle\leq C(ε1N12α+H12(ε32Nα+ε12Nα))\displaystyle C\left(\varepsilon^{-1}N^{\frac{1}{2}-\alpha}+H^{-\frac{1}{2}}(\varepsilon^{\frac{3}{2}}N^{-\alpha}+\varepsilon^{-\frac{1}{2}}N^{-\alpha})\right)
\displaystyle\leq Cε1N12α.\displaystyle C\varepsilon^{-1}N^{\frac{1}{2}-\alpha}.

A similar bound can be readily obtained for E2,E3E_{2},E_{3}, and E4E_{4}. Let’s focus on estimating E41E_{41} for the concer layers, as the other concer layers in the decomposition from Assumption 2.1 follow a similar way. By applying inequalities (4.3) and (4.6), we arrive at

(TΩcΔE41𝒬0ΔE41Ti2)12\displaystyle\left(\sum_{T\in\Omega_{c}}\|\Delta E_{41}-\mathcal{Q}_{0}\Delta E_{41}\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\leq Cεkhk32(Ωce2xεe2yε𝑑x𝑑y)12\displaystyle C\varepsilon^{-k}h^{k-\frac{3}{2}}\cdot\left(\int_{\Omega_{c}}e^{-\frac{2x}{\varepsilon}}e^{-\frac{2y}{\varepsilon}}\,dxdy\right)^{\frac{1}{2}}
\displaystyle\leq Cε12N(k32)lnk32N.\displaystyle C\varepsilon^{-\frac{1}{2}}N^{-(k-\frac{3}{2})}\ln^{k-\frac{3}{2}}N.
(TΩr1ΔE41𝒬0ΔE41T12)12\displaystyle\left(\sum_{T\in\Omega_{r}^{1}}\|\Delta E_{41}-\mathcal{Q}_{0}\Delta E_{41}\|_{\partial T_{1}}^{2}\right)^{\frac{1}{2}}\leq CTΩr1(ΔE41T1+h212𝒬0ΔE41T)\displaystyle C\sum_{T\in\Omega_{r}^{1}}\left(\left\|\Delta E_{41}\right\|_{\partial T_{1}}+h_{2}^{-\frac{1}{2}}\left\|\mathcal{Q}_{0}\Delta E_{41}\right\|_{T}\right)
\displaystyle\leq Cε1(Nλ1λe2xεe2yε𝑑x)12+h12ΔE41Ωr1\displaystyle C\varepsilon^{-1}\left(N\int_{\lambda}^{1-\lambda}e^{-\frac{2x}{\varepsilon}}e^{-\frac{2y}{\varepsilon}}\,dx\right)^{\frac{1}{2}}+h^{-\frac{1}{2}}\left\|\Delta E_{41}\right\|_{\Omega_{r}^{1}}
\displaystyle\leq C(ε1ε12N12α+h12Nα)\displaystyle C\left(\varepsilon^{-1}\cdot\varepsilon^{\frac{1}{2}}N^{\frac{1}{2}-\alpha}+h^{-\frac{1}{2}}\cdot N^{-\alpha}\right)
\displaystyle\leq Cε12N12α,\displaystyle C\varepsilon^{-\frac{1}{2}}N^{\frac{1}{2}-\alpha},
(TΩr2ΔE41𝒬0ΔE41T12)12\displaystyle\left(\sum_{T\in\Omega_{r}^{2}}\|\Delta E_{41}-\mathcal{Q}_{0}\Delta E_{41}\|_{\partial T_{1}}^{2}\right)^{\frac{1}{2}}\leq CTΩr2(ΔE41T1+h212𝒬0ΔE41T)\displaystyle C\sum_{T\in\Omega_{r}^{2}}\left(\left\|\Delta E_{41}\right\|_{\partial T_{1}}+h_{2}^{-\frac{1}{2}}\left\|\mathcal{Q}_{0}\Delta E_{41}\right\|_{T}\right)
\displaystyle\leq Cε1(N01e2xεe2yε𝑑x)12+H12ΔE41Ωr2\displaystyle C\varepsilon^{-1}\left(N\int_{0}^{1}e^{-\frac{2x}{\varepsilon}}e^{-\frac{2y}{\varepsilon}}\,dx\right)^{\frac{1}{2}}+H^{-\frac{1}{2}}\left\|\Delta E_{41}\right\|_{\Omega_{r}^{2}}
\displaystyle\leq C(ε1ε12N12α+H12Nα)\displaystyle C\left(\varepsilon^{-1}\cdot\varepsilon^{\frac{1}{2}}N^{\frac{1}{2}-\alpha}+H^{-\frac{1}{2}}\cdot N^{-\alpha}\right)
\displaystyle\leq Cε12N12α.\displaystyle C\varepsilon^{-\frac{1}{2}}N^{\frac{1}{2}-\alpha}.

By combining the aforementioned proofs, we establish inequality (4.8). Likewise, for inequalities (4.9), (4.10), (4.11), and (4.12), the proof follows a similar way as above. Therefore, we omit the detailed explanation. ∎

5 Error estimate

In this section, the objective is to provide error estimates for the WG solution uNu_{N} obtained from (3.4).

5.1 Error equation

We introduce notation used in error analysis to represent the error between the finite element solution and the L2L^{2} projection of the exact solution, as follows

eN=𝒬NuuN={e0,eb,𝐞g}.\displaystyle e_{N}=\mathcal{Q}_{N}u-u_{N}=\{e_{0},e_{b},\mathbf{e}_{g}\}.

The convergence analysis relies on the error equation, and in the following lemma, we will establish an equation that the error eNe_{N} satisfies.

Lemma 5.1.

Suppose uu and uN={u0,ub,𝐮g}VNu_{N}=\{u_{0},u_{b},\mathbf{u}_{g}\}\in V_{N} represent the solutions of (1.1) and (3.4), respectively. Then for any vVN0v\in V_{N}^{0}, we have

a(eN,v)=l1(u,v)l2(u,v)+l3(u,v)+s(𝒬Nu,v),\displaystyle a(e_{N},v)=l_{1}(u,v)-l_{2}(u,v)+l_{3}(u,v)+s(\mathcal{Q}_{N}u,v), (5.1)

where

l1(u,v)=T𝒯Nε2Δu𝒬0Δu,(v0𝐯g)𝐧T,\displaystyle l_{1}(u,v)=\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}\left\langle\Delta u-\mathcal{Q}_{0}\Delta u,\left(\nabla v_{0}-\mathbf{v}_{g}\right)\cdot\mathbf{n}\right\rangle_{\partial T},
l2(u,v)=T𝒯Nε2(Δu𝒬0Δu)𝐧,v0vbT,\displaystyle l_{2}(u,v)=\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}\left\langle\nabla\left(\Delta u-\mathcal{Q}_{0}\Delta u\right)\cdot\mathbf{n},v_{0}-v_{b}\right\rangle_{\partial T},
l3(u,v)=T𝒯N(u𝐐Nu)𝐧,v0vbT.\displaystyle l_{3}(u,v)=\sum_{T\in\mathcal{T}_{N}}\left\langle\left(\nabla u-\mathbf{Q}_{N}\nabla u\right)\cdot\mathbf{n},v_{0}-v_{b}\right\rangle_{\partial T}.
Proof.

