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

An arbitrary order Reconstructed Discontinuous Approximation to Fourth-order Curl Problem

Ruo Li CAPT, LMAM and School of Mathematical Sciences, Peking University, Beijing 100871, P.R. China [email protected] Qicheng Liu School of Mathematical Sciences, Peking University, Beijing 100871, P.R. China [email protected]  and  Shuhai Zhao School of Mathematical Sciences, Peking University, Beijing 100871, P.R. China [email protected]
Abstract.

We present an arbitrary order discontinuous Galerkin finite element method for solving the fourth-order curl problem using a reconstructed discontinuous approximation method. It is based on an arbitrarily high-order approximation space with one unknown per element in each dimension. The discrete problem is based on the symmetric IPDG method. We prove a priori error estimates under the energy norm and the L2L^{2} norm and show numerical results to verify the theoretical analysis.
keywords: fourth-order curl problem, patch reconstruction, discontinuous Galerkin method

1. Introduction

We are concerned in this paper with the fourth-order curl problem, which has applications in inverse electromagnetic scattering, and magnetohydrodynamics (MHD) when modeling magnetized plasmas. Discretizing the fourth-order curl operator is one of the keys to simulate these models. Additionally, the fourth-order curl operator plays an important role in approximating the Maxwell transmission eigenvalue problem. Therefore, it is important to design highly efficient and accurate numerical methods for fourth-order curl problems [2, 17].

Finite element methods (FEMs) are a widely used numerical scheme for solving partial differential equations. The design of FEMs for fourth-order curl problems is challenging. Nevertheless, in recent years, researchers have proposed and analysed various FEMs for these problems to address this difficulty. According to whether the finite element space is contained within the space of the exact solution, finite element methods can be classified into conforming methods and nonconforming methods. For conforming methods, one needs to construct H(curl2)H(\mathrm{curl}^{2})-conforming finite elements, we refer to [22, 6] for some recent works. Due to the high order of the fourth-order curl operator, it is still a difficult task to implement a high-order conforming space. Therefore, many researches focus on nonconforming methods. The discontinuous Galerkin (DG) methods use completely discontinuous piecewise polynomial to approximate the solution of partial differential equations. These methods enforce numerical solutions to be close to the exact solution by adding penalties. DG methods are commonly used in solving complex problems due to their simplicity and flexibility in approximation space. For fourth-order curl problems, we refer to [24, 5, 21, 20, 23] for some DG methods.

A significant drawback of DG methods is the large number of degrees of freedom in DG space, which results in high computational costs. This drawback is a matter of concern. In this paper, we propose an arbitrary order discontinuous Galerkin FEM for solving the fourth-order curl problem. The method is based on a reconstructed approximation space that has only one unknown per dimension in each element. The construction of the approximation space includes creating an element patch per element and solving a local least squares fitting problem to obtain a local high-order polynomial. Methods based on the reconstructed spaces are called reconstructed discontinuous approximation (RDA) methods, which have been successfully applied to a series of classical problems [13, 10, 16, 15, 8, 14]. The reconstructed space is a subspace of the standard DG space, which can approximate functions to high-order accuracy and inherits the flexibility on the mesh partition in the meanwhile. One advantage of this space is that it has very few degrees of freedom, which gives high approximation efficiency of finite element. Using this new space, we design the discrete scheme for the fourth-order curl problem under the symmetric interior penalty discontinuous Galerkin (IPDG) framework. We prove the convergence rates under the energy norm and the L2L^{2} norm, and numerical experiments are conducted to verify the theoretical analysis ,showing our algorithm is simple to implement and can reach high-order accuracy.

The rest of this paper is organized as follows. In Section 2, we introduce the fourth-order curl problem without div-free condition and give the basic notations about the Sobolev spaces and the partition. We also recall two commonly used inequalities in this section. In Section 3, we establish the reconstruction operator and the corresponding approximation space. Some basic properties of the reconstruction are also proven. In Section 4, we define the discrete variational form for the fourth-order curl problem and analyse the error under the energy norm and the L2L^{2} norm, and prove that the convergence rate is optimal with respect to the energy norm. In Section 5, we carried out some numerical examples to validate our theoretical results. Finally, a brief conclusion is given in Section 6.

2. Preliminaries

Let Ωd(d=2,3)\Omega\subset\mathbb{R}^{d}(d=2,3) be a bounded polygonal (polyhedral) domain with a Lipschitz boundary Ω\partial\Omega.
Given a bounded domain DΩD\subset\Omega, we follow the standard definitions to the space L2(D),L2(D)dL^{2}(D),L^{2}(D)^{d}, the spaces Hq(D),Hq(D)dH^{q}(D),H^{q}(D)^{d} with the regular exponent q0q\geq 0. The corresponding norms and seminorms are defined as

Hq(D):=(Hq(D)2)1/2,||Hq(D):=(||Hq(D)2)1/2,\|\cdot\|_{H^{q}(D)}:=\left(\|\cdot\|_{H^{q}(D)}^{2}\right)^{1/2},\,\,|\cdot|_{H^{q}(D)}:=\left(|\cdot|_{H^{q}(D)}^{2}\right)^{1/2},

respectively.

Throughout this paper, for two vectors 𝒂=(a1,,ad)d,𝒃=(b1,,bd)d\boldsymbol{a}=(a_{1},\ldots,a_{d})\in\mathbb{R}^{d},\boldsymbol{b}=(b_{1},\ldots,b_{d})\in\mathbb{R}^{d}, the cross product 𝒂×𝒃\boldsymbol{a}\times\boldsymbol{b} is defined as

𝒂×𝒃:={a1b2a2b1,d=2,(a2b3a3b2,a3b1a1b3,a1b2a2b1)T,d=3.\boldsymbol{a}\times\boldsymbol{b}:=\begin{cases}a_{1}b_{2}-a_{2}b_{1},\quad d=2,\\ (a_{2}b_{3}-a_{3}b_{2},a_{3}b_{1}-a_{1}b_{3},a_{1}b_{2}-a_{2}b_{1})^{T},\quad d=3.\\ \end{cases}

The cross product between a vector and a scalar will be used in two dimensions, and in this case for a vector 𝒂=(a1,a2)2\boldsymbol{a}=(a_{1},a_{2})\in\mathbb{R}^{2} and a scalar bb\in\mathbb{R}, 𝒂×b\boldsymbol{a}\times b is defined as 𝒂×b:=(a2b,a1b)T\boldsymbol{a}\times b:=(a_{2}b,-a_{1}b)^{T}. Specifically, for a vector-valued function 𝒖d\boldsymbol{u}\in\mathbb{R}^{d}, the curl of 𝒖\boldsymbol{u} reads

curl𝒖:=×𝒖={u2xu1y,d=2,(u3yu2z,u1zu3x,u2xu1y)T,d=3,\mathrm{curl}\boldsymbol{u}:=\nabla\times\boldsymbol{u}=\begin{cases}\frac{\partial u_{2}}{\partial x}-\frac{\partial u_{1}}{\partial y},\quad d=2,\\ (\frac{\partial u_{3}}{\partial y}-\frac{\partial u_{2}}{\partial z},\frac{\partial u_{1}}{\partial z}-\frac{\partial u_{3}}{\partial x},\frac{\partial u_{2}}{\partial x}-\frac{\partial u_{1}}{\partial y})^{T},\quad d=3,\end{cases}

and for the scalar-valued function q(x,y)q(x,y), we let ×q\nabla\times q be the formal adjoint, which reads

×q=(qy,qx)T.\nabla\times q=\left(\frac{\partial q}{\partial y},-\frac{\partial q}{\partial x}\right)^{T}.

For convenience, we mainly use the notations for the three-dimensional case.

In this paper, we consider the following fourth-order curl problem

(1) {curl4𝒖+𝒖=𝒇, in Ω,𝒖×𝒏=g1, on Ω,(×𝒖)×𝒏=g2, on Ω,\left\{\begin{aligned} &\mathrm{curl}^{4}\boldsymbol{u}+\boldsymbol{u}=\boldsymbol{f},&\quad\text{ in }\Omega,\\ &\boldsymbol{u}\times\boldsymbol{n}=g_{1},&\quad\text{ on }\partial\Omega,\\ &(\nabla\times\boldsymbol{u})\times\boldsymbol{n}=g_{2},&\quad\text{ on }\partial\Omega,\end{aligned}\right.

For the problem domain Ω\Omega, we define the space

H(curls,Ω)\displaystyle H(\mathrm{curl}^{s},\Omega) :={𝒗L2(Ω)d|curlj𝒗L2(Ω),1js},\displaystyle:=\left\{\boldsymbol{v}\in L^{2}(\Omega)^{d}\ |\ \mathrm{curl}^{j}\boldsymbol{v}\in L^{2}(\Omega),1\leq j\leq s\right\},

and

H0(curl2,Ω)\displaystyle H_{0}(\mathrm{curl}^{2},\Omega) :={𝒗L2(Ω)d|curlj𝒗L2(Ω),curlj1𝒗=0onΩ,1j2},\displaystyle:=\left\{\boldsymbol{v}\in L^{2}(\Omega)^{d}\ |\ \mathrm{curl}^{j}\boldsymbol{v}\in L^{2}(\Omega),\ \mathrm{curl}^{j-1}\boldsymbol{v}=0\ \text{on}\ \partial\Omega,1\leq j\leq 2\right\},

For the weak formulation we are going to find 𝒖H(curl2,Ω)\boldsymbol{u}\in H(\mathrm{curl}^{2},\Omega) such that

a(𝒖,𝒗)=(𝒇,𝒗)L2(Ω),𝒗H0(curl2,Ω),\displaystyle a(\boldsymbol{u},\boldsymbol{v})=(\boldsymbol{f},\boldsymbol{v})_{L^{2}(\Omega)},\ \forall\boldsymbol{v}\in H_{0}(\mathrm{curl}^{2},\Omega),

where

a(𝒖,𝒗):=(curl2𝒖,curl2𝒗)L2(Ω)+(𝒖,𝒗)L2(Ω).a(\boldsymbol{u},\boldsymbol{v}):=(\mathrm{curl}^{2}\boldsymbol{u},\mathrm{curl}^{2}\boldsymbol{v})_{L^{2}(\Omega)}+(\boldsymbol{u},\boldsymbol{v})_{L^{2}(\Omega)}.

We refer to [18, 23] for some regularity results of this problem.

Next, we define some notations about the mesh. Let 𝒯h\mathcal{T}_{h} be a regular and quasi-uniform partition Ω\Omega into disjoint open triangles (tetrahedra). Let h\mathcal{E}_{h} denote the set of all d1d-1 dimensional faces of 𝒯h\mathcal{T}_{h}, and we decompose h\mathcal{E}_{h} into h=hihb\mathcal{E}_{h}=\mathcal{E}_{h}^{i}\cup\mathcal{E}_{h}^{b}, where hi\mathcal{E}_{h}^{i} and hb\mathcal{E}_{h}^{b} are the sets of interior faces and boundary faces, respectively. We let

hK:=diam(K),K𝒯h,he:=diam(e),eh,h_{K}:=\text{diam}(K),\quad\forall K\in\mathcal{T}_{h},\quad h_{e}:=\text{diam}(e),\quad\forall e\in\mathcal{E}_{h},

and define h:=maxK𝒯hhKh:=\max_{K\in\mathcal{T}_{h}}h_{K}. The quasi-uniformity of the mesh 𝒯h\mathcal{T}_{h} is in the sense that there exists a constant ν>0\nu>0 such that hνminK𝒯hρKh\leq\nu\min_{K\in\mathcal{T}_{h}}\rho_{K}, where ρK\rho_{K} is the diameter of the largest ball inscribed in KK. One can get the inverse inequality and the trace inequality from the regularity of the mesh, which are commonly used in the analysis.

