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 mixed finite element method with boundary value correction for the Darcy flow on curved domains

Yongli Hou and Yanqiu Wang 111Corresponding author
Abstract

We propose a boundary value correction method for the Brezzi-Douglas-Marini mixed finite element discretization of the Darcy flow with non-homogeneous Neumann boundary condition on 2D curved domains. The discretization is defined on a body-fitted triangular mesh, i.e. the boundary nodes of the mesh lie on the curved physical boundary. However, the boundary edges of the triangular mesh, which are straight, may not coincide with the curved physical boundary. A boundary value correction technique is then designed to transform the Neumann boundary condition from the physical boundary to the boundary of the triangular mesh. One advantage of the boundary value correction method is that it avoids using curved mesh elements and thus reduces the complexity of implementation. We prove that the proposed method reaches optimal convergence for arbitrary order discretizations. Supporting numerical results are presented.

Key words: mixed finite element method, Neumann boundary condition, curved domain, boundary value correction method

1 Introduction

Many practical problems arising in science and engineering are posed on the curved domain Ω\Omega. Classical finite element methods defined on a polygonal approximation domain Ωh\Omega_{h} often suffer from an additional geometric error due to the difference between Ω\Omega and Ωh\Omega_{h}. This leads to a loss of accuracy for higher-order discretizations [29, 30]. Moreover, when the problem is equipped with an essential boundary condition, effective ways to transfer the boundary condition from Ω\partial\Omega to Ωh\partial\Omega_{h} must be designed. Different strategies have been proposed to solve this problem, for example, the isoparametric finite element method [20, 22] and the isogeometric analysis [21, 17]. Another strategy is the boundary value correction method [8], which shifts the essential boundary condition from Ω\partial\Omega to Ωh\partial\Omega_{h}, and results in a modified variational formulation. One advantage of the boundary value correction method is that it avoids using curved mesh elements and thus reduces the complexity of implementation. Recently, there are many works utilizing the boundary correction strategy, including the discontinuous Galerkin method via extensions from subdomains [15, 16], transferring technique based on the path integral [26], the shifted boundary method [24, 1], the cutFEM [12], the boundary-corrected virtual element method [2] and the boundary-corrected weak Galerkin method [23], etc.

For the mixed finite element method (MFEM), the Neumann boundary condition becomes essential. Again, higher-order MFEM suffers from a loss of accuracy on curved domains. In [3, 4, 5], the authors studied a parametric Raviart-Thomas MFEM on curved domains, which is a generalization of the isoparametric method to the MFEM. In [27], a cutFEM method is proposed in the curved domain, which is based on a primal mixed formulation [7]. The authors of [27] show that the cutFEM method reaches suboptimal convergence.

In this paper, we propose a new boundary corrected MFEM based on the primal mixed formulation [7], and prove that it has optimal convergence rate. Similarly to [18], we weakly impose the Neumann boundary condition. Then, a boundary value correction technique is designed to pull the Neumann boundary data from Ω\partial\Omega to Ωh\partial\Omega_{h}. However, unlike the boundary correction in [8, 1, 24], which considers the Dirichlet boundary condition, the Neumann boundary condition involves the outward normal vector on the boundary, which is different in Ω\partial\Omega and Ωh\partial\Omega_{h}. This poses extra difficulty in the design and analysis of our scheme. Finally, following [14], we added a term (div,div)(\mathrm{div}\,\cdot,\mathrm{div}\,\cdot) to the discrete formulation to increase stability.

The paper is organized as follows. In Section 2, we introduce some notation and settings. In Section 3, we describe the model problem and introduce the boundary value correction method. In Section 4, the discrete space and the discrete form are defined and analyzed. In Section 5, we prove the optimal error estimate. In Section 6, numerical results are presented. Finally, we draw our conclusions in Section 7.