Using the definition of weak Laplacian (3.2), integration by parts and Lemma 4.1, for any vVN0v\in V_{N}^{0}, we yield

(Δw𝒬Nu,Δwv)T\displaystyle\left(\Delta_{w}\mathcal{Q}_{N}u,\Delta_{w}v\right)_{T} =(v0,Δ(Δw𝒬Nu))T+𝐯g𝐧,Δw𝒬NuTvb,(Δw𝒬Nu)𝐧T\displaystyle=\left(v_{0},\Delta\left(\Delta_{w}\mathcal{Q}_{N}u\right)\right)_{T}+\left\langle\mathbf{v}_{g}\cdot\mathbf{n},\Delta_{w}\mathcal{Q}_{N}u\right\rangle_{\partial T}-\left\langle v_{b},\nabla\left(\Delta_{w}\mathcal{Q}_{N}u\right)\cdot\mathbf{n}\right\rangle_{\partial T}
=(Δv0,Δw𝒬Nu)T+v0,(Δw𝒬Nu)𝐧Tv0𝐧,Δw𝒬NuT\displaystyle=\left(\Delta v_{0},\Delta_{w}\mathcal{Q}_{N}u\right)_{T}+\left\langle v_{0},\nabla\left(\Delta_{w}\mathcal{Q}_{N}u\right)\cdot\mathbf{n}\right\rangle_{\partial T}-\left\langle\nabla v_{0}\cdot\mathbf{n},\Delta_{w}\mathcal{Q}_{N}u\right\rangle_{\partial T}
+𝐯g𝐧,Δw𝒬NuTvb,(Δw𝒬Nu)𝐧T\displaystyle\quad+\left\langle\mathbf{v}_{g}\cdot\mathbf{n},\Delta_{w}\mathcal{Q}_{N}u\right\rangle_{\partial T}-\left\langle v_{b},\nabla\left(\Delta_{w}\mathcal{Q}_{N}u\right)\cdot\mathbf{n}\right\rangle_{\partial T}
=(Δv0,Δw𝒬Nu)T+v0vb,(Δw𝒬Nu)𝐧T(v0𝐯g)𝐧,Δw𝒬NuT\displaystyle=\left(\Delta v_{0},\Delta_{w}\mathcal{Q}_{N}u\right)_{T}+\left\langle v_{0}-v_{b},\nabla\left(\Delta_{w}\mathcal{Q}_{N}u\right)\cdot\mathbf{n}\right\rangle_{\partial T}-\left\langle\left(\nabla v_{0}-\mathbf{v}_{g}\right)\cdot\mathbf{n},\Delta_{w}\mathcal{Q}_{N}u\right\rangle_{\partial T}
=(Δv0,𝒬0Δu)T+v0vb,(𝒬0Δu)𝐧T(v0𝐯g)𝐧,𝒬0ΔuT\displaystyle=\left(\Delta v_{0},\mathcal{Q}_{0}\Delta u\right)_{T}+\left\langle v_{0}-v_{b},\nabla\left(\mathcal{Q}_{0}\Delta u\right)\cdot\mathbf{n}\right\rangle_{\partial T}-\left\langle\left(\nabla v_{0}-\mathbf{v}_{g}\right)\cdot\mathbf{n},\mathcal{Q}_{0}\Delta u\right\rangle_{\partial T}
=(Δu,Δv0)T+v0vb,(𝒬0Δu)𝐧T(v0𝐯g)𝐧,𝒬0ΔuT,\displaystyle=\left(\Delta u,\Delta v_{0}\right)_{T}+\left\langle v_{0}-v_{b},\nabla\left(\mathcal{Q}_{0}\Delta u\right)\cdot\mathbf{n}\right\rangle_{\partial T}-\left\langle\left(\nabla v_{0}-\mathbf{v}_{g}\right)\cdot\mathbf{n},\mathcal{Q}_{0}\Delta u\right\rangle_{\partial T},

which implies that

T𝒯h(Δu,Δv0)T=(Δw𝒬Nu,Δwv)NT𝒯hv0vb,(𝒬0Δu)𝐧T+T𝒯h(v0𝐯g)𝐧,𝒬0ΔuT.\begin{split}\sum_{T\in\mathcal{T}_{h}}\left(\Delta u,\Delta v_{0}\right)_{T}=&\left(\Delta_{w}\mathcal{Q}_{N}u,\Delta_{w}v\right)_{N}-\sum_{T\in\mathcal{T}_{h}}\left\langle v_{0}-v_{b},\nabla\left(\mathcal{Q}_{0}\Delta u\right)\cdot\mathbf{n}\right\rangle_{\partial T}\\ &+\sum_{T\in\mathcal{T}_{h}}\left\langle\left(\nabla v_{0}-\mathbf{v}_{g}\right)\cdot\mathbf{n},\mathcal{Q}_{0}\Delta u\right\rangle_{\partial T}.\end{split} (5.2)