Lemma 1.

There exists a constant CC independent of the mesh size hh, such that

(2) vL2(K)2C(hK1vL2(K)2+hKvL2(K)2),vH1(K).\|v\|^{2}_{L^{2}(\partial K)}\leq C\left(h_{K}^{-1}\|v\|^{2}_{L^{2}(K)}+h_{K}\|\nabla v\|_{L^{2}(K)}^{2}\right),\quad\forall v\in H^{1}(K).
Lemma 2.

There exists a constant CC independent of the mesh size hh, such that

(3) vHq(K)ChKqvL2(K),vl(K),\|v\|_{H^{q}(K)}\leq Ch_{K}^{-q}\|v\|_{L^{2}(K)},\quad\forall v\in\mathbb{P}_{l}(K),

where l(K)\mathbb{P}_{l}(K) is the space of polynomial on KK with the degree no more than ll.

We refer to [1] for more details of these inequalities.

Finally, we introduce the trace operators that will be used in our numerical schemes. For ehie\in\mathcal{E}_{h}^{i}, we denote by K+K^{+} and KK^{-} the two neighbouring elements that share the boundary ee, and 𝐧+,𝐧\boldsymbol{\mathrm{n}}^{+},\boldsymbol{\mathrm{n}}^{-} the unit out normal vector on ee, respectively. We define the jump operator [[]][[\cdot]] and the average operator {}\{\cdot\} as

(4) [[𝒒]]\displaystyle[[\boldsymbol{q}]] :=𝒒|K+×𝐧++𝒒|K×𝐧, for vector-valued function,\displaystyle:=\boldsymbol{q}|_{K^{+}}\times\boldsymbol{\mathrm{n}}^{+}+\boldsymbol{q}|_{K^{-}}\times\boldsymbol{\mathrm{n}}^{-},\quad\text{ for vector-valued function},
[[curl𝒒]]\displaystyle[[\mathrm{curl}\boldsymbol{q}]] :=curl𝒒|K+×𝐧++curl𝒒|K×𝐧, for vector-valued function,\displaystyle:=\mathrm{curl}\boldsymbol{q}|_{K^{+}}\times\boldsymbol{\mathrm{n}}^{+}+\mathrm{curl}\boldsymbol{q}|_{K^{-}}\times\boldsymbol{\mathrm{n}}^{-},\quad\text{ for vector-valued function},
{v}\displaystyle\{v\} :=12(v|K++v|K), for scalar-valued function,\displaystyle:=\frac{1}{2}(v|_{K^{+}}+v|_{K^{-}}),\quad\text{ for scalar-valued function},
{𝒒}\displaystyle\{\boldsymbol{q}\} :=12(𝒒|K++𝒒|K), for vector-valued function.\displaystyle:=\frac{1}{2}(\boldsymbol{q}|_{K^{+}}+\boldsymbol{q}|_{K^{-}}),\quad\text{ for vector-valued function}.

For ehbe\in\mathcal{E}_{h}^{b}, we let K𝒯hK\in\mathcal{T}_{h} such that eKe\in\partial K and 𝐧\boldsymbol{\mathrm{n}} is the unit out normal vector. We define

(5) [[𝒒]]:=𝒒|K×𝐧,[[curl𝒒]]:=curl𝒒|K×𝐧 for vector-valued function,\displaystyle[[\boldsymbol{q}]]:=\boldsymbol{q}|_{K}\times\boldsymbol{\mathrm{n}},\quad[[\mathrm{curl}\boldsymbol{q}]]:=\mathrm{curl}\boldsymbol{q}|_{K}\times\boldsymbol{\mathrm{n}}\text{ for vector-valued function},
{v}:=v|K for scalar-valued function,{𝒒}:=𝒒|K for vector-valued function.\displaystyle\{v\}:=v|_{K}\quad\text{ for scalar-valued function},\qquad\{\boldsymbol{q}\}:=\boldsymbol{q}|_{K}\quad\text{ for vector-valued function}.

Throughout this paper, CC and CC with subscripts denote the generic constants that may differ between lines but are independent of the mesh size.

3. Reconstructed Discontinuous Space

In this section, we will introduce a linear reconstruction operator to obtain a discontinuous approximation space for the given mesh 𝒯h\mathcal{T}_{h}. The reconstructed space can achieve a high-order accuracy while the number of degrees of freedom in each dimension remain the same as the number of elements in 𝒯h\mathcal{T}_{h}. The construction of the operator includes the following steps.

Step 1. For each K𝒯hK\in\mathcal{T}_{h}, we construct an element patch S(K)S(K), which consists of KK itself and some surrounding elements. The size of the patch is controlled with a given threshold #S\#S. The construction of the element patch S(K)S(K) is conducted by a recursive algorithm. We begin by setting S0(K)={K}S_{0}(K)=\{K\}, and define St(K)S_{t}(K) recursively:

(6) St(K)=KSt1(K)K′′Δ(K)K′′,t=0,1,S_{t}(K)=\bigcup_{K^{\prime}\in S_{t-1}(K)}\bigcup_{K^{\prime\prime}\in\Delta(K^{\prime})}K^{\prime\prime},\quad t=0,1,\dots

where Δ(K):={K𝒯h|K¯K¯}\Delta(K):=\{K^{\prime}\in\mathcal{T}_{h}\ |\ \overline{K}\cap\overline{K^{\prime}}\neq\emptyset\}. The recursion stops once tt meets the condition that the cardinality #St(K)#S\#S_{t}(K)\geq\#S, and we let the patch S(K):=St(K)S(K):=S_{t}(K). We apply the recursive algorithm (6) to all elements in 𝒯h\mathcal{T}_{h}.

Step 2. For each K𝒯hK\in\mathcal{T}_{h}, we solve a local least squares fitting problem on the patch S(K)S(K). We let 𝒙K\boldsymbol{x}_{K} be the barycenter of the element KK and mark barycenters of all elements as collocation points. Let I(K)I(K) be the set of collocation points located inside the domain of S(K)S(K),

I(K):={𝒙K|KS(K)},I(K):=\{\boldsymbol{x}_{K^{\prime}}\ |\ K^{\prime}\in S(K)\},

Let Uh0U_{h}^{0} be the piecewise constant space with respect to 𝒯h\mathcal{T}_{h},

Uh0:={vhL2(Ω)|vh|K0(K),K𝒯h}.U_{h}^{0}:=\{v_{h}\in L^{2}(\Omega)\ |\ v_{h}|_{K}\in\mathbb{P}_{0}(K),\ \forall K\in\mathcal{T}_{h}\}.

and we denote by 𝐔h0:=(Uh0)d\boldsymbol{\mathrm{U}}_{h}^{0}:=(U_{h}^{0})^{d} the vector-valued piecewise constant space. Given a function 𝒈h𝐔h0\boldsymbol{g}_{h}\in\boldsymbol{\mathrm{U}}_{h}^{0} and an integer m1m\geq 1, for the element K𝒯hK\in\mathcal{T}_{h}, we consider the following local constrained least squares problem:

(7) 𝒑S(K)=argmin𝒒m(S(K))d𝒙I(K)𝒒(𝒙)𝒈h(𝒙)l22,s.t. 𝒒(𝒙K)=𝒈h(𝒙K),\boldsymbol{p}_{S(K)}=\mathop{\arg\min}_{\boldsymbol{q}\in\mathbb{P}_{m}(S(K))^{d}}\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{q}(\boldsymbol{x})-\boldsymbol{g}_{h}(\boldsymbol{x})\|_{l^{2}}^{2},\quad\text{s.t. }\boldsymbol{q}(\boldsymbol{x}_{K})=\boldsymbol{g}_{h}(\boldsymbol{x}_{K}),

where I(K):={𝒙K~|K~S(K)𝒯h}I(K):=\{\boldsymbol{x}_{\widetilde{K}}\ |\ \widetilde{K}\in S(K)\cap\mathcal{T}_{h}\} denotes the set of collocation points located in S(K)S(K), and l2\|\cdot\|_{l^{2}} denotes the discrete l2l^{2} norm for vectors. We make the following geometrical assumption on the location of collocation points [11, 10]:

Assumption 1.

For any element patch S(K)S(K) and any polynomial pm(S(K))p\in\mathbb{P}_{m}(S(K)), p|I(K)=0p|_{I(K)}=0 implies p|S(K)=0p|_{S(K)}=0.

The Assumption 1 excludes the case that the points in I(K)I(K) are on an algebraic curve of degree mm and requires cardinality #I(K)dim(m())\#I(K)\geq\mathrm{dim}(\mathbb{P}_{m}(\cdot)). Under this assumption, the fitting problem (7) admits a unique solution. By solving (7) and considering the restriction (𝒑S(K))|K(\boldsymbol{p}_{S(K)})|_{K}, we actually gain a local polynomial defined on KK. Since 𝒑S(K)\boldsymbol{p}_{S(K)} is sought in the least squares sense, we note that 𝒑S(K)\boldsymbol{p}_{S(K)} depends linearly on the given function 𝒈h\boldsymbol{g}_{h}. This property inspires us to define a linear operator from the piecewise constant space 𝐔h0\boldsymbol{\mathrm{U}}_{h}^{0} to a piecewise polynomial space with respect to 𝒯h\mathcal{T}_{h}.

We define a local reconstruction operator K\mathcal{R}_{K} for elements in 𝒯h\mathcal{T}_{h},

K:𝐔h0(m(K))d,𝒈h(𝒑S(K))|K,K𝒯h.\begin{aligned} \mathcal{R}_{K}:\boldsymbol{\mathrm{U}}_{h}^{0}&\rightarrow(\mathbb{P}_{m}(K))^{d},\\ \boldsymbol{g}_{h}&\rightarrow(\boldsymbol{p}_{S(K)})|_{K},\\ \end{aligned}\quad\forall K\in\mathcal{T}_{h}.