2 Notations and preliminaries

Let Ω\Omega be a bounded open set in 2\mathbb{R}^{2} with Lipschitz continuous and piecewise C2C^{2} boundary Γ\Gamma. Let 𝒯h\mathcal{T}_{h} be a body-fitted triangulation of Ω\Omega, i.e., all boundary vertices of 𝒯h\mathcal{T}_{h} lie on Γ\Gamma. We assume that 𝒯h\mathcal{T}_{h} is geometrically conforming, shape regular, and quasi-uniform. Denote by Ωh\Omega_{h} the polygonal region occupied by the triangular mesh 𝒯h\mathcal{T}_{h}, and by Γh\Gamma_{h} the boundary of Ωh\Omega_{h}. When Ω\Omega has a curved boundary, the polygonal region Ωh\Omega_{h} does not coincide with Ω\Omega, as shown in Fig. 1. Denote by hb\mathcal{E}_{h}^{b} the set of all boundary edges in 𝒯h\mathcal{T}_{h} and by 𝒯hb\mathcal{T}_{h}^{b} all mesh elements that contain at least one edge in hb\mathcal{E}_{h}^{b}. For each ehbe\in\mathcal{E}_{h}^{b}, denote by e~Γ\tilde{e}\subset\Gamma the curved edge cut by two end points of ee (see Fig. 1). We further assume that each e~\tilde{e} is C2C^{2} continuous, i.e. e~\tilde{e} cannot cross the connection point of two C2C^{2} continuous pieces of Γ\Gamma. Note that e~\tilde{e} may coincide with ee when part of Γ\Gamma is flat.

We use Ωhe\Omega_{h}^{e} to denote a crescent-shaped region surrounded by ee and e~\tilde{e}, as shown in Fig. 1. There are three possibilities: (a) ΩheΩ\Ωh\Omega_{h}^{e}\subset\Omega\backslash\Omega_{h}; (b) ΩheΩh\Ω\Omega_{h}^{e}\subset\Omega_{h}\backslash\Omega; (c) e=e~e=\tilde{e} and Ωhe=\Omega_{h}^{e}=\emptyset. If Ω\Omega is convex, we have ΩhΩ\Omega_{h}\subset\Omega. But when Ω\Omega is not convex, there is no inclusion relationship between Ω\Omega and Ωh\Omega_{h}.

Ωhe\Omega_{h}^{e}eee~\widetilde{e}
Ωhe\Omega_{h}^{e}eee~\widetilde{e}
e=e~e=\widetilde{e}
Figure 1: (a) The case when ΩheΩ\Ωh\Omega_{h}^{e}\subset\Omega\backslash\Omega_{h}. (b) The case when ΩheΩh\Ω\Omega_{h}^{e}\subset\Omega_{h}\backslash\Omega. (c) The case when e=e~e=\widetilde{e} and Ωhe=\Omega_{h}^{e}=\emptyset.

Throughout the paper, we assume that the mesh 𝒯h\mathcal{T}_{h} is geometrically conforming, shape regular, and quasi-uniform [10]. For each triangle K𝒯hK\in{\mathcal{T}}_{h}, denote by hKh_{K} the diameter of KK. Let h=maxK𝒯hhKh=\max_{K\in{\mathcal{T}}_{h}}h_{K}. Similarly, for each edge ehbe\in\mathcal{E}_{h}^{b}, denote by heh_{e} the length of ee. Denote by Pk(K)P_{k}(K) the space of polynomials on K𝒯hK\in\mathcal{T}_{h} with degree less than or equal to kk. We use the standard notation for Sobolev spaces [10]. Let Hm(S)H^{m}(S), for mm\in\mathbb{R} and S2S\subset\mathbb{R}^{2} be the usual Sobolev space equipped with the norm m,S\|\cdot\|_{m,S} and the seminorm ||m,S|\cdot|_{m,S}. When m=0m=0, the space H0(S)H^{0}(S) coincides with the square integrable space L2(S)L^{2}(S). Denote by L02(S)L_{0}^{2}(S) the subspace of L2(S)L^{2}(S) consisting of functions with mean value zero. The above notation extends to a portion sΓs\subset\Gamma or sΓhs\subset\Gamma_{h}. For example, m,s\|\cdot\|_{m,s} is the Sobolev norm on Hm(s)H^{m}(s). We also use the notation (,)S(\cdot,\cdot)_{S} to indicate the L2L^{2} inner-product on S2S\subset\mathbb{R}^{2}, and ,s\langle\cdot,\cdot\rangle_{s} to indicate the duality pair on sΓs\subset\Gamma or sΓhs\subset\Gamma_{h}.