Likewise, we deduce from integration by parts and Lemma 4.2 the following

(w𝒬Nu,wv)T\displaystyle\left(\nabla_{w}\mathcal{Q}_{N}u,\nabla_{w}v\right)_{T} =(v0,w𝒬Nu)T+vb,w𝒬Nu𝐧T\displaystyle=-\left(v_{0},\nabla\cdot\nabla_{w}\mathcal{Q}_{N}u\right)_{T}+\left\langle v_{b},\nabla_{w}\mathcal{Q}_{N}u\cdot\mathbf{n}\right\rangle_{\partial T}
=(v0,w𝒬Nu)T+vbv0,w𝒬Nu𝐧T\displaystyle=\left(\nabla v_{0},\nabla_{w}\mathcal{Q}_{N}u\right)_{T}+\left\langle v_{b}-v_{0},\nabla_{w}\mathcal{Q}_{N}u\cdot\mathbf{n}\right\rangle_{\partial T}
=(v0,𝐐Nu)T+vbv0,𝐐Nu𝐧T\displaystyle=\left(\nabla v_{0},\mathbf{Q}_{N}\nabla u\right)_{T}+\left\langle v_{b}-v_{0},\mathbf{Q}_{N}\nabla u\cdot\mathbf{n}\right\rangle_{\partial T}
=(v0,u)Tv0vb,(𝐐Nu)𝐧T,\displaystyle=\left(\nabla v_{0},\nabla u\right)_{T}-\left\langle v_{0}-v_{b},\left(\mathbf{Q}_{N}\nabla u\right)\cdot\mathbf{n}\right\rangle_{\partial T},

which implies that

T𝒯h(u,v0)T=(w𝒬Nu,wv)N+T𝒯hv0vb,𝐐Nu𝐧T.\begin{split}\sum_{T\in\mathcal{T}_{h}}\left(\nabla u,\nabla v_{0}\right)_{T}=\left(\nabla_{w}\mathcal{Q}_{N}u,\nabla_{w}v\right)_{N}+\sum_{T\in\mathcal{T}_{h}}\left\langle v_{0}-v_{b},\mathbf{Q}_{N}\nabla u\cdot\mathbf{n}\right\rangle_{\partial T}.\end{split} (5.3)

Test equation (1.1) with the vector v0v_{0} of v={v0,vb,𝐯g}VN0v=\{v_{0},v_{b},\mathbf{v}g\}\in V_{N}^{0}, we find

ε2(Δ2u,v0)(Δu,v0)=(f,v0).\displaystyle\varepsilon^{2}\left(\Delta^{2}u,v_{0}\right)-\left(\Delta u,v_{0}\right)=\left(f,v_{0}\right).

Using the boundary conditions that 𝐯g𝐧\mathbf{v}_{g}\cdot\mathbf{n} and vbv_{b} vanish on Ω\partial\Omega, along with integration by parts, we derive

ε2(Δ2u,v0)(Δu,v0)=\displaystyle\varepsilon^{2}\left(\Delta^{2}u,v_{0}\right)-\left(\Delta u,v_{0}\right)= T𝒯h[ε2(Δu,Δv0)Tε2Δu,v0𝐧T+(u,v0)T\displaystyle\sum_{T\in\mathcal{T}_{h}}\Big{[}\varepsilon^{2}\left(\Delta u,\Delta v_{0}\right)_{T}-\varepsilon^{2}\left\langle\Delta u,\nabla v_{0}\cdot\mathbf{n}\right\rangle_{\partial T}+\left(\nabla u,\nabla v_{0}\right)_{T}
+ε2(Δu)𝐧,v0Tu𝐧,v0T]\displaystyle+\varepsilon^{2}\left\langle\nabla(\Delta u)\cdot\mathbf{n},v_{0}\right\rangle_{\partial T}-\left\langle\nabla u\cdot\mathbf{n},v_{0}\right\rangle_{\partial T}\Big{]}
=\displaystyle= T𝒯h[ε2(Δu,Δv0)Tε2Δu,(v0𝐯g)𝐧T+(u,v0)T\displaystyle\sum_{T\in\mathcal{T}_{h}}\Big{[}\varepsilon^{2}\left(\Delta u,\Delta v_{0}\right)_{T}-\varepsilon^{2}\left\langle\Delta u,\left(\nabla v_{0}-\mathbf{v}_{g}\right)\cdot\mathbf{n}\right\rangle_{\partial T}+\left(\nabla u,\nabla v_{0}\right)_{T}
+ε2(Δu)𝐧,v0vbTu𝐧,v0vbT].\displaystyle+\varepsilon^{2}\left\langle\nabla(\Delta u)\cdot\mathbf{n},v_{0}-v_{b}\right\rangle_{\partial T}-\left\langle\nabla u\cdot\mathbf{n},v_{0}-v_{b}\right\rangle_{\partial T}\Big{]}.

Upon applying the previously mentioned equation together with (5.2) and (5.3), we get

ε2(Δw𝒬Nu,Δwv)N+(w𝒬Nu,wv)N=(f,v0)+l1(u,v)l2(u,v)+l3(u,v).\displaystyle\varepsilon^{2}\left(\Delta_{w}\mathcal{Q}_{N}u,\Delta_{w}v\right)_{N}+\left(\nabla_{w}\mathcal{Q}_{N}u,\nabla_{w}v\right)_{N}=\left(f,v_{0}\right)+l_{1}(u,v)-l_{2}(u,v)+l_{3}(u,v).

By adding s(𝒬Nu,v)s\left(\mathcal{Q}_{N}u,v\right) to both sides of the above equation, we arrive at

a(𝒬Nu,v)=(f,v0)+l1(u,v)l2(u,v)+l3(u,v)+s(𝒬Nu,v).a\left(\mathcal{Q}_{N}u,v\right)=\left(f,v_{0}\right)+l_{1}(u,v)-l_{2}(u,v)+l_{3}(u,v)+s\left(\mathcal{Q}_{N}u,v\right). (5.4)

Subtracting (3.4) from (5.4) yields the error equation as follows

a(eN,v)=l1(u,v)l2(u,v)+l3(u,v)+s(𝒬Nu,v),\displaystyle a(e_{N},v)=l_{1}(u,v)-l_{2}(u,v)+l_{3}(u,v)+s(\mathcal{Q}_{N}u,v),

for all vVN0v\in V_{N}^{0}. This completes the derivation of (5.1). ∎

5.2 Error estimate

The following theorem is the estimate for the error function eNe_{N} in the triple-bar norm (3.5), which is an H2H^{2}-equivalent norm in VN0V_{N}^{0}.

Theorem 5.1.

Consider the weak Galerkin finite element solution from (3.4), denoted as uNVNu_{N}\in V_{N}. Assuming that uHk+1(Ω)u\in H^{k+1}(\Omega), it follows that there exists a constant CC such that

|eN|C(N(k1)lnk32N).{|||}e_{N}{|||}\leq C\left(N^{-(k-1)}\ln^{k-\frac{3}{2}}N\right). (5.5)
Proof.