Further, the linear reconstruction operator \mathcal{R} for 𝒯h\mathcal{T}_{h} is defined in a piecewise manner as

:𝐔h0𝐔hm,𝒈h𝒈h,(𝒈h)|K:=K𝒈h,K𝒯h,\begin{aligned} \mathcal{R}:\boldsymbol{\mathrm{U}}_{h}^{0}&\rightarrow\boldsymbol{\mathrm{U}}_{h}^{m},\\ \boldsymbol{g}_{h}&\rightarrow\mathcal{R}\boldsymbol{g}_{h},\\ \end{aligned}\quad(\mathcal{R}\boldsymbol{g}_{h})|_{K}:=\mathcal{R}_{K}\boldsymbol{g}_{h},\quad\forall K\in\mathcal{T}_{h},

where 𝐔hm\boldsymbol{\mathrm{U}}_{h}^{m} is the image space of the operator \mathcal{R}. Clearly, by the operator \mathcal{R}, any piecewise constant function 𝒈h\boldsymbol{g}_{h} will be mapped into a piecewise polynomial function 𝒈h𝐔hm\mathcal{R}\boldsymbol{g}_{h}\in\boldsymbol{\mathrm{U}}_{h}^{m}. Then, we present more details to the space 𝐔hm\boldsymbol{\mathrm{U}}_{h}^{m}, which are piecewise polynomial spaces for the partition 𝒯h\mathcal{T}_{h}. For any element KK, we pick up a function 𝒘K,j𝐔h0\boldsymbol{w}_{K,j}\in\boldsymbol{\mathrm{U}}_{h}^{0} such that

𝒘K,j(𝒙)={𝒆j,𝒙K,𝟎,otherwise,1jd,\boldsymbol{w}_{K,j}(\boldsymbol{x})=\begin{cases}\boldsymbol{e}_{j},&\boldsymbol{x}\in K,\\ \boldsymbol{0},&\text{otherwise},\\ \end{cases}\quad 1\leq j\leq d,

where 𝒆j\boldsymbol{e}_{j} is the unit vector in d\mathbb{R}^{d} whose jj-th entry is 11. We let 𝝀K,j:=𝒘K,j(K𝒯hjd)\boldsymbol{\lambda}_{K,j}:=\mathcal{R}\boldsymbol{w}_{K,j}(K\in\mathcal{T}_{h}\leq j\leq d) and we state that the space 𝐔hm\boldsymbol{\mathrm{U}}_{h}^{m} is spanned by {𝝀K,j}\{\boldsymbol{\lambda}_{K,j}\}.

Lemma 3.

the functions {𝛌K,j}(K𝒯h,1jd)\{\boldsymbol{\lambda}_{K,j}\}(K\in\mathcal{T}_{h},1\leq j\leq d) are linearly independent and the space 𝐔hm=span({𝛌K,j})\boldsymbol{\mathrm{U}}_{h}^{m}=\text{span}(\{\boldsymbol{\lambda}_{K,j}\}).

Proof.

For any K𝒯hK\in\mathcal{T}_{h}, assume there exists a group of coefficients {aK,j}\{a_{K,j}\} such that

(8) K𝒯hj=1daK,j𝝀K,j(𝒙)=0,𝒙d.\sum_{K\in\mathcal{T}_{h}}\sum_{j=1}^{d}a_{K,j}\boldsymbol{\lambda}_{K,j}(\boldsymbol{x})=0,\quad\forall\boldsymbol{x}\in\mathbb{R}^{d}.

From the constraint to the problem (7), we have that

𝝀K,j(𝒙K)={𝒆j,K=K,𝟎,otherwise.\boldsymbol{\lambda}_{K,j}(\boldsymbol{x}_{K^{\prime}})=\begin{cases}\boldsymbol{e}_{j},&K^{\prime}=K,\\ \boldsymbol{0},&\text{otherwise}.\\ \end{cases}

We choose 𝒙=𝒙K\boldsymbol{x}=\boldsymbol{x}_{K} for all K𝒯hK\in\mathcal{T}_{h} in (8), and it can be seen that aK,j=0a_{K,j}=0. Thus, the functions {𝝀K,j}(K𝒯h,1jd)\{\boldsymbol{\lambda}_{K,j}\}(K\in\mathcal{T}_{h},1\leq j\leq d) are linearly independent. Obviously, there holds 𝝀K,j𝐔hm\boldsymbol{\lambda}_{K,j}\in\boldsymbol{\mathrm{U}}_{h}^{m}, and we note that the dimension of {𝝀K,j}\{\boldsymbol{\lambda}_{K,j}\} is d#𝒯hd\#\mathcal{T}_{h}. By the linear property of \mathcal{R}, we have that the space 𝐔hm\boldsymbol{\mathrm{U}}_{h}^{m} is spanned by {𝝀K,j}\{\boldsymbol{\lambda}_{K,j}\}. This completes the proof. ∎

Lemma 3 claims that the basis functions of 𝐔hm\boldsymbol{\mathrm{U}}_{h}^{m} are indeed the group of functions {𝝀K,j}\{\boldsymbol{\lambda}_{K,j}\}, and for any function 𝒈h𝐔h0\boldsymbol{g}_{h}\in\boldsymbol{\mathrm{U}}_{h}^{0}, one can explicitly write 𝒈h\mathcal{R}\boldsymbol{g}_{h} as

(9) 𝒈h=K𝒯hj=1dgh,j(𝒙K)𝝀K,j.\mathcal{R}\boldsymbol{g}_{h}=\sum_{K\in\mathcal{T}_{h}}\sum_{j=1}^{d}g_{h,j}(\boldsymbol{x}_{K})\boldsymbol{\lambda}_{K,j}.

From the problem (7), the basis function 𝝀K,j\boldsymbol{\lambda}_{K,j} vanishes on the element KK^{\prime} that KS(K)K\not\in S(K^{\prime}). This fact indicates 𝝀K,j\boldsymbol{\lambda}_{K,j} has a finite support set that supp(𝝀K,j)={K𝒯h|KS(K)}\text{supp}(\boldsymbol{\lambda}_{K,j})=\{K^{\prime}\in\mathcal{T}_{h}\ |\ K\in S(K^{\prime})\}, and supp(𝝀K,j)\text{supp}(\boldsymbol{\lambda}_{K,j}) is different from the patch S(K)S(K).

We extend the operator \mathcal{R} to act on smooth functions. For any 𝒈Hm+1(Ω)d\boldsymbol{g}\in H^{m+1}(\Omega)^{d}, we define a piecewise constant function 𝒈h𝐔h0\boldsymbol{g}_{h}\in\boldsymbol{\mathrm{U}}_{h}^{0} as

𝒈h(𝒙K):=𝒈(𝒙K),K𝒯h,\boldsymbol{g}_{h}(\boldsymbol{x}_{K}):=\boldsymbol{g}(\boldsymbol{x}_{K}),\quad\forall K\in\mathcal{T}_{h},

and we directly define 𝒈:=𝒈h\mathcal{R}\boldsymbol{g}:=\mathcal{R}\boldsymbol{g}_{h}. In this way, any smooth function in Hm+1(Ω)dH^{m+1}(\Omega)^{d} is mapped into a piecewise polynomial function with respect to 𝒯h\mathcal{T}_{h} by the operator \mathcal{R}. For any 𝒈Hm+1(Ω)d\boldsymbol{g}\in H^{m+1}(\Omega)^{d}, 𝒈\mathcal{R}\boldsymbol{g} can also be written as (9).

Next, we will focus on the approximation property of the operator \mathcal{R}. We first define a constant Λm,K\Lambda_{m,K} for every element patch,

Λm,K2:=maxpm(S(K))pL2hd𝒙I(K)p(𝒙)2.\Lambda^{2}_{m,K}:=\max_{p\in\mathbb{P}_{m}(S(K))}\frac{\|p\|_{L^{2}}}{h^{d}\sum_{\boldsymbol{x}\in I(K)}p(\boldsymbol{x})^{2}}.

Assumption 1 as well as the norm equivalence in finite dimensional spaces actually ensures Λm,K<\Lambda_{m,K}<\infty.

Lemma 4.

For any element K𝒯hK\in\mathcal{T}_{h}, there holds

(10) K𝒈L2(K)(1+2Λm,K#I(K))hd/2max𝒙I(K)|𝒈|,𝒈Hm+1(Ω)d.\|\mathcal{R}_{K}\boldsymbol{g}\|_{L^{2}(K)}\leq(1+2\Lambda_{m,K}\sqrt{\#I(K)})h^{d/2}\max_{\boldsymbol{x}\in I(K)}|\boldsymbol{g}|,\quad\forall\boldsymbol{g}\in H^{m+1}(\Omega)^{d}.
Proof.

We define a polynomial space

~m(S(K))d:={𝒗m(S(K))d|𝒗(𝒙K)=𝒈(𝒙K)}.\widetilde{\mathbb{P}}_{m}(S(K))^{d}:=\left\{\boldsymbol{v}\in\mathbb{P}_{m}(S(K))^{d}\ |\ \boldsymbol{v}(\boldsymbol{x}_{K})=\boldsymbol{g}(\boldsymbol{x}_{K})\right\}.

Clearly, any polynomial in ~m(S(K))d\widetilde{\mathbb{P}}_{m}(S(K))^{d} satisfies the constraint to (7). Since 𝒑:=𝒑S(K)\boldsymbol{p}:=\boldsymbol{p}_{S(K)} is the solution to (7), for any ε\varepsilon\in\mathbb{R} and any 𝒒~m(S(K))d\boldsymbol{q}\in\widetilde{\mathbb{P}}_{m}(S(K))^{d}, we have that 𝒑+ε(𝒒𝒈(𝒙K))~m(S(K))d\boldsymbol{p}+\varepsilon(\boldsymbol{q}-\boldsymbol{g}(\boldsymbol{x}_{K}))\in\widetilde{\mathbb{P}}_{m}(S(K))^{d} and

𝒙I(K)𝒑(𝒙)+ε(𝒒(𝒙)𝒈(𝒙K))𝒈(𝒙)l22𝒙I(K)𝒑(𝒙)𝒈(𝒙)l22.\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{p}(\boldsymbol{x})+\varepsilon(\boldsymbol{q}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K}))-\boldsymbol{g}(\boldsymbol{x})\|_{l^{2}}^{2}\geq\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x})\|_{l^{2}}^{2}.

Because ε\varepsilon is arbitrary, the above inequality implies

𝒙I(K)(𝒑(𝒙)𝒈(𝒙))(𝒒(𝒙)𝒈(𝒙K))=0.\sum_{\boldsymbol{x}\in I(K)}(\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}))\cdot(\boldsymbol{q}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K}))=0.

for any 𝒒~m(S(K))d\boldsymbol{q}\in\widetilde{\mathbb{P}}_{m}(S(K))^{d}. By letting 𝒒=𝒑\boldsymbol{q}=\boldsymbol{p} and applying Cauchy-Schwarz inequality,

0\displaystyle 0 =𝒙I(K)(𝒈(𝒙)𝒈(𝒙K)+𝒈(𝒙K)𝒑(𝒙))(𝒑(𝒙)𝒈(𝒙K))\displaystyle=\sum_{\boldsymbol{x}\in I(K)}(\boldsymbol{g}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})+\boldsymbol{g}(\boldsymbol{x}_{K})-\boldsymbol{p}(\boldsymbol{x}))\cdot(\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K}))
=𝒙I(K)(𝒑(𝒙)𝒈(𝒙K)l22+(𝒑(𝒙)𝒈(𝒙K))(𝒈(𝒙)𝒈(𝒙K)))\displaystyle=\sum_{\boldsymbol{x}\in I(K)}\left(-\|\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})\|^{2}_{l^{2}}+(\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K}))\cdot(\boldsymbol{g}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K}))\right)
12𝒙I(K)𝒑(𝒙)𝒈(𝒙K)l22+12𝒙I(K)𝒈(𝒙)𝒈(𝒙K)l22,\displaystyle\leq-\frac{1}{2}\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})\|^{2}_{l^{2}}+\frac{1}{2}\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{g}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})\|^{2}_{l^{2}},

and we obtain

(11) 𝒙I(K)𝒑(𝒙)𝒈(𝒙K)l22𝒙I(K)𝒈(𝒙)𝒈(𝒙K)l22.\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})\|_{l^{2}}^{2}\leq\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{g}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})\|_{l^{2}}^{2}.