All of the above notations extend to vector-valued functions, following the usual rule of product spaces. We use bold face letters to distinguish the vector-valued function spaces from the scalar-valued ones. For example, 𝑷k(K)=[Pk(K)]2\bm{P}_{k}(K)=[P_{k}(K)]^{2} and 𝑯m(S)=[Hm(S)]2\bm{H}^{m}(S)=[H^{m}(S)]^{2} are the spaces of vector-valued functions with each component in Pk(K)P_{k}(K) and Hm(S)H^{m}(S), respectively. Define

H(div,S)\displaystyle H(\mathrm{div},S) ={𝒗𝑳2(S):div𝒗L2(S)},\displaystyle=\{\bm{v}\in\bm{L}^{2}(S):\mathrm{div}\,\bm{v}\in L^{2}(S)\},
Hm(div,S)\displaystyle H^{m}(\mathrm{div},S) ={𝒗𝑯m(S):div𝒗Hm(S)}.\displaystyle=\{\bm{v}\in\bm{H}^{m}(S):\mathrm{div}\,\bm{v}\in H^{m}(S)\}.

The norm in H(div,S)H(\mathrm{div},S) is defined by

𝒗H(div,S)=(𝒗0,S2+div𝒗0,S2)1/2.\|\bm{v}\|_{H(\mathrm{div},S)}=(\|\bm{v}\|_{0,S}^{2}+\|\mathrm{div}\,\bm{v}\|_{0,S}^{2})^{1/2}.

Throughout the paper, we use \lesssim to denote less than or equal up to a constant, and the analogous notation \gtrsim to denote greater than or equal up to a constant.

Finally, we introduce some well-known inequalities.

Lemma 2.1.

(Trace Inequality [10]). For K𝒯hK\in\mathcal{T}_{h}, we have

v0,KhK1/2v0,K+hK1/2|v|1,K,vH1(K).\displaystyle\|v\|_{0,\partial K}{\lesssim}~{}h_{K}^{-1/2}\|v\|_{0,K}+h_{K}^{1/2}|v|_{1,K},\qquad\forall v\in H^{1}(K).
Lemma 2.2.

(Inverse Inequality [10]). Given integers 0ml0\leq m\leq l. For K𝒯hK\in\mathcal{T}_{h}, we have

|q|l,KhKml|q|m,K,qPl(K).\displaystyle|q|_{l,K}{\lesssim}~{}h_{K}^{m-l}|q|_{m,K},\qquad\forall q\in P_{l}(K).
Lemma 2.3.

(Trace-inverse Inequality [10]). Given an integer l0l\geq 0. For K𝒯hK\in\mathcal{T}_{h} and ee be an edge of KK, we have

q0,ehK1/2q0,K,qPl(K).\displaystyle\|q\|_{0,e}{\lesssim}~{}h_{K}^{-1/2}\|q\|_{0,K},\qquad\forall q\in P_{l}(K).
Lemma 2.4.

(Bramble-Hilbert Lemma [10]). Given an integer k0k\geq 0. For K𝒯hK\in\mathcal{T}_{h} and integers 0mk+10\leq m\leq k+1, one has

infξPk(K)(i=0mhKi|vξ|i,K)hKm|v|m,K,vHm(K).\displaystyle\inf_{\xi\in P_{k}(K)}\bigg{(}\sum_{i=0}^{m}h_{K}^{i}|v-\xi|_{i,K}\bigg{)}\lesssim h_{K}^{m}|v|_{m,K},\qquad\forall v\in H^{m}(K).

3 Model problem and boundary value correction method

In this section, we briefly describe the model problem and the boundary value correction method [8]. Consider the Darcy’s equation with the Neumann boundary condition

𝒖+p=\displaystyle\bm{u}+\nabla p= 0,\displaystyle~{}0,\quad inΩ,\displaystyle\text{in}\,\Omega, (3.1)
div𝒖=\displaystyle\mathrm{div}\,\bm{u}= f,\displaystyle~{}f,\quad inΩ,\displaystyle\text{in}\,\Omega,
𝒖𝐧=\displaystyle\bm{u}\cdot\bm{\mathop{}\mathrm{n}}= gN,\displaystyle~{}g_{N},\quad onΓ,\displaystyle\text{on}\,\Gamma,

with fL2(Ω)f\in L^{2}(\Omega) and gNH1/2(Γ)g_{N}\in H^{-1/2}(\Gamma). Here, 𝐧\bm{\mathop{}\mathrm{n}} denotes the unit outward normal on Γ\Gamma. Problem (3.1) admits a unique weak solution (𝒖,p)H(div,Ω)×L02(Ω)(\bm{u},p)\in H(\mathrm{div},\Omega)\times L_{0}^{2}(\Omega) as long as the following compatibility condition holds:

Ωfdx=ΓgNds.\displaystyle\int_{\Omega}f\mathop{}\mathrm{d}x=\int_{\Gamma}g_{N}\mathop{}\mathrm{d}s. (3.2)

The weak formulation of Problem (3.1) can be written as: Find (𝒖,p)H(div,Ω)×L02(Ω)(\bm{u},p)\in H(\mathrm{div},\Omega)\times L_{0}^{2}(\Omega) satisfying 𝒖𝐧=gN\bm{u}\cdot\bm{\mathop{}\mathrm{n}}=g_{N} on Γ\Gamma, such that

{(𝒖,𝒗)Ω(div𝒗,p)Ω=0,𝒗H0(div,Ω),(div𝒖,q)Ω=(f,q)Ω,qL02(Ω),\left\{\begin{aligned} (\bm{u},\bm{v})_{\Omega}-(\mathrm{div}\,\bm{v},p)_{\Omega}&=0,\quad&\forall&\bm{v}\in H_{0}(\mathrm{div},\Omega),\\ -(\mathrm{div}\,\bm{u},q)_{\Omega}&=-(f,q)_{\Omega},\quad&\forall&q\in L_{0}^{2}(\Omega),\end{aligned}\right. (3.3)

where H0(div,Ω)={𝒗H(div,Ω):𝒗𝒏=0 on Γ}H_{0}(\mathrm{div},\Omega)=\{\bm{v}\in H(\mathrm{div},\Omega):\bm{v}\cdot\bm{n}=0\textrm{ on }\Gamma\}. The existence and uniqueness of a weak solution to the mixed problem (3.3) can be found in [10].

We use the Brezzi-Douglas-Marini (BDM) finite element method [11] on triangulation 𝒯h\mathcal{T}_{h} to discretize Equation (3.3). The discretization is defined on Ωh\Omega_{h} rather than on Ω\Omega. Note that in (3.3), the Neumann boundary condition 𝒖𝐧=gN\bm{u}\cdot\bm{\mathop{}\mathrm{n}}=g_{N} on Γ\Gamma becomes essential. This causes the main difficulty in its finite element discretization when Γ\Gamma is curved and hence not equal to Γh\Gamma_{h}. Another difficulty is that the compatibility condition (3.2) holds only on Ω\Omega but not on Ωh\Omega_{h}. We shall take care of these issues one by one. First, we use a boundary correction technique [8] to transform the boundary data from Γ\Gamma to Γh\Gamma_{h}.

Assume that there exists a map ρh:ΓhΓ\rho_{h}:\Gamma_{h}\rightarrow\Gamma such that

ρh(𝒙h):=𝒙h+δh(𝒙h)𝝂h(𝒙h),\displaystyle\rho_{h}(\bm{x}_{h}):=\bm{x}_{h}+\delta_{h}(\bm{x}_{h})\bm{\nu}_{h}(\bm{x}_{h}), (3.4)

as shown in Fig. 2, where 𝝂h\bm{\nu}_{h} is the unit vector pointing from 𝒙hΓh\bm{x}_{h}\in\Gamma_{h} to ρh(𝒙h)Γ\rho_{h}(\bm{x}_{h})\in\Gamma, and δh(𝒙h)=|ρh(𝒙h)𝒙h|\delta_{h}(\bm{x}_{h})=|\rho_{h}(\bm{x}_{h})-\bm{x}_{h}|. For convenience, we denote 𝒙:=ρh(𝒙h)\bm{x}:=\rho_{h}(\bm{x}_{h}). Define 𝐧~(𝒙h):=𝐧ρh(𝒙h)\widetilde{\bm{\mathop{}\mathrm{n}}}(\bm{x}_{h}):=\bm{\mathop{}\mathrm{n}}\circ\rho_{h}(\bm{x}_{h}), i.e. the unit outward normal vector on 𝒙Γ\bm{x}\in\Gamma is pulled to the corresponding 𝒙h\bm{x}_{h}. Denote by 𝐧h\bm{\mathop{}\mathrm{n}}_{h} the unit outward normal vector on Γh\Gamma_{h}. Since Γ\Gamma is piecewise C2C^{2} continuous, there exists a map ρh\rho_{h} satisfying [12]