Upon substituting v=eNv=e_{N} into the error equation (5.1), we derive the following equation,

|eN|2=l1(u,eN)l2(u,eN)+l3(u,eN)+s(𝒬Nu,eN).{|||}e_{N}{|||}^{2}=l_{1}(u,e_{N})-l_{2}(u,e_{N})+l_{3}(u,e_{N})+s(\mathcal{Q}_{N}u,e_{N}). (5.6)

Using the Cauchy-Schwarz inequality, the meshwidth characteristics (2.2) and the inequality (4.8), we derive

l1(u,eN)=T𝒯Nε2Δu𝒬0Δu,(e0𝐞g)𝐧Ti=1,2(T𝒯Nε2hΔu𝒬0ΔuTi2)12(T𝒯Nε2h1e0𝐞gTi2)12Cεh12i=1,2(T𝒯NΔu𝒬0ΔuTi2)12|eN|Cεh12(N(k32)+ε1N(k32)lnk32N+ε1N12α)|eN|C(εN(k1)+ε12N(k1)lnk1N+ε12N12α)|eN|.\begin{split}l_{1}\left(u,e_{N}\right)&=\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}\left\langle\Delta u-\mathcal{Q}_{0}\Delta u,\left(\nabla e_{0}-\mathbf{e}_{g}\right)\cdot\mathbf{n}\right\rangle_{\partial T}\\ &\leq\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h\|\Delta u-\mathcal{Q}_{0}\Delta u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-1}\|\nabla e_{0}-\mathbf{e}_{g}\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\\ &\leq C\varepsilon h^{\frac{1}{2}}\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\Delta u-\mathcal{Q}_{0}\Delta u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}{|||}e_{N}{|||}\\ &\leq C\varepsilon h^{\frac{1}{2}}\left(N^{-(k-\frac{3}{2})}+\varepsilon^{-1}N^{-(k-\frac{3}{2})}\ln^{k-\frac{3}{2}}N+\varepsilon^{-1}N^{\frac{1}{2}-\alpha}\right){|||}e_{N}{|||}\\ &\leq C\left(\varepsilon N^{-(k-1)}+\varepsilon^{\frac{1}{2}}N^{-(k-1)}\ln^{k-1}N+\varepsilon^{\frac{1}{2}}N^{\frac{1}{2}-\alpha}\right){|||}e_{N}{|||}.\end{split} (5.7)

From both the Cauchy-Schwarz inequality and (2.2) and the inequality (4.9), it can be deduced that

l2(u,eN)=T𝒯Nε2(Δu𝒬0Δu)𝐧,e0ebTi=1,2(T𝒯Nε2h2H(Δu𝒬0Δu)Ti2)12(T𝒯Nε2h2H1e0ebTi2)12CεhH12i=1,2(T𝒯N(Δu𝒬0Δu)Ti2)12|eN|CεhH12(N(k52)+ε2N(k52)lnk52N+ε2N32α)|eN|C(εN(k1)+N(k1)lnk32N+N1α)|eN|.\begin{split}l_{2}\left(u,e_{N}\right)&=\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}\left\langle\nabla\left(\Delta u-\mathcal{Q}_{0}\Delta u\right)\cdot\mathbf{n},e_{0}-e_{b}\right\rangle_{\partial T}\\ &\leq\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{2}H\|\nabla\left(\Delta u-\mathcal{Q}_{0}\Delta u\right)\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-2}H^{-1}\|e_{0}-e_{b}\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\\ &\leq C\varepsilon hH^{\frac{1}{2}}\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\nabla\left(\Delta u-\mathcal{Q}_{0}\Delta u\right)\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}{|||}e_{N}{|||}\\ &\leq C\varepsilon hH^{\frac{1}{2}}\left(N^{-(k-\frac{5}{2})}+\varepsilon^{-2}N^{-(k-\frac{5}{2})}\ln^{k-\frac{5}{2}}N+\varepsilon^{-2}N^{\frac{3}{2}-\alpha}\right){|||}e_{N}{|||}\\ &\leq C\left(\varepsilon N^{-(k-1)}+N^{-(k-1)}\ln^{k-\frac{3}{2}}N+N^{1-\alpha}\right){|||}e_{N}{|||}.\end{split} (5.8)

Similarly, it follows from the Cauchy-Schwarz inequality and (2.2) and (4.10) that

l3(u,eN)=T𝒯N(u𝐐Nu)𝐧,e0ebTi=1,2(T𝒯NHu𝐐NuTi2)12(T𝒯NH1e0ebTi2)12CH12i=1,2(T𝒯Nu𝐐NuTi2)12|eN|CH12(N(k12)+N(k12)lnk12N+N12α)|eN|C(Nk+Nklnk12N+Nα)|eN|.\begin{split}l_{3}\left(u,e_{N}\right)&=\sum_{T\in\mathcal{T}_{N}}\left\langle\left(\nabla u-\mathbf{Q}_{N}\nabla u\right)\cdot\mathbf{n},e_{0}-e_{b}\right\rangle_{\partial T}\\ &\leq\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}H\|\nabla u-\mathbf{Q}_{N}\nabla u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\left(\sum_{T\in\mathcal{T}_{N}}H^{-1}\|e_{0}-e_{b}\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\\ &\leq CH^{\frac{1}{2}}\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\nabla u-\mathbf{Q}_{N}\nabla u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}{|||}e_{N}{|||}\\ &\leq CH^{\frac{1}{2}}\left(N^{-(k-\frac{1}{2})}+N^{-(k-\frac{1}{2})}\ln^{k-\frac{1}{2}}N+N^{\frac{1}{2}-\alpha}\right){|||}e_{N}{|||}\\ &\leq C\left(N^{-k}+N^{-k}\ln^{k-\frac{1}{2}}N+N^{-\alpha}\right){|||}e_{N}{|||}.\end{split} (5.9)