Combining the definition of the constantΛm,K\Lambda_{m,K}, we get

𝒑𝒈(𝒙K)L2(K)\displaystyle\|\boldsymbol{p}-\boldsymbol{g}(\boldsymbol{x}_{K})\|_{L^{2}(K)} Λm,K2hd𝒙I(K)𝒑(𝒙)𝒈(𝒙K)l22Λm,K2hd𝒙I(K)𝒈(𝒙)𝒈(𝒙K)l22\displaystyle\leq\Lambda^{2}_{m,K}h^{d}\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})\|_{l^{2}}^{2}\leq\Lambda^{2}_{m,K}h^{d}\sum_{\boldsymbol{x}\in I(K)}\|\boldsymbol{g}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x}_{K})\|_{l^{2}}^{2}
4dΛm,K2hd#I(K)max𝒙I(K)|𝒈|2,\displaystyle\leq 4d\Lambda^{2}_{m,K}h^{d}\#I(K)\max_{\boldsymbol{x}\in I(K)}|\boldsymbol{g}|^{2},

hence

𝒑L2(K)\displaystyle\|\boldsymbol{p}\|_{L^{2}(K)} 𝒑𝒈(𝒙K)L2(K)+hd/2|𝒈(𝒙K)|\displaystyle\leq\|\boldsymbol{p}-\boldsymbol{g}(\boldsymbol{x}_{K})\|_{L^{2}(K)}+h^{d/2}|\boldsymbol{g}(\boldsymbol{x}_{K})| (1+2Λm,K#I(K))hd/2max𝒙I(K)|𝒈|\displaystyle\leq(1+2\Lambda_{m,K}\sqrt{\#I(K)})h^{d/2}\max_{\boldsymbol{x}\in I(K)}|\boldsymbol{g}|

and completes the proof. ∎

Assumption 2.

For every element patch S(K)(K𝒯h)S(K)(K\in\mathcal{T}_{h}), there exist constants RR and rr which are independent of KK such that BrS(K)BRB_{r}\subset S(K)\subset B_{R}, and S(K)S(K) is star-shaped with respect to BrB_{r}, where BρB_{\rho} is a disk with the radius ρ\rho.

From the stability result (10), we can prove the approximation results.

Lemma 5.

For any K𝒯hK\in\mathcal{T}_{h}, there exists a constant CC such that

(12) 𝒈𝒈Hq(K)\displaystyle\|\boldsymbol{g}-\mathcal{R}\boldsymbol{g}\|_{H^{q}(K)} CΛmhKm+1q𝒈Hm+1(S(K)),0qm\displaystyle\leq C\Lambda_{m}h_{K}^{m+1-q}\|\boldsymbol{g}\|_{H^{m+1}(S(K))},\quad 0\leq q\leq m

for any 𝐠Hm+1(Ω)d\boldsymbol{g}\in H^{m+1}(\Omega)^{d}, where we set

(13) Λm:=maxK𝒯h(1+Λm,K#I(K)).\Lambda_{m}:=\max_{K\in\mathcal{T}_{h}}\left(1+\Lambda_{m,K}\sqrt{\#I(K)}\right).
Proof.

According to Assumption2, there exists a polynomial 𝒑m(S(K))d\boldsymbol{p}\in\mathbb{P}_{m}(S(K))^{d}, such that

𝒈𝒑Hq(S(K))\displaystyle\|\boldsymbol{g}-\boldsymbol{p}\|_{H^{q}(S(K))} hKm+1q𝒈Hm+1(S(K))\displaystyle\leq h_{K}^{m+1-q}\|\boldsymbol{g}\|_{H^{m+1}(S(K))}
𝒈𝒑L(S(K))\displaystyle\|\boldsymbol{g}-\boldsymbol{p}\|_{L^{\infty}(S(K))} hKm+1d/2𝒈Hm+1(S(K))\displaystyle\leq h_{K}^{m+1-d/2}\|\boldsymbol{g}\|_{H^{m+1}(S(K))}

Combining the stablility result of the reconstruction operator, we have

𝒈𝒈L2(K)\displaystyle\|\boldsymbol{g}-\mathcal{R}\boldsymbol{g}\|_{L^{2}(K)} 𝒈𝒑L2(K)+𝒑𝒈L2(K)\displaystyle\leq\|\boldsymbol{g}-\boldsymbol{p}\|_{L^{2}(K)}+\|\boldsymbol{p}-\mathcal{R}\boldsymbol{g}\|_{L^{2}(K)}
𝒈𝒑L2(K)+(1+2Λm,K#I(K))hd/2max𝒙I(K)|𝒑(𝒙)𝒈(𝒙)|\displaystyle\leq\|\boldsymbol{g}-\boldsymbol{p}\|_{L^{2}(K)}+(1+2\Lambda_{m,K}\sqrt{\#I(K)})h^{d/2}\max_{\boldsymbol{x}\in I(K)}|\boldsymbol{p}(\boldsymbol{x})-\boldsymbol{g}(\boldsymbol{x})|

By the arbitraryness of 𝒑\boldsymbol{p},

𝒈𝒈L2(K)\displaystyle\|\boldsymbol{g}-\mathcal{R}\boldsymbol{g}\|_{L^{2}(K)} Chm+1Λm𝒈Hm+1(S(K))\displaystyle\leq Ch^{m+1}\Lambda_{m}\|\boldsymbol{g}\|_{H^{m+1}(S(K))}

the case q>0q>0 follows from the inverse inequality

𝒈𝒈Hq(K)\displaystyle\|\boldsymbol{g}-\mathcal{R}\boldsymbol{g}\|_{H^{q}(K)} 𝒈𝒑Hq(K)+𝒑𝒈Hq(K)\displaystyle\leq\|\boldsymbol{g}-\boldsymbol{p}\|_{H^{q}(K)}+\|\boldsymbol{p}-\mathcal{R}\boldsymbol{g}\|_{H^{q}(K)}
Chm+1q𝒈Hm+1(K)+Chq𝒑𝒈L2(K)\displaystyle\leq Ch^{m+1-q}\|\boldsymbol{g}\|_{H^{m+1}(K)}+Ch^{-q}\|\boldsymbol{p}-\mathcal{R}\boldsymbol{g}\|_{L^{2}(K)}
CΛmhm+1q𝒈Hm+1(S(K))\displaystyle\leq C\Lambda_{m}h^{m+1-q}\|\boldsymbol{g}\|_{H^{m+1}(S(K))}

From (12), it can be seen that the operator \mathcal{R} has an approximation error of degree O(hm+1q)O(h^{m+1-q}) under the case that Λm\Lambda_{m} admits an upper bound independent of the mesh size hh. However, it is not trivial to bound the constant Λm\Lambda_{m}, see Remark 1. We can prove that under some mild conditions on the element patch, the constant admits a uniform upper bound.

Remark 1.

There is a constant Λ(m,S(K))\Lambda(m,S(K)) defined in [10] as

Λ(m,S(K)):=maxpm(S(K))max𝒙S(K)|p(𝒙)|max𝒙I(K)|p(𝒙)|.\Lambda(m,S(K)):=\max_{p\in\mathbb{P}_{m}(S(K))}\frac{\max_{\boldsymbol{x}\in S(K)}|p(\boldsymbol{x})|}{\max_{\boldsymbol{x}\in I(K)}|p(\boldsymbol{x})|}.

It can be shown that Λm,KΛ(m,S(K))\Lambda_{m,K}\leq\Lambda(m,S(K)). Hence when the constant Λ(m,S(K))\Lambda(m,S(K)) is uniformly bounded, so is Λm,K\Lambda_{m,K}. For the patch S(K)S(K), consider the special case that the corresponding collocation points set I(K)I(K) has that #I(K)=dim(m())\#I(K)=\mathrm{dim}(\mathbb{P}_{m}(\cdot)), then the constant Λ(m,S(K))\Lambda(m,S(K)) is equal to the Lebesgue constant [19] and the solution to the least squares problem (7) is just the Lagrange interpolation polynomial, however, in this case the constant Λ(m,S(K))\Lambda(m,S(K)) may depend on hh and grow very fast, which will further damage the convergence. To ensure the stability of the reconstruction operator, we are acquired to choose a sufficiently large #S\#S so that the constant Λ(m,S(K))\Lambda(m,S(K)) admits a uniform upper bound, thus. We refer to [10, Lemma 6] and [11, Lemma 3.4] for the details of this statement. The wide element patch will bring more computational cost for filling the stiffness matrix and increase the width of the banded structure. This is just the price we have to pay for the uniform bound of Λ(m,S(K))\Lambda(m,S(K)).

4. Approximation to Fourth-order Curl Problem

We define the approximation problem to solve the problem (1) based on the space 𝐔hm\boldsymbol{\mathrm{U}}_{h}^{m} constructed in the previous section: Seek 𝒖𝒉𝐔hm\boldsymbol{u_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m} (m2m\geq 2) such that

(14) Bh(𝒖𝒉,𝒗𝒉)=lh(𝒗𝒉),𝒗𝒉𝐔hm,B_{h}(\boldsymbol{u_{h}},\boldsymbol{v_{h}})=l_{h}(\boldsymbol{v_{h}}),\quad\forall\boldsymbol{v_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m},

where we define the bilinear form Bh(,)B_{h}(\cdot,\cdot) on Uh:=𝐔hm+H(curl2,Ω){𝒗:curlj𝒗1/2+σ,Ω<,0j3}U_{h}:=\boldsymbol{\mathrm{U}}_{h}^{m}+H(\mathrm{curl}^{2},\Omega)\cap\{\boldsymbol{v}:\|\mathrm{curl}^{j}\boldsymbol{v}\|_{1/2+\sigma,\Omega}<\infty,0\leq j\leq 3\}, as in [4].