In the same way, considering s(𝒬Nu,eN)s(\mathcal{Q}_{N}u,e_{N}), it follows from the Cauchy-Schwarz inequality and (2.2) and (4.11)-(4.12) that

|i=1,2T𝒯Nε2h1𝒬0u𝐐gu,e0𝐞gTi|\displaystyle\bigg{|}\sum_{i=1,2}\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-1}\left\langle\nabla\mathcal{Q}_{0}u-\mathbf{Q}_{g}\nabla u,\nabla e_{0}-\mathbf{e}_{g}\right\rangle_{\partial T_{i}}\bigg{|}
i=1,2(T𝒯Nε2h1𝒬0uuTi2)12(T𝒯Nε2h1e0𝐞gTi2)12\displaystyle\leq\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-1}\|\nabla\mathcal{Q}_{0}u-\nabla u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-1}\|\nabla e_{0}-\mathbf{e}_{g}\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}
Cεh12i=1,2(T𝒯N𝒬0uuTi2)12|eN|\displaystyle\leq C\varepsilon h^{-\frac{1}{2}}\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\nabla\mathcal{Q}_{0}u-\nabla u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}{|||}e_{N}{|||}
Cεh12(N(k12)+N(k12)lnk12N+N32α)|eN|\displaystyle\leq C\varepsilon h^{-\frac{1}{2}}\left(N^{-(k-\frac{1}{2})}+N^{-(k-\frac{1}{2})}\ln^{k-\frac{1}{2}}N+N^{\frac{3}{2}-\alpha}\right){|||}e_{N}{|||}
C(ε12N(k1)+ε12N(k1)lnk1N+ε12N2α)|eN|,\displaystyle\leq C\left(\varepsilon^{\frac{1}{2}}N^{-(k-1)}+\varepsilon^{\frac{1}{2}}N^{-(k-1)}\ln^{k-1}N+\varepsilon^{\frac{1}{2}}N^{2-\alpha}\right){|||}e_{N}{|||}, (5.10)
|i=1,2T𝒯Nε2h2H1𝒬0u𝒬bu,e0ebTi|\displaystyle\bigg{|}\sum_{i=1,2}\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-2}H^{-1}\left\langle\mathcal{Q}_{0}u-\mathcal{Q}_{b}u,e_{0}-e_{b}\right\rangle_{\partial T_{i}}\bigg{|}
i=1,2(T𝒯Nε2h2H1𝒬0uuTi2)12(T𝒯Nε2h2H1e0ebTi2)12\displaystyle\leq\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-2}H^{-1}\|\mathcal{Q}_{0}u-u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\left(\sum_{T\in\mathcal{T}_{N}}\varepsilon^{2}h^{-2}H^{-1}\|e_{0}-e_{b}\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}
Cεh1H12i=1,2(T𝒯N𝒬0uuTi2)12|eN|\displaystyle\leq C\varepsilon h^{-1}H^{-\frac{1}{2}}\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\mathcal{Q}_{0}u-u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}{|||}e_{N}{|||}
Cεh1H12(N(k+12)+εN(k+12)lnk+12N+εN12α)\displaystyle\leq C\varepsilon h^{-1}H^{-\frac{1}{2}}\left(N^{-(k+\frac{1}{2})}+\varepsilon N^{-(k+\frac{1}{2})}\ln^{k+\frac{1}{2}}N+\varepsilon N^{\frac{1}{2}-\alpha}\right)
C(N(k1)+εN(k1)lnk12N+εN2α),\displaystyle\leq C\left(N^{-(k-1)}+\varepsilon N^{-(k-1)}\ln^{k-\frac{1}{2}}N+\varepsilon N^{2-\alpha}\right), (5.11)

and

|i=1,2T𝒯NH1𝒬0u𝒬bu,e0ebTi|\displaystyle\bigg{|}\sum_{i=1,2}\sum_{T\in\mathcal{T}_{N}}H^{-1}\left\langle\mathcal{Q}_{0}u-\mathcal{Q}_{b}u,e_{0}-e_{b}\right\rangle_{\partial T_{i}}\bigg{|}
i=1,2(T𝒯NH1𝒬0uuTi2)12(T𝒯NH1e0ebTi2)12\displaystyle\leq\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}H^{-1}\|\mathcal{Q}_{0}u-u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}\left(\sum_{T\in\mathcal{T}_{N}}H^{-1}\|e_{0}-e_{b}\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}
CH12i=1,2(T𝒯N𝒬0uuTi2)12|eN|\displaystyle\leq CH^{-\frac{1}{2}}\sum_{i=1,2}\left(\sum_{T\in\mathcal{T}_{N}}\|\mathcal{Q}_{0}u-u\|_{\partial T_{i}}^{2}\right)^{\frac{1}{2}}{|||}e_{N}{|||}
CH12(N(k+12)+εN(k+12)lnk+12N+εN12α)\displaystyle\leq CH^{-\frac{1}{2}}\left(N^{-(k+\frac{1}{2})}+\varepsilon N^{-(k+\frac{1}{2})}\ln^{k+\frac{1}{2}}N+\varepsilon N^{\frac{1}{2}-\alpha}\right)
C(Nk+εNklnk+12N+εN1α).\displaystyle\leq C\left(N^{-k}+\varepsilon N^{-k}\ln^{k+\frac{1}{2}}N+\varepsilon N^{1-\alpha}\right). (5.12)

Substituting (5.7)-(5.2) into (5.6) yields

|eN|2C(N(k1)lnk32N)|eN|.\displaystyle{|||}e_{N}{|||}^{2}\leq C\left(N^{-(k-1)}\ln^{k-\frac{3}{2}}N\right){|||}e_{N}{|||}.

which implies (5.5). This completes the proof of the theorem. ∎

6 Numerical Experiments

In this section we compute two numerical examples where we select α=k+1\alpha=k+1 for creating the Shishkin mesh. Consider the singularly perturbed fourth-order problem that seeks solution u=u(x,y)u=u(x,y) satisfying

ε2Δ2uΔu\displaystyle\varepsilon^{2}\Delta^{2}u-\Delta u =f,inΩ,\displaystyle=f,\quad\text{in}~{}\Omega,
u=𝐧u\displaystyle u=\partial_{\mathbf{n}}u =0,onΩ,\displaystyle=0,\quad\text{on}~{}\partial\Omega,

where Ω=(0,1)2\Omega=(0,1)^{2}, ff and ε\varepsilon will choose later. Tables 1-6 display the errors |𝒬NuuN|{|||}\mathcal{Q}_{N}u-u_{N}{|||} for several different ε\varepsilon and NN.

Example 6.1.

Choose f(x,y)f(x,y) such that the exact solution is u(x,y)=g(x)g(y)u(x,y)=g(x)g(y), where

g(x)=12[sin(πx)+πε1e1/ε(ex/ε+e(x1)/ε1e1/ε)].g(x)=\frac{1}{2}\left[\sin(\pi x)+\frac{\pi\varepsilon}{1-e^{-1/\varepsilon}}\left(e^{-x/\varepsilon}+e^{(x-1)/\varepsilon}-1-e^{-1/\varepsilon}\right)\right].

The numerical results are shown in Table 1-3.

In the case of k=3k=3, the error and the order of convergence on Shishkin mesh and uniform mesh are listed in Table 1 and Table 2, respectively. Compared with the numerical results of the WG method for the problem on uniform mesh, our results with Shishkin mesh are better and an ε\varepsilon-independent asymptotically optimal order of convergence is achieved in all cases.

Table 1: Numerical results for Example 6.1 on Shishkin mesh.
ε\varepsilon N=8N=8 N=16N=16 N=32N=32 N=64N=64 N=128N=128
1e-00 1.01e-03 2.61e-04 6.58e-05 1.65e-05 4.12e-06
1.96 1.99 2.00 2.00
1e-01 3.77e-03 1.06e-03 2.75e-04 6.94e-05 1.74e-05
1.83 1.95 1.99 2.00
1e-02 1.17e-02 6.43e-03 3.03e-03 1.25e-03 4.59e-04
0.86 1.09 1.28 1.44
1e-03 3.81e-03 2.08e-03 9.73e-04 4.00e-04 1.46e-04
0.87 1.10 1.28 1.45
1e-04 1.22e-03 6.59e-04 3.08e-04 1.27e-04 4.64e-05
0.89 1.10 1.28 1.45
1e-05 4.18e-04 2.09e-04 9.75e-05 4.01e-05 1.47e-05
1.00 1.10 1.28 1.45
1e-06 2.09e-04 6.71e-05 3.09e-05 1.27e-05 4.64e-06
1.64 1.12 1.28 1.45
1e-07 1.74e-04 2.44e-05 9.84e-06 4.01e-06 1.47e-06
2.84 1.31 1.29 1.45
Table 2: Numerical results for Example 6.1 on uniform mesh.
ε\varepsilon N=8N=8 N=16N=16 N=32N=32 N=64N=64 N=128N=128
1e-00 1.01e-03 2.61e-04 6.58e-05 1.65e-05 4.12e-06
1.96 1.99 2.00 2.00
1e-01 3.77e-03 1.06e-03 2.75e-04 6.94e-05 1.74e-05
1.83 1.95 1.99 2.00
1e-02 3.87e-02 2.06e-02 8.03e-03 2.62e-03 7.53e-04
0.91 1.36 1.62 1.80
1e-03 1.24e-02 1.60e-02 1.73e-02 1.44e-02 8.63e-03
-0.37 -0.12 0.26 0.74
1e-04 1.30e-03 1.83e-03 2.57e-03 3.56e-03 4.74e-03
-0.49 -0.49 -0.47 -0.41
1e-05 1.33e-04 1.85e-04 2.62e-04 3.70e-04 5.21e-04
-0.47 -0.50 -0.50 -0.49
1e-06 1.96e-05 1.87e-05 2.62e-05 3.71e-05 5.25e-05
0.07 -0.49 -0.50 -0.50
1e-07 1.30e-05 2.24e-06 2.63e-06 3.71e-06 5.25e-06
2.54 -0.23 -0.50 -0.50

In the case of k=4k=4, the numerical results are shown in Table 3, which has yielded the asymptotically optimal order.

Table 3: Numerical results for Example 6.1 on Shishkin mesh.
ε\varepsilon N=8N=8 N=16N=16 N=32N=32 N=64N=64
1e-00 3.07e-05 3.90e-06 4.89e-07 6.12e-08
2.98 3.00 3.00
1e-01 3.92e-04 5.35e-05 6.84e-06 8.61e-07
2.87 2.97 2.99
1e-02 6.08e-03 2.56e-03 8.25e-04 2.11e-04
1.25 1.63 1.97
1e-03 1.98e-03 8.29e-04 2.66e-04 6.77e-05
1.26 1.64 1.97
1e-04 6.28e-04 2.63e-04 8.43e-05 2.15e-05
1.26 1.64 1.97
1e-05 1.99e-04 8.32e-05 2.67e-05 6.79e-06
1.26 1.64 1.97
1e-06 6.33e-05 2.63e-05 8.44e-06 2.24e-06
1.27 1.64 1.91
Example 6.2.

Let g(x)g(x) be as in Example 6.1 and set

p(y)=2y(1y2)+ε[ld(12y)3ql+(3ld)ey/ε+(3l+d)e(y1)/ε],p(y)=2y(1-y^{2})+\varepsilon\left[ld(1-2y)-3\frac{q}{l}+\left(\frac{3}{l}-d\right)e^{-y/\varepsilon}+\left(\frac{3}{l}+d\right)e^{(y-1)/\varepsilon}\right],

with l=1e1/εl=1-e^{-1/\varepsilon}, q=2lq=2-l and d=1/(q2εl)d=1/(q-2\varepsilon l). Then choose f(x,y)f(x,y) such that the exact solution of (1.1) is u(x,y)=g(x)p(y)u(x,y)=g(x)p(y). The numerical results are shown in Table 4-6.

The error and the order of convergence on Shishkin mesh and uniform mesh with k=3k=3 are listed in Table 4 and Table 5, respectively. And also we get better results on Shishkin mesh than uniform mesh.

Table 4: Numerical results for Example 6.2 on Shishkin mesh.
ε\varepsilon N=8N=8 N=16N=16 N=32N=32 N=64N=64 N=128N=128
1e-00 1.66e-04 4.24e-05 1.07e-05 2.67e-06 6.67e-07
1.97 1.99 2.00 2.00
1e-01 6.48e-03 1.82e-03 4.72e-04 1.19e-04 2.99e-05
1.83 1.94 1.99 2.00
1e-02 2.10e-02 1.16e-02 5.44e-03 2.24e-03 8.25e-04
0.86 1.09 1.28 1.44
1e-03 6.86e-03 3.74e-03 1.75e-03 7.20e-04 2.64e-04
0.87 1.10 1.28 1.45
1e-04 2.18e-03 1.19e-03 5.55e-04 2.28e-04 8.35e-05
0.88 1.10 1.28 1.45
1e-05 7.13e-04 3.76e-04 1.76e-04 7.22e-05 2.64e-05
0.92 1.10 1.28 1.45
1e-06 2.87e-04 1.20e-04 5.56e-05 2.28e-05 8.36e-06
1.27 1.11 1.28 1.45
1e-07 2.00e-04 4.01e-05 1.76e-05 7.29e-06 2.64e-06
2.32 1.19 1.27 1.46
Table 5: Numerical results for Example 6.2 on uniform mesh.
ε\varepsilon N=8N=8 N=16N=16 N=32N=32 N=64N=64 N=128N=128
1e-00 1.66E-04 4.24E-05 1.07E-05 2.67E-06 6.67E-07
1.97 1.99 2.00 2.00
1e-01 6.48E-03 1.82E-03 4.72E-04 1.19E-04 2.99E-05
1.83 1.94 1.99 2.00
1e-02 6.96E-02 3.70E-02 1.44E-02 4.71E-03 1.35E-03
0.91 1.36 1.62 1.80
1e-03 2.22E-02 2.88E-02 3.12E-02 2.60E-02 1.56E-02
-0.37 -0.12 0.26 0.74
1e-04 2.35E-03 3.30E-03 4.63E-03 6.41E-03 8.54E-03
-0.49 -0.49 -0.47 -0.41
1e-05 2.37E-04 3.34E-04 4.72E-04 6.66E-04 9.38E-04
-0.49 -0.50 -0.50 -0.49
1e-06 2.85E-05 3.35E-05 4.73E-05 6.69E-05 9.45E-05
-0.23 -0.50 -0.50 -0.50
1e-07 1.45E-05 3.62E-06 4.73E-06 6.69E-06 9.46E-06
2.00 -0.39 -0.50 -0.50