Bh(𝒖,𝒗)\displaystyle B_{h}(\boldsymbol{u},\boldsymbol{v}) :=K𝒯hKcurl2𝒖curl2𝒗d𝒙+K𝒯hK𝒖𝒗d𝒙\displaystyle:=\sum_{K\in\mathcal{T}_{h}}\int_{K}\mathrm{curl}^{2}\boldsymbol{u}\cdot\mathrm{curl}^{2}\boldsymbol{v}\ \mathrm{d}\boldsymbol{\boldsymbol{x}}+\sum_{K\in\mathcal{T}_{h}}\int_{K}\boldsymbol{u}\cdot\boldsymbol{v}\ \mathrm{d}\boldsymbol{\boldsymbol{x}}
+ehe([[𝒖]]{curl3𝒗}+[[curl𝒖]]{curl2𝒗})d𝒔+ehe([[𝒗]]{curl3𝒖}+[[curl𝒗]]{curl2𝒖})d𝒔\displaystyle+\sum_{e\in\mathcal{E}_{h}}\int_{e}\left([[\boldsymbol{u}]]\cdot\{\mathrm{curl}^{3}\boldsymbol{v}\}+[[\mathrm{curl}\boldsymbol{u}]]\cdot\{\mathrm{curl}^{2}\boldsymbol{v}\}\right)\mathrm{d}\boldsymbol{\boldsymbol{s}}+\sum_{e\in\mathcal{E}_{h}}\int_{e}\left([[\boldsymbol{v}]]\cdot\{\mathrm{curl}^{3}\boldsymbol{u}\}+[[\mathrm{curl}\boldsymbol{v}]]\cdot\{\mathrm{curl}^{2}\boldsymbol{u}\}\right)\mathrm{d}\boldsymbol{\boldsymbol{s}}
+ehe(μ1[[𝒖]][[𝒗]]+μ2[[curl𝒖]][[curl𝒗]])d𝒔\displaystyle+\sum_{e\in\mathcal{E}_{h}}\int_{e}\left(\mu_{1}[[\boldsymbol{u}]]\cdot[[\boldsymbol{v}]]+\mu_{2}[[\mathrm{curl}\boldsymbol{u}]]\cdot[[\mathrm{curl}\boldsymbol{v}]]\right)\mathrm{d}\boldsymbol{\boldsymbol{s}}

The parameter μ1\mu_{1} and μ2\mu_{2} are positive penalties which are set by

μ1|e=ηhe3,μ2|e=ηhe. for eh.\displaystyle\mu_{1}|_{e}=\frac{\eta}{h_{e}^{3}},\quad\mu_{2}|_{e}=\frac{\eta}{h_{e}}.\quad\text{ for }e\in\mathcal{E}_{h}.

The linear form lh()l_{h}(\cdot) is defined for 𝒗Uh\boldsymbol{v}\in U_{h},

lh()\displaystyle l_{h}(\cdot) :=K𝒯hK𝒇𝒗d𝒙\displaystyle:=\sum_{K\in\mathcal{T}_{h}}\int_{K}\boldsymbol{f}\cdot\boldsymbol{v}\mathrm{d}\boldsymbol{\boldsymbol{x}} +ehbe(g1(curl3𝒗𝒉+μ1𝒗𝒉×𝐧)+g2(curl2𝒗𝒉+μ2curl𝒗𝒉×𝐧))d𝒔.\displaystyle+\sum_{e\in\mathcal{E}_{h}^{b}}\int_{e}\left(g_{1}\cdot(\mathrm{curl}^{3}\boldsymbol{v_{h}}+\mu_{1}\boldsymbol{v_{h}}\times\boldsymbol{\mathrm{n}})+g_{2}\cdot(\mathrm{curl}^{2}\boldsymbol{v_{h}}+\mu_{2}\mathrm{curl}\boldsymbol{v_{h}}\times\boldsymbol{\mathrm{n}})\right)\mathrm{d}\boldsymbol{\boldsymbol{s}}.

We introduce two energy norms for the space UhU_{h},

𝒖DG2\displaystyle\|\boldsymbol{u}\|_{\mathrm{DG}}^{2} :=K𝒯hK|𝒖|2d𝒙+K𝒯hK|curl2𝒖|2d𝒙+ehehK3|[[𝒖]]|2d𝒔+ehehK1|[[curl𝒖]]|2d𝒔,\displaystyle:=\sum_{K\in\mathcal{T}_{h}}\int_{K}|\boldsymbol{u}|^{2}\mathrm{d}\boldsymbol{\boldsymbol{x}}+\sum_{K\in\mathcal{T}_{h}}\int_{K}|\mathrm{curl}^{2}\boldsymbol{u}|^{2}\mathrm{d}\boldsymbol{\boldsymbol{x}}+\sum_{e\in\mathcal{E}_{h}}\int_{e}h_{K}^{-3}|[[\boldsymbol{u}]]|^{2}\mathrm{d}\boldsymbol{\boldsymbol{s}}+\sum_{e\in\mathcal{E}_{h}}\int_{e}h_{K}^{-1}|[[\mathrm{curl}\boldsymbol{u}]]|^{2}\mathrm{d}\boldsymbol{\boldsymbol{s}},

and

|𝒖|2\displaystyle|\!|\!|\boldsymbol{u}|\!|\!|^{2} :=𝒖DG2+ehehe3|{curl3𝒖}|2d𝒔+ehehe|{curl2𝒖}|2d𝒔.\displaystyle:=\|\boldsymbol{u}\|_{\mathrm{DG}}^{2}+\sum_{e\in\mathcal{E}_{h}}\int_{e}h_{e}^{3}|\{\mathrm{curl}^{3}\boldsymbol{u}\}|^{2}\mathrm{d}\boldsymbol{\boldsymbol{s}}+\sum_{e\in\mathcal{E}_{h}}\int_{e}h_{e}|\{\mathrm{curl}^{2}\boldsymbol{u}\}|^{2}\mathrm{d}\boldsymbol{\boldsymbol{s}}.

We claim that the two energy norms are equivalent over the space 𝐔hm\boldsymbol{\mathrm{U}}_{h}^{m}.

Lemma 6.

For any 𝐮𝐡𝐔hm\boldsymbol{u_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m}, there exists a constant CC, such that

(15) 𝒖𝒉DG|𝒖𝒉|C𝒖𝒉DG.\|\boldsymbol{u_{h}}\|_{\mathrm{DG}}\leq|\!|\!|\boldsymbol{u_{h}}|\!|\!|\leq C\|\boldsymbol{u_{h}}\|_{\mathrm{DG}}.
Proof.

We only need to prove |𝒖𝒉|C𝒖𝒉DG|\!|\!|\boldsymbol{u_{h}}|\!|\!|\leq C\|\boldsymbol{u_{h}}\|_{\mathrm{DG}}. For ehie\in\mathcal{E}_{h}^{i}, we denote the two neighbor elements of ee by K+K^{+} and KK^{-}. We have

he3/2{curl3𝒖𝒉}L2(e)C(he3/2curl3𝒖𝒉L2(eK+)+he3/2curl3𝒖𝒉L2(eK))\|h_{e}^{3/2}\{\mathrm{curl}^{3}\boldsymbol{u_{h}}\}\|_{L^{2}(e)}\leq C\left(\|h_{e}^{3/2}\mathrm{curl}^{3}\boldsymbol{u_{h}}\|_{L^{2}(e\cap\partial K^{+})}+\|h_{e}^{3/2}\mathrm{curl}^{3}\boldsymbol{u_{h}}\|_{L^{2}(e\cap\partial K^{-})}\right)

By the trace inequalities (2), and the inverse inequality (3), we obtain that

he3/2curl3𝒖𝒉L2(eK±)Ccurl2𝒖𝒉L2(K±)\|h_{e}^{3/2}\mathrm{curl}^{3}\boldsymbol{u_{h}}\|_{L^{2}(e\cap\partial K^{\pm})}\leq C\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|_{L^{2}(K^{\pm})}

For ehbe\in\mathcal{E}_{h}^{b}, let eKe\subset K. Similarly, we have

he3/2curl3𝒖𝒉L2(eK)Ccurl2𝒖𝒉L2(K)\|h_{e}^{3/2}\mathrm{curl}^{3}\boldsymbol{u_{h}}\|_{L^{2}(e\cap\partial K)}\leq C\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|_{L^{2}(K)}

The term he1/2{curl2𝒖𝒉}L2(e)\|h_{e}^{1/2}\{\mathrm{curl}^{2}\boldsymbol{u_{h}}\}\|_{L^{2}(e)} can bounded by the same way. Thus, by summing over all ehe\in\mathcal{E}_{h} , we conclude that

ehhe1/2{curl2𝒖𝒉}L2(e)2+ehhe3/2{curl3𝒖𝒉}L2(e)2CK𝒯hcurl2𝒖𝒉L2(K)2,\displaystyle\sum_{e\in\mathcal{E}_{h}}\|h_{e}^{1/2}\{\mathrm{curl}^{2}\boldsymbol{u_{h}}\}\|_{L^{2}(e)}^{2}+\sum_{e\in\mathcal{E}_{h}}\|h_{e}^{3/2}\{\mathrm{curl}^{3}\boldsymbol{u_{h}}\}\|_{L^{2}(e)}^{2}\leq C\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|_{L^{2}(K)}^{2},

which can immediately leads us to (15) and completes the proof. ∎

Now we are ready to prove the coercivity and the continuity of the bilinear form Bh(,)B_{h}(\cdot,\cdot).

Theorem 1.

Let Bh(,)B_{h}(\cdot,\cdot) be the bilinear form with sufficiently large penalty η\eta. Then there exists a positive constant CC such that

(16) Bh(𝒖𝒉,𝒖𝒉)C|𝒖𝒉|2,𝒖𝒉𝐔hm.B_{h}(\boldsymbol{u_{h}},\boldsymbol{u_{h}})\geq C|\!|\!|\boldsymbol{u_{h}}|\!|\!|^{2},\quad\forall\boldsymbol{u_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m}.
Proof.

From Lemma 15, we only need to establish the coercivity over the norm DG\|\cdot\|_{\mathrm{DG}}. For the face ehie\in\mathcal{E}_{h}^{i}, let ee be shared by the neighbor elements KK^{-} and K+K^{+}. We apply the Cauchy-Schwarz inequality,

e2[[𝒖𝒉]]\displaystyle-\int_{e}2[[\boldsymbol{u_{h}}]]\cdot {curl3𝒖𝒉}d𝒔1ϵhe3/2[[𝒖𝒉]]L2(e)2ϵhe3/2{curl3𝒖𝒉}L2(e)2\displaystyle\{\mathrm{curl}^{3}\boldsymbol{u_{h}}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}\geq-\frac{1}{\epsilon}\|h_{e}^{-3/2}[[\boldsymbol{u_{h}}]]\|^{2}_{L^{2}(e)}-\epsilon\|h_{e}^{3/2}\{\mathrm{curl}^{3}\boldsymbol{u_{h}}\}\|^{2}_{L^{2}(e)}
1ϵhe3/2[[𝒖𝒉]]L2(e)2ϵ2he3/2curl3𝒖𝒉L2(eK)2ϵ2he3/2curl3𝒖𝒉L2(eK+)2,\displaystyle\geq-\frac{1}{\epsilon}\|h_{e}^{-3/2}[[\boldsymbol{u_{h}}]]\|^{2}_{L^{2}(e)}-\frac{\epsilon}{2}\|h_{e}^{3/2}\mathrm{curl}^{3}\boldsymbol{u_{h}}\|^{2}_{L^{2}(e\cap\partial K^{-})}-\frac{\epsilon}{2}\|h_{e}^{3/2}\mathrm{curl}^{3}\boldsymbol{u_{h}}\|^{2}_{L^{2}(e\cap\partial K^{+})},

for any ϵ>0\epsilon>0. From the trace inequalities (2), and the inverse inequality (3), we deduce that

he3/2curl3𝒖𝒉L2(eiK±)Ccurl2𝒖𝒉L2(K±)\|h_{e}^{3/2}\mathrm{curl}^{3}\boldsymbol{u_{h}}\|_{L^{2}(e^{i}\cap\partial K^{\pm})}\leq C\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|_{L^{2}(K^{\pm})}

Thus, we have

(17) ehie0e12[[𝒖𝒉]]{curl3𝒖𝒉}d𝒔ehi1ϵhe3/2[[𝒖𝒉]]L2(e0e1)2CϵK𝒯hcurl2𝒖𝒉L2(K0K1)2.-\sum_{e\in\mathcal{E}_{h}^{i}}\int_{e^{0}\cup e^{1}}2[[\boldsymbol{u_{h}}]]\cdot\{\mathrm{curl}^{3}\boldsymbol{u_{h}}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}\geq-\sum_{e\in\mathcal{E}_{h}^{i}}\frac{1}{\epsilon}\|h_{e}^{-3/2}[[\boldsymbol{u_{h}}]]\|^{2}_{L^{2}(e^{0}\cup e^{1})}-C\epsilon\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|^{2}_{L^{2}(K^{0}\cup K^{1})}.

For the face ehbe\in\mathcal{E}_{h}^{b}, we can similarly derive that

(18) ehbe0e12[[𝒖𝒉]]{curl3𝒖𝒉}d𝒔ehb1ϵhe3/2[[𝒖𝒉]]L2(e)2CϵK𝒯hcurl2𝒖𝒉L2(K)2.-\sum_{e\in\mathcal{E}_{h}^{b}}\int_{e^{0}\cup e^{1}}2[[\boldsymbol{u_{h}}]]\cdot\{\mathrm{curl}^{3}\boldsymbol{u_{h}}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}\geq-\sum_{e\in\mathcal{E}_{h}^{b}}\frac{1}{\epsilon}\|h_{e}^{-3/2}[[\boldsymbol{u_{h}}]]\|^{2}_{L^{2}(e)}-C\epsilon\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|^{2}_{L^{2}(K)}.

By employing the same method to the term e0e12[[curl𝒖𝒉]]{curl2𝒖𝒉}d𝒔\int_{e^{0}\cup e^{1}}2[[\mathrm{curl}\boldsymbol{u_{h}}]]\{\mathrm{curl}^{2}\boldsymbol{u_{h}}\}\mathrm{d}\boldsymbol{\boldsymbol{s}} , we can obtain that

(19) ehe2[[curl𝒖𝒉]]{curl2𝒖𝒉}d𝒔1ϵehhe1/2[[curl𝒖𝒉]]L2(e)2CϵK𝒯hcurl2𝒖𝒉L2(K)2,-\sum_{e\in\mathcal{E}_{h}}\int_{e}2[[\mathrm{curl}\boldsymbol{u_{h}}]]\{\mathrm{curl}^{2}\boldsymbol{u_{h}}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}\geq-\frac{1}{\epsilon}\sum_{e\in\mathcal{E}_{h}}\|h_{e}^{-1/2}[[\mathrm{curl}\boldsymbol{u_{h}}]]\|^{2}_{L^{2}(e)}-C\epsilon\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|^{2}_{L^{2}(K)},

Combining the inequalities (17), (18), (19), we conclude that there exists a constant CC such that

Bh(𝒖𝒉,𝒖𝒉)\displaystyle B_{h}(\boldsymbol{u_{h}},\boldsymbol{u_{h}}) (1Cϵ)K𝒯hcurl2𝒖𝒉L2(K)2+K𝒯h𝒖𝒉L2(K)2\displaystyle\geq(1-C\epsilon)\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{u_{h}}\|^{2}_{L^{2}(K)}+\sum_{K\in\mathcal{T}_{h}}\|\boldsymbol{u_{h}}\|^{2}_{L^{2}(K)}
+(η1ϵ)eh(he1/2[[curl𝒖𝒉]]L2(e)2+he3/2[[𝒖𝒉]]L2(e)2)\displaystyle+(\eta-\frac{1}{\epsilon})\sum_{e\in\mathcal{E}_{h}}(\|h_{e}^{-1/2}[[\mathrm{curl}\boldsymbol{u_{h}}]]\|^{2}_{L^{2}(e)}+\|h_{e}^{-3/2}[[\boldsymbol{u_{h}}]]\|^{2}_{L^{2}(e)})

for any ϵ>0\epsilon>0. We can let ϵ=1/(2C)\epsilon=1/(2C) and select a sufficiently large η\eta to ensure Bh(𝒖𝒉,𝒖𝒉)C𝒖𝒉DG2B_{h}(\boldsymbol{u_{h}},\boldsymbol{u_{h}})\geq C\|\boldsymbol{u_{h}}\|_{\mathrm{DG}}^{2}, which completes the proof. ∎

Theorem 2.

There exists a positive constant CC such that

(20) |Bh(𝒖,𝒗)|C|𝒖||𝒗|,𝒖,𝒗Uh.|B_{h}(\boldsymbol{u},\boldsymbol{v})|\leq C|\!|\!|\boldsymbol{u}|\!|\!||\!|\!|\boldsymbol{v}|\!|\!|,\quad\forall\boldsymbol{u},\boldsymbol{v}\in U_{h}.
Proof.

By directly using the Cauchy-Schwarz inequality,

Bh(𝒖,𝒗)C\displaystyle B_{h}(\boldsymbol{u},\boldsymbol{v})\leq C (K𝒯hcurl2𝒖L2(K)2+eh(he3/2[[𝒖]]L2(e)2+he1/2[[curl𝒖]]L2(e)2\displaystyle\left(\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{u}\|_{L^{2}(K)}^{2}+\sum_{e\in\mathcal{E}_{h}}(\|h_{e}^{-3/2}[[\boldsymbol{u}]]\|^{2}_{L^{2}(e)}+\|h_{e}^{-1/2}[[\mathrm{curl}\boldsymbol{u}]]\|^{2}_{L^{2}(e)}\right.
+he3/2{curl3𝒖}L2(e)2+he1/2{curl2𝒖}L2(e)2))1/2\displaystyle\left.+\|h_{e}^{3/2}\{\mathrm{curl}^{3}\boldsymbol{u}\}\|^{2}_{L^{2}(e)}+\|h_{e}^{1/2}\{\mathrm{curl}^{2}\boldsymbol{u}\}\|^{2}_{L^{2}(e)})\right)^{1/2}\cdot
(K𝒯hcurl2𝒗L2(K)2+eh(he3/2[[𝒗]]L2(e)2+he1/2[[curl𝒗]]L2(e)2\displaystyle\left(\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{v}\|_{L^{2}(K)}^{2}+\sum_{e\in\mathcal{E}_{h}}(\|h_{e}^{-3/2}[[\boldsymbol{v}]]\|^{2}_{L^{2}(e)}+\|h_{e}^{-1/2}[[\mathrm{curl}\boldsymbol{v}]]\|^{2}_{L^{2}(e)}\right.
+he3/2{curl3𝒗}L2(e)2+he1/2{curl2𝒗}L2(e)2))1/2\displaystyle\left.+\|h_{e}^{3/2}\{\mathrm{curl}^{3}\boldsymbol{v}\}\|^{2}_{L^{2}(e)}+\|h_{e}^{1/2}\{\mathrm{curl}^{2}\boldsymbol{v}\}\|^{2}_{L^{2}(e)})\right)^{1/2}

which completes the proof. ∎

Now we verify the Galerkin orthogonality.

Lemma 7.

Suppose 𝐮\boldsymbol{u} is the exact solution to the problem (1), and 𝐮𝐡𝐔hm\boldsymbol{u_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m} is the numerical solution to the discrete problem (14), then

(21) Bh(𝒖𝒖𝒉,𝒗𝒉)=0,𝒗𝒉𝐔hm.B_{h}(\boldsymbol{u}-\boldsymbol{u_{h}},\boldsymbol{v_{h}})=0,\quad\forall\boldsymbol{v_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m}.
Proof.

We first have for the two jump terms

[[𝒖]]|e=𝟎,[[curl𝒖]]|e=0,ehi,i=0,1.[[\boldsymbol{u}]]|_{e}=\boldsymbol{0},\quad[[\mathrm{curl}\boldsymbol{u}]]|_{e}=0,\quad\quad\forall e\in\mathcal{E}_{h}^{i},\,\,i=0,1.

Taking the exact solution into Bh(,)B_{h}(\cdot,\cdot), we have that

Bh(u,𝒗𝒉)\displaystyle B_{h}(u,\boldsymbol{v_{h}}) =K𝒯hKcurl2𝒖curl2𝒗𝒉d𝒙+K𝒯hK𝒖𝒗𝒉d𝒙\displaystyle=\sum_{K\in\mathcal{T}_{h}}\int_{K}\mathrm{curl}^{2}\boldsymbol{u}\cdot\mathrm{curl}^{2}\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}+\sum_{K\in\mathcal{T}_{h}}\int_{K}\boldsymbol{u}\cdot\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}
+ehe[[𝒗𝒉]]{curl3𝒖}d𝒔+ehe[[curl𝒗𝒉]]{curl2𝒖}d𝒔\displaystyle+\sum_{e\in\mathcal{E}_{h}}\int_{e}[[\boldsymbol{v_{h}}]]\cdot\{\mathrm{curl}^{3}\boldsymbol{u}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}+\sum_{e\in\mathcal{E}_{h}}\int_{e}[[\mathrm{curl}\boldsymbol{v_{h}}]]\{\mathrm{curl}^{2}\boldsymbol{u}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}
+ehbe(g1(curl3𝒗𝒉+μ1𝒗𝒉×𝐧)+g2(curl2𝒗𝒉+μ2curl𝒗𝒉×𝐧))d𝒔\displaystyle+\sum_{e\in\mathcal{E}_{h}^{b}}\int_{e}\left(g_{1}\cdot(\mathrm{curl}^{3}\boldsymbol{v_{h}}+\mu_{1}\boldsymbol{v_{h}}\times\boldsymbol{\mathrm{n}})+g_{2}\cdot(\mathrm{curl}^{2}\boldsymbol{v_{h}}+\mu_{2}\mathrm{curl}\boldsymbol{v_{h}}\times\boldsymbol{\mathrm{n}})\right)\mathrm{d}\boldsymbol{\boldsymbol{s}}

We multiply the test function 𝒗𝒉\boldsymbol{v_{h}} at both side of equation (1), and apply the integration by parts to get