In the case of k=4k=4, the numerical results are shown in Table 6, which has also yielded the asymptotically optimal order, independent of ε\varepsilon.

Table 6: Numerical results for Example 6.2 on Shishkin mesh.
ε\varepsilon N=8N=8 N=16N=16 N=32N=32 N=64N=64
1e-00 3.84e-06 4.86e-07 6.09e-08 7.62e-09
2.98 3.00 3.00
1e-01 6.93e-04 9.45e-05 1.21e-05 1.52e-06
2.87 2.97 2.99
1e-02 1.09e-02 4.60e-03 1.48e-03 3.79e-04
1.25 1.63 1.97
1e-03 3.57e-03 1.49e-03 4.79e-04 1.22e-04
1.26 1.64 1.97
1e-04 1.13e-03 4.74e-04 1.52e-04 3.87e-05
1.26 1.64 1.97
1e-05 3.58e-04 1.50e-04 4.81e-05 1.22e-05
1.26 1.64 1.97
1e-06 1.14e-04 4.74e-05 2.53e-05 3.90e-06
1.26 0.90 2.70
Remark 6.1.

The numerical examples show that we obtain the asymptotically optimal error estimate. In fact, the lnk32N\ln^{k-\frac{3}{2}}N in the error results will have an impact on the convergence order, and with the refinement of the Shishkin mesh, the influence of lnk32N\ln^{k-\frac{3}{2}}N on the convergence order will gradually become smaller. The convergence order is reduced by about half order when k=3k=3 and by one order in the case of k=4k=4. Our numerical results also confirm this well.

7 Conclusion

In this paper, we use the weak Galerkin finite element method to solve the singularly perturbed fourth-order boundary value problem in 2D domain. By constructing the Shishkin mesh which is suitable for the problem, we give the corresponding numerical algorithm and the error estimate in the H2H^{2} discrete norm. Compared with solving the problem on the uniform mesh, the WG method obtains better results on the Shishkin mesh, which is verified by the numerical results. Moreover, the results of our numerical experiments are consistent with the error estimation theory, and the asymptotically optimal convergence order is obtained.

Statements and Declarations

Funding. This work was supported by the National Natural Science Foundation of China (Grant No. 12101039, 12271208).

Data Availability. The code used in this work will be made available upon request to the authors.

Competing Interests. The authors have no relevant financial or non-financial interests to disclose.