K𝒯hK\displaystyle\sum_{K\in\mathcal{T}_{h}}\int_{K} 𝒇𝒗𝒉d𝒙=K𝒯hKcurl4𝒖𝒗𝒉d𝒙+K𝒯hK𝒖𝒗𝒉d𝒙\displaystyle\boldsymbol{f}\cdot\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}=\sum_{K\in\mathcal{T}_{h}}\int_{K}\mathrm{curl}^{4}\boldsymbol{u}\cdot\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}+\sum_{K\in\mathcal{T}_{h}}\int_{K}\boldsymbol{u}\cdot\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}
=K𝒯hKcurl2𝒖curl2𝒗𝒉d𝒙+K𝒯hK𝒖𝒗𝒉d𝒙\displaystyle=\sum_{K\in\mathcal{T}_{h}}\int_{K}\mathrm{curl}^{2}\boldsymbol{u}\cdot\mathrm{curl}^{2}\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}+\sum_{K\in\mathcal{T}_{h}}\int_{K}\boldsymbol{u}\cdot\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}
+K𝒯h(Kcurl3𝒖𝒗𝒉×𝐧d𝒔+Kcurl2𝒖(curl𝒗𝒉×𝐧)d𝒔)\displaystyle+\sum_{K\in\mathcal{T}_{h}}\left(\int_{\partial K}\mathrm{curl}^{3}\boldsymbol{u}\cdot\boldsymbol{v_{h}}\times\boldsymbol{\mathrm{n}}\mathrm{d}\boldsymbol{\boldsymbol{s}}+\int_{\partial K}\mathrm{curl}^{2}\boldsymbol{u}\cdot(\mathrm{curl}\boldsymbol{v_{h}}\times\boldsymbol{\mathrm{n}})\mathrm{d}\boldsymbol{\boldsymbol{s}}\right)
=K𝒯hKcurl2𝒖curl2𝒗𝒉d𝒙+K𝒯hK𝒖𝒗𝒉d𝒙\displaystyle=\sum_{K\in\mathcal{T}_{h}}\int_{K}\mathrm{curl}^{2}\boldsymbol{u}\cdot\mathrm{curl}^{2}\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}+\sum_{K\in\mathcal{T}_{h}}\int_{K}\boldsymbol{u}\cdot\boldsymbol{v_{h}}\mathrm{d}\boldsymbol{\boldsymbol{x}}
+ehe[[𝒗𝒉]]{curl3𝒖}d𝒔+ehe[[curl𝒗𝒉]]{curl2𝒖}d𝒔d𝒔\displaystyle+\sum_{e\in\mathcal{E}_{h}}\int_{e}[[\boldsymbol{v_{h}}]]\cdot\{\mathrm{curl}^{3}\boldsymbol{u}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}+\sum_{e\in\mathcal{E}_{h}}\int_{e}[[\mathrm{curl}\boldsymbol{v_{h}}]]\{\mathrm{curl}^{2}\boldsymbol{u}\}\mathrm{d}\boldsymbol{\boldsymbol{s}}\mathrm{d}\boldsymbol{\boldsymbol{s}}

Thus, by simply calculating, we obtain that

Bh(𝒖𝒉,𝒗𝒉)=lh(𝒗𝒉)=Bh(𝒖,𝒗𝒉),B_{h}(\boldsymbol{u_{h}},\boldsymbol{v_{h}})=l_{h}(\boldsymbol{v_{h}})=B_{h}(\boldsymbol{u},\boldsymbol{v_{h}}),

which completes the proof. ∎

Then we establish the interpolation error estimate of the reconstruction operator.

Lemma 8.

For 0hh00\leq h\leq h_{0} and m2m\geq 2, there exists a constant CC such that

(22) |𝒗𝒗|CΛmhm1𝒗Hm+1(Ω),vHs(Ω),s=max(4,m+1).|\!|\!|\boldsymbol{v}-\mathcal{R}\boldsymbol{v}|\!|\!|\leq C\Lambda_{m}h^{m-1}\|\boldsymbol{v}\|_{H^{m+1}(\Omega)},\quad\forall v\in H^{s}(\Omega),\quad s=\max(4,m+1).
Proof.

From Lemma LABEL:le_localapproximation, we can show that

K𝒯hcurl2𝒗curl2(𝒗)L2(K)2\displaystyle\sum_{K\in\mathcal{T}_{h}}\|\mathrm{curl}^{2}\boldsymbol{v}-\mathrm{curl}^{2}(\mathcal{R}\boldsymbol{v})\|^{2}_{L^{2}(K)} K𝒯hCΛm2hK2m2𝒗Hm+1(S(K))2\displaystyle\leq\sum_{K\in\mathcal{T}_{h}}C\Lambda_{m}^{2}h_{K}^{2m-2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(S(K))}
CΛm2h2m2𝒗Hm+1(Ω)2CΛm2h2m2𝒗Hm+1(Ω)2\displaystyle\leq C\Lambda_{m}^{2}h^{2m-2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(\Omega)}\leq C\Lambda_{m}^{2}h^{2m-2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(\Omega)}

,also

K𝒯h𝒗(𝒗)L2(K)2\displaystyle\sum_{K\in\mathcal{T}_{h}}\|\boldsymbol{v}-(\mathcal{R}\boldsymbol{v})\|^{2}_{L^{2}(K)} K𝒯hCΛm2hK2m+2𝒗Hm+1(S(K))2\displaystyle\leq\sum_{K\in\mathcal{T}_{h}}C\Lambda_{m}^{2}h_{K}^{2m+2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(S(K))}
CΛm2h2m+2𝒗Hm+1(Ω)2CΛm2h2m+2𝒗Hm+1(Ω)2,\displaystyle\leq C\Lambda_{m}^{2}h^{2m+2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(\Omega)}\leq C\Lambda_{m}^{2}h^{2m+2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(\Omega)},

By the trace estimate (2) and the mesh regularity,

ehhe1[[curl(𝒗𝒗)]]L2(e)2\displaystyle\sum_{e\in\mathcal{E}_{h}}h_{e}^{-1}\|[[\mathrm{curl}(\boldsymbol{v}-\mathcal{R}\boldsymbol{v})]]\|^{2}_{L^{2}(e)} CK𝒯h(hK2curl(𝒗𝒗)L2(K)2+curl(𝒗𝒗)H1(K)2)\displaystyle\leq C\sum_{K\in\mathcal{T}_{h}}\left(h_{K}^{-2}\|\mathrm{curl}(\boldsymbol{v}-\mathcal{R}\boldsymbol{v})\|^{2}_{L^{2}(K)}+\|\mathrm{curl}(\boldsymbol{v}-\mathcal{R}\boldsymbol{v})\|^{2}_{H^{1}(K)}\right)
CΛm2h2m2𝒗Hm+1(Ω)2.\displaystyle\leq C\Lambda_{m}^{2}h^{2m-2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(\Omega)}.

and

ehhe3[[𝒗𝒗]]L2(e)2\displaystyle\sum_{e\in\mathcal{E}_{h}}h_{e}^{-3}\|[[\boldsymbol{v}-\mathcal{R}\boldsymbol{v}]]\|^{2}_{L^{2}(e)} CK𝒯h(hK4𝒗𝒗L2(K)2+hK2𝒗𝒗H1(K)2)\displaystyle\leq C\sum_{K\in\mathcal{T}_{h}}\left(h_{K}^{-4}\|\boldsymbol{v}-\mathcal{R}\boldsymbol{v}\|^{2}_{L^{2}(K)}+h_{K}^{-2}\|\boldsymbol{v}-\mathcal{R}\boldsymbol{v}\|^{2}_{H^{1}(K)}\right)
CΛm2h2m2𝒗Hm+1(Ω)2.\displaystyle\leq C\Lambda_{m}^{2}h^{2m-2}\|\boldsymbol{v}\|^{2}_{H^{m+1}(\Omega)}.

which completes the proof. ∎

Now we are ready to present the a priori error estimate under |||||||\!|\!|\cdot|\!|\!| within the standard Lax-Milgram framework.

Theorem 3.

Suppose the problem (1) has a solution 𝐮Hs(Ω)\boldsymbol{u}\in H^{s}(\Omega), where s=max(4,m+1)s=\max(4,m+1), m2m\geq 2 and Λm\Lambda_{m} has a uniform upper bound independent of hh. Let the bilinear form Bh(,)B_{h}(\cdot,\cdot) be defined with a sufficiently large η\eta and 𝐮𝐡𝐔hm\boldsymbol{u_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m} be the numerical solution to the problem (14). Then for hh0h\leq h_{0} there exists a constant CC such that

(23) |𝒖𝒖𝒉|CΛmhm1𝒖Hm+1(Ω).|\!|\!|\boldsymbol{u}-\boldsymbol{u_{h}}|\!|\!|\leq C\Lambda_{m}h^{m-1}\|\boldsymbol{u}\|_{H^{m+1}(\Omega)}.
Proof.

From (16), (20) and (21), we have that for any 𝒗𝒉𝐔hm\boldsymbol{v_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m},

|𝒖𝒉𝒗𝒉|2\displaystyle|\!|\!|\boldsymbol{u_{h}}-\boldsymbol{v_{h}}|\!|\!|^{2} CBh(𝒖𝒉𝒗𝒉,𝒖𝒉𝒗𝒉)=CBh(𝒖𝒖𝒉,𝒖𝒉𝒗𝒉)\displaystyle\leq CB_{h}(\boldsymbol{u_{h}}-\boldsymbol{v_{h}},\boldsymbol{u_{h}}-\boldsymbol{v_{h}})=CB_{h}(\boldsymbol{u}-\boldsymbol{u_{h}},\boldsymbol{u_{h}}-\boldsymbol{v_{h}})
C|𝒖𝒗𝒉||𝒖𝒉𝒗𝒉|.\displaystyle\leq C|\!|\!|\boldsymbol{u}-\boldsymbol{v_{h}}|\!|\!||\!|\!|\boldsymbol{u_{h}}-\boldsymbol{v_{h}}|\!|\!|.

By the triangle inequality, there holds

|𝒖𝒖𝒉||𝒖𝒗𝒉|+|𝒗𝒉𝒖𝒉|Cinf𝒗𝒉𝐔hm|𝒖𝒗𝒉|.|\!|\!|\boldsymbol{u}-\boldsymbol{u_{h}}|\!|\!|\leq|\!|\!|\boldsymbol{u}-\boldsymbol{v_{h}}|\!|\!|+|\!|\!|\boldsymbol{v_{h}}-\boldsymbol{u_{h}}|\!|\!|\leq C\inf_{\boldsymbol{v_{h}}\in\boldsymbol{\mathrm{U}}_{h}^{m}}|\!|\!|\boldsymbol{u}-\boldsymbol{v_{h}}|\!|\!|.

Let 𝒗𝒉=𝒖\boldsymbol{v_{h}}=\mathcal{R}\boldsymbol{u}, by the inequality (22), we arrive at

|𝒖𝒖𝒉|CΛm|𝒖𝒖|Chm1𝒖Hm+1(Ω),|\!|\!|\boldsymbol{u}-\boldsymbol{u_{h}}|\!|\!|\leq C\Lambda_{m}|\!|\!|\boldsymbol{u}-\mathcal{R}\boldsymbol{u}|\!|\!|\leq Ch^{m-1}\|\boldsymbol{u}\|_{H^{m+1}(\Omega)},

which completes the proof. ∎

Corollary 1.

Under the same assumptions, there exists a constant CC such that

(24) 𝒖𝒖𝒉L2(Ω)CΛmhm1𝒖Hm+1(Ω).\|\boldsymbol{u}-\boldsymbol{u_{h}}\|_{L^{2}(\Omega)}\leq C\Lambda_{m}h^{m-1}\|\boldsymbol{u}\|_{H^{m+1}(\Omega)}.
Remark 2.

We have proved that under the energy norm |||||||\!|\!|\cdot|\!|\!| the numerical solution 𝐮h\boldsymbol{u}_{h} has the optimal convergence rate. It can be seen that |||||||\!|\!|\cdot|\!|\!| is stronger than L2(Ω)\|\cdot\|_{L^{2}(\Omega)} from the definition of |||||||\!|\!|\cdot|\!|\!|. Hence, from (24) we can conclude that the numerical solution 𝐮h\boldsymbol{u}_{h} at least have the O(hm1)O(h^{m-1}) convergence rate in L2L^{2} norm, which is exactly the convergence rate indicated by our numerical tests.

5. Numerical Results

In this section, we conduct a series numerical experiments to test the performance of our method. For the accuracy 2m42\leq m\leq 4, the threshold #S\#S we used in the examples are listed in Tab. 1. For all examples, the boundary g1,g2g_{1},g_{2} and the right hand side ff in the equation (1) are chosen according to the exact solution.

Example 1. We first consider a fourth-order curl problem defined on the squared domain Ω=(0,1)2\Omega=(0,1)^{2} . The exact solution is chosen by

𝒖(x,y)=[3πsin2(πy)cos(πy)sin3(πx)3πsin2(πx)cos(πx)sin3(πy)],\boldsymbol{u}(x,y)=\begin{bmatrix}3\pi\sin^{2}(\pi y)\cos(\pi y)\sin^{3}(\pi x)\\ -3\pi\sin^{2}(\pi x)\cos(\pi x)\sin^{3}(\pi y)\\ \end{bmatrix},

We solve the problem on a sequence of meshes with the size h=1/8,1/16,1/32,1/64,1/128h=1/8,1/16,1/32,1/64,1/128. The convergence histories under the DG\|\cdot\|_{\mathrm{DG}} (which is equivalent to |||||||\!|\!|\cdot|\!|\!| )and L2(Ω)\|\cdot\|_{L^{2}(\Omega)} are shown in Fig. 1. The error under the energy norm is decreasing at the speed O(hm1)O(h^{m-1}) for fixed mm. For L2L^{2} error, the speed is the same as the energy norm. These results coincide with the theoretical analysis in Theorem 24 .

mm 2 3 4
#S\#S 12 20 27
mm 2 3
#S\#S 25 47
Table 1. The #S\#S used in 2D and 3D examples.
Refer to caption
Refer to caption
Figure 1. The convergence histories under the L2(Ω)\|\cdot\|_{L^{2}(\Omega)} (left) and theDG\|\cdot\|_{\mathrm{DG}} (right) in Example 1.

Example 2. We solve a three-dimensional problem in this case. The computation domain is an unit cube Ω=(0,1)3\Omega=(0,1)^{3} . We select the exact solution as

𝒖(x,y,z)=[sin(πy)sin(πz)sin(πz)sin(πx)sin(πx)sin(πy)],\boldsymbol{u}(x,y,z)=\begin{bmatrix}\sin(\pi y)\sin(\pi z)\\ \sin(\pi z)\sin(\pi x)\\ \sin(\pi x)\sin(\pi y)\\ \end{bmatrix},

We use the tetrahedra meshes generated by the Gmsh software [3]. We solve the problem on three different meshes with the reconstruction order m=2,3m=2,3. The convergence order in both norms is shown in Fig. 2, also consistent with our theoretical predictions.

Refer to caption
Refer to caption
Figure 2. The convergence histories under the L2(Ω)\|\cdot\|_{L^{2}(\Omega)} (left) and the DG\|\cdot\|_{\mathrm{DG}}(right) in Example 2.

6. Conclusion

In this paper, we proposed an IPDG method for the fourth-order curl problem with the reconstructed discontinuous approximation. The approximation space is based on patch reconstruction from piecewise constant sapce and can approximate functions up to high order accuracy. We show numercial experiments in two and three dimensions to examine the order of convergence under the energy norm and the L2L^{2} norm.

Appendix A Calculating the Reconstruction Constant

This appendix gives the method to compute the constant Λm\Lambda_{m} for a given mesh 𝒯h\mathcal{T}_{h}. For any element K𝒯hK\in\mathcal{T}_{h}, let p1,p2,,plp_{1},p_{2},\ldots,p_{l} be a group of standard orthogonal basis functions in m(K)\mathbb{P}_{m}(K) under the L2L^{2} inner product (,)L2(K)(\cdot,\cdot)_{L^{2}(K)}. Then, any polynomial qm(K)q\in\mathbb{P}_{m}(K) can be expanded by a group coefficients 𝒂={aj}j=1ll\boldsymbol{a}=\{a_{j}\}_{j=1}^{l}\in\mathbb{R}^{l} such that q=j=1lajpjq=\sum_{j=1}^{l}a_{j}p_{j}. We can naturally extend qq and all pjp_{j} to the domain S(K)S(K) by the polynomial extension. The main step to get Λm\Lambda_{m} is to compute Λm,K\Lambda_{m,K} for all elements, and Λm,K\Lambda_{m,K} can be represented as

Λm,K2=max𝒂l|𝒂|l22hKd𝒂TBK𝒂,BK={bij}l×l,bij=𝒙I(K)pi(𝒙)pj(𝒙).\Lambda_{m,K}^{2}=\max_{\boldsymbol{a}\in\mathbb{R}^{l}}\frac{|\boldsymbol{a}|_{l^{2}}^{2}}{h_{K}^{d}\boldsymbol{a}^{T}B_{K}\boldsymbol{a}},\quad B_{K}=\{b_{ij}\}_{l\times l},\quad b_{ij}=\sum_{\boldsymbol{x}\in I(K)}p_{i}(\boldsymbol{x})p_{j}(\boldsymbol{x}).

From the matrix representation, Λm,K=(hKdσmin(BK))1/2\Lambda_{m,K}=(h_{K}^{d}\sigma_{\min}(B_{K}))^{-1/2}, where σmin(BK)\sigma_{\min}(B_{K}) is the smallest singular value to BKB_{K}. Hence, it is enough to observe the minimum value of σmin(BK)\sigma_{\min}(B_{K}) for all elements, and Λm\Lambda_{m} can be computed by (13).

As we remarked in Section 3, when the patch is wide enough, Λm\Lambda_{m} will admit a uniform upper bound independent of the mesh size. Here, we will show Λm\Lambda_{m} for different size of the patch. We consider the triangular mesh with h=1/40h=1/40 and the tetrahedral mesh with h=1/16h=1/16 in two and three dimensions, which are used in Example 1 and Example 2. The values of Λm\Lambda_{m} are collected in Fig. 3.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3. Λm\Lambda_{m} in 2d with h=1/40h=1/40 (row 1) / Λm\Lambda_{m} in 3d with h=1/16h=1/16 (row 2).

References

  • [1] S. C. Brenner and L. R. Scott, The Mathematical Theory of Finite Element Methods, third ed., Texts in Applied Mathematics, vol. 15, Springer, New York, 2008.
  • [2] Fioralba Cakoni and Houssem Haddar, A variational approach for the solution of the electromagnetic interior transmission problem for anisotropic media, Inverse Problems and Imaging 1 (2007), 443–456.
  • [3] C. Geuzaine and J. F. Remacle, Gmsh: A 3-D finite element mesh generator with built-in pre- and post-processing facilities, Internat. J. Numer. Methods Engrg. 79 (2009), no. 11, 1309–1331.
  • [4] Jiayu Han and Zhimin Zhang, An hp-version interior penalty discontinuous galerkin method for the quad-curl eigenvalue problem, BIT Numerical Mathematics 63 (2023), article number 56.
  • [5] Qingguo Hong, Jun Hu, Shi Shu, and Jinchao Xu, A discontinuous galerkin method for the fourth-order curl problem, Journal of Computational Mathematics 30 (2012), 565–578.
  • [6] Kaibo Hu, Qian Zhang, and Zhimin Zhang, Simple curl-curl-conforming finite elements in two dimensions, SIAM Journal on Scientific Computing 42 (2020), no. 6, A3859–A3877.
  • [7] T. J. R. Hughes, G. Engel, L. Mazzei, and M. G. Larson, A comparison of discontinuous and continuous Galerkin methods based on error estimates, conservation, robustness and efficiency, Discontinuous Galerkin methods (Newport, RI, 1999), Lect. Notes Comput. Sci. Eng., vol. 11, Springer, Berlin, 2000, pp. 135–146.
  • [8] R. Li, Q. Liu, and F. Yang, A reconstructed discontinuous approximation on unfitted meshes to H(curl)H({\rm curl}) and H(div)H({\rm div}) interface problems, Comput. Methods Appl. Mech. Engrg. 403 (2023), no. part A, Paper No. 115723, 27.
  • [9] R. Li, P. Ming, Z. Sun, F. Yang, and Z. Yang, A discontinuous Galerkin method by patch reconstruction for biharmonic problem, J. Comput. Math. 37 (2019), no. 4, 563–580.
  • [10] R. Li, P. Ming, Z. Sun, and Z. Yang, An arbitrary-order discontinuous Galerkin method with one unknown per element, J. Sci. Comput. 80 (2019), no. 1, 268–288.
  • [11] R. Li, P. Ming, and F. Tang, An efficient high order heterogeneous multiscale method for elliptic problems, Multiscale Model. Simul. 10 (2012), no. 1, 259–283.
  • [12] R. Li, Z. Sun, and F. Yang, Solving eigenvalue problems in a discontinuous approximate space by patch reconstruction, SIAM J. Sci. Comput. 41 (2019), no. 5, A3381–A3400.
  • [13] R. Li and F. Yang, A discontinuous Galerkin method by patch reconstruction for elliptic interface problem on unfitted mesh, SIAM J. Sci. Comput. 42 (2020), no. 2, A1428–A1457.
  • [14] by same author, A least squares method for linear elasticity using a patch reconstructed space, Comput. Methods Appl. Mech. Engrg. 363 (2020), no. 1, 112902.
  • [15] by same author, A sequential least squares method for Poisson equation using a patch reconstructed space, SIAM J. Numer. Anal. 58 (2020), no. 1, 353–374.
  • [16] by same author, A reconstructed discontinuous approximation to Monge-Ampère equation in least square formulation, Adv. Appl. Math. Mech. 15 (2023), no. 5, 1109–1141. MR 4613677
  • [17] Peter Monk and Jiguang Sun, Finite element methods for maxwell’s transmission eigenvalues, SIAM Journal on Scientific Computing 34 (2012), B247–B264.
  • [18] Serge Nicaise, Singularities of the quad curl problem, Journal of Differential Equations 264 (2018), 5025–5069.
  • [19] M. J. D. Powell, Approximation theory and methods, Cambridge University Press, Cambridge-New York, 1981.
  • [20] Jiguang Sun, A mixed fem for the quad-curl eigenvalue problem, Numerische Mathematik 132 (2013), 185–200.
  • [21] Jiguang Sun, Qian Zhang, and Zhimin Zhang, A curl-conforming weak galerkin method for the quad-curl problem, BIT Numerical Mathematics 59 (2019), 1093–1114.
  • [22] Qian Zhang, Lixiu Wang, and Zhimin Zhang, H(curl2curl^{2})-conforming finite elements in 2 dimensions and applications to the quad-curl problem, SIAM Journal on Scientific Computing 41 (2019), A1527–A1547.
  • [23] Shuo Zhang, Mixed schemes for quad-curl equations, ESAIM: Mathematical Modelling and Numerical Analysis 52 (2018), 147–161.
  • [24] Bin Zheng, Qiya Hu, and Jinchao Xu, A nonconforming finite element method for fourth order curl equations in r3r^{3}, Math. Comput. 80 (2010), 1871–1886.