References

  • [1] N. S. Bahvalov, On the optimization of the methods for solving boundary value problems in the presence of a boundary layer, Ž. Vyčisl. Mat i Mat. Fiz., In Russian, 9 (1969), pp. 841–859.
  • [2] S. C. Brenner, T. Gudi, M. Neilan, and L.-y. Sung, C0C^{0} penalty methods for the fully nonlinear monge-ampère equation, Math. Comp., 80 (2011), pp. 1979–1995.
  • [3] S. C. Brenner and M. Neilan, A C0C^{0} interior penalty method for a fourth order elliptic singular perturbation problem, SIAM J. Numer. Anal., 49 (2011), pp. 869–892.
  • [4] G. Chen and X. Xie, A robust weak Galerkin finite element method for linear elasticity with strong symmetric stresses, Comput. Methods Appl. Math., 16 (2016), pp. 389–408.
  • [5] H. Chen, S. Chen, and L. Xiao, Uniformly convergent C0C^{0}-nonconforming triangular prism element for fourth-order elliptic singular perturbation problem, Numer. Methods Partial Differential Equations, 30 (2014), pp. 1785–1796.
  • [6] L. Chen, J. Wang, Y. Wang, and X. Ye, An auxiliary space multigrid preconditioner for the weak Galerkin method, Comput. Math. Appl., 70 (2015), pp. 330–344.
  • [7] P. Constantinou and C. Xenophontos, An hphp finite element method for a 4th order singularly perturbed boundary value problem in two dimensions, Comput. Math. Appl., 74 (2017), pp. 1565–1575.
  • [8] M. Cui, X. Ye, and S. Zhang, A modified weak Galerkin finite element method for the biharmonic equation on polytopal meshes, Commun. Appl. Math. Comput., 3 (2021), pp. 91–105.
  • [9] M. Cui and S. Zhang, On the uniform convergence of the weak Galerkin finite element method for a singularly-perturbed biharmonic equation, J. Sci. Comput., 82 (2020), pp. Paper No. 5, 15.
  • [10] S. Franz and H.-G. Roos, Robust error estimation in energy and balanced norms for singularly perturbed fourth order problems, Comput. Math. Appl., 72 (2016), pp. 233–247.
  • [11] S. Franz, H.-G. Roos, and A. Wachtel, A C0C^{0} interior penalty method for a singularly-perturbed fourth-order elliptic problem on a layer-adapted mesh, Numer. Methods Partial Differential Equations, 30 (2014), pp. 838–861.
  • [12] E. H. Georgoulis, Discontinuous Galerkin methods on shape-regular and anisotropic meshes, PhD thesis, University of Oxford D. Phil. Thesis, 2003.
  • [13] J. Guzmán, D. Leykekhman, and M. Neilan, A family of non-conforming elements and the analysis of Nitsche’s method for a singularly perturbed fourth order problem, Calcolo, 49 (2012), pp. 95–125.
  • [14] G. Harper, J. Liu, S. Tavener, and B. Zheng, Lowest-order weak Galerkin finite element methods for linear elasticity on rectangular and brick meshes, J. Sci. Comput., 78 (2019), pp. 1917–1941.
  • [15] S. C. Hongru Chen, Uniformly convergent nonconforming element for 3D fourth order elliptic singular perturbation problem, J. Comput. Math., 32 (2014), pp. 687–695.
  • [16] X. Hu, L. Mu, and X. Ye, A weak Galerkin finite element method for the Navier-Stokes equations, J. Comput. Appl. Math., 362 (2019), pp. 614–625.
  • [17] W. Huang and Y. Wang, Discrete maximum principle for the weak Galerkin method for anisotropic diffusion problems, Commun. Comput. Phys., 18 (2015), pp. 65–90.
  • [18] X. Liu, J. Li, and Z. Chen, A weak Galerkin finite element method for the Navier-Stokes equations, J. Comput. Appl. Math., 333 (2018), pp. 442–457.
  • [19] X. Meng and M. Stynes, Convergence analysis of the Adini element on a Shishkin mesh for a singularly perturbed fourth-order problem in two dimensions, Adv. Comput. Math., 45 (2019), pp. 1105–1128.
  • [20] J. J. H. Miller, E. O’Riordan, and G. I. Shishkin, Fitted numerical methods for singular perturbation problems, World Scientific Publishing Co., Inc., River Edge, NJ, 1996.
  • [21] L. Mu, J. Wang, and X. Ye, A stable numerical algorithm for the Brinkman equations by weak Galerkin finite element methods, J. Comput. Phys., 273 (2014), pp. 327–342.
  • [22] 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.
  • [23] T. K. Nilssen, X.-C. Tai, and R. Winther, A robust nonconforming H2H^{2}-element, Math. Comp., 70 (2001), pp. 489–505.
  • [24] P. Panaseti, A. Zouvani, N. Madden, and C. Xenophontos, A C1C^{1}-conforming hp finite element method for fourth order singularly perturbed boundary value problems, Appl. Numer. Math., 104 (2016), pp. 81–97.
  • [25] H.-G. Roos, M. Stynes, and L. Tobiska, Robust numerical methods for singularly perturbed differential equations, vol. 24 of Springer Series in Computational Mathematics, Springer-Verlag, Berlin, second ed., 2008.
  • [26] G. I. Shishkin, Grid approximation of singularly perturbed elliptic and parabolic equations, Keldysh Institute of Applied Mathematics, Moscow, In Russian, 1990.
  • [27] C. Wang, J. Wang, R. Wang, and R. Zhang, A locking-free weak Galerkin finite element method for elasticity problems in the primal formulation, J. Comput. Appl. Math., 307 (2016), pp. 346–366.
  • [28] 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.
  • [29] J. Wang and X. Ye, A weak Galerkin finite element method for the stokes equations, Adv. Comput. Math., 42 (2016), pp. 155–174.
  • [30] J. Wang, X. Ye, Q. Zhai, and R. Zhang, Discrete maximum principle for the P1P_{1}-P0P_{0} weak Galerkin finite element approximations, J. Comput. Phys., 362 (2018), pp. 114–130.
  • [31] L. Wang, Y. WU, and X. Xie, Uniformly stable rectangular elements for fourth order elliptic singular perturbation problems, Numer. Methods Partial Differential Equations, 29 (2013), pp. 721–737.
  • [32] R. Wang, X. Wang, Q. Zhai, and R. Zhang, A weak Galerkin finite element scheme for solving the stationary Stokes equations, J. Comput. Appl. Math., 302 (2016), pp. 171–185.
  • [33] R. Wang, R. Zhang, X. Zhang, and Z. Zhang, Supercloseness analysis and polynomial preserving recovery for a class of weak Galerkin methods, Numer. Methods Partial Differential Equations, 34 (2018), pp. 317–335.
  • [34] W. Wang, X. Huang, K. Tang, and R. Zhou, Morley-Wang-Xu element methods with penalty for a fourth order elliptic singular perturbation problem, Adv. Comput. Math., 44 (2018), pp. 1041–1061.
  • [35] X. Wang, Q. Zhai, and R. Zhang, The weak Galerkin method for solving the incompressible Brinkman flow, J. Comput. Appl. Math., 307 (2016), pp. 13–24.
  • [36] Y. Wang, Y. Li, and X. Meng, An upwind finite volume element method on a Shishkin mesh for singularly perturbed convection-diffusion problems, J. Comput. Appl. Math., 438 (2024), pp. Paper No. 115493, 20.
  • [37] Z. Wang, R. Wang, and J. Liu, Robust weak Galerkin finite element solvers for Stokes flow based on a lifting operator, Comput. Math. Appl., 125 (2022), pp. 90–100.
  • [38] P. Xie, D. Shi, and H. Li, A new robust C0C^{0}-type nonconforming triangular element for singular perturbation problems, Appl. Math. Comput., 217 (2010), pp. 3832–3843.
  • [39] S.-Y. Yi, A lowest-order weak Galerkin method for linear elasticity, J. Comput. Appl. Math., 350 (2019), pp. 286–298.
  • [40] Q. Zhai, R. Zhang, and L. Mu, A new weak Galerkin finite element scheme for the Brinkman model, Commun. Comput. Phys., 19 (2016), pp. 1409–1434.
  • [41] J. Zhang and X. Liu, Uniform convergence of a weak Galerkin finite element method on Shishkin mesh for singularly perturbed convection-diffusion problems in 2D, Appl. Math. Comput., 432 (2022), pp. Paper No. 127346, 12.
  • [42] J. Zhang and X. Liu, Uniform convergence of a weak Galerkin method for singularly perturbed convection-diffusion problems, Math. Comput. Simulation, 200 (2022), pp. 393–403.
  • [43] J. Zhang, K. Zhang, J. Li, and X. Wang, A weak Galerkin finite element method for the Navier-Stokes equations, Commun. Comput. Phys., 23 (2018), pp. 706–746.
  • [44] R. Zhang and Q. Zhai, A weak Galerkin finite element scheme for the biharmonic equations by using polynomials of reduced order, J. Sci. Comput., 64 (2015), pp. 559–585.
  • [45] T. Zhang and T. Lin, The weak Galerkin finite element method for incompressible flow, J. Math. Anal. Appl., 464 (2018), pp. 247–265.
  • [46] P. Zhu and S. Xie, A uniformly convergent weak Galerkin finite element method on Shishkin mesh for 1D convection-diffusion problem, J. Sci. Comput., 85 (2020), pp. Paper No. 34, 22.