δ=sup𝒙hΓhδh(𝒙h)h2,𝐧~𝐧hL(Γh)h.\displaystyle\delta=\sup_{\bm{x}_{h}\in\Gamma_{h}}\delta_{h}(\bm{x}_{h})\lesssim h^{2},\qquad\|\widetilde{\bm{\mathop{}\mathrm{n}}}-\bm{\mathop{}\mathrm{n}}_{h}\|_{L^{\infty}(\Gamma_{h})}\lesssim h. (3.5)
Remark 3.1.

In this paper, we do not specify the map ρh\rho_{h}. We only require that the distance function δh(𝐱h)\delta_{h}(\bm{x}_{h}) satisfies (3.5). For body-fitted meshes, this surely holds by simply setting 𝛎h=𝐧h\bm{\nu}_{h}=\bm{\mathop{}\mathrm{n}}_{h}. In fact, (3.5) can be true on certain unfitted meshes [12]. The method and analysis in this paper extend to unfitted meshes that satisfy (3.5), with a little modification.

𝒙h{\bm{x}_{h}}Γh\Gamma_{h}Γ\Gammaδh(𝒙h)\delta_{h}(\bm{x}_{h})𝒙\bm{x}𝝂h\bm{\nu}_{h}
Figure 2: The distance δh(𝒙h)\delta_{h}(\bm{x}_{h}) and the unit vector 𝝂𝒉\bm{\nu_{h}} on Γh\Gamma_{h}.

For a sufficiently smooth vector-valued function 𝒗\bm{v} defined between Γ\Gamma and Γh\Gamma_{h}, consider its mm-th order Taylor expansion along 𝝂𝒉\bm{\nu_{h}}

𝒗(ρh(𝒙h))\displaystyle\bm{v}(\rho_{h}(\bm{x}_{h})) =j=0mδhj(𝒙h)j!𝝂hj𝒗(𝒙h)+Rm𝒗(𝒙h)\displaystyle=\sum_{j=0}^{m}\frac{\delta_{h}^{j}(\bm{x}_{h})}{j!}\partial_{\bm{\nu}_{h}}^{j}\bm{v}(\bm{x}_{h})+R^{m}\bm{v}(\bm{x}_{h})
=:(Tm𝒗)(𝒙h)+Rm𝒗(𝒙h),\displaystyle=:(T^{m}\bm{v})(\bm{x}_{h})+R^{m}\bm{v}(\bm{x}_{h}),

where 𝝂hj\partial_{\bm{\nu}_{h}}^{j} is the jj-th order directional derivative and the remainder term satisfies

|Rm𝒗(𝒙h)|=o(δm).|R^{m}\bm{v}(\bm{x}_{h})|=o(\delta^{m}).

For simplicity of notation, denote

(Tm𝒗)(𝒙h):=𝒗(𝒙h)+(T1m𝒗)(𝒙h)where(T1m𝒗)(𝒙h):=j=1mδhj(𝒙h)j!𝝂hj𝒗(𝒙h).\displaystyle(T^{m}\bm{v})(\bm{x}_{h}):=\bm{v}(\bm{x}_{h})+(T_{1}^{m}\bm{v})(\bm{x}_{h})\quad\textrm{where}\quad(T_{1}^{m}\bm{v})(\bm{x}_{h}):=\sum_{j=1}^{m}\frac{\delta_{h}^{j}(\bm{x}_{h})}{j!}\partial_{\bm{\nu}_{h}}^{j}\bm{v}(\bm{x}_{h}). (3.6)

Both Tm𝒗T^{m}\bm{v} and T1m𝒗T_{1}^{m}\bm{v} are functions defined on Γh\Gamma_{h}. Denote by 𝒗~(𝒙h):=𝒗ρh(𝒙h)\widetilde{\bm{v}}(\bm{x}_{h}):=\bm{v}\circ\rho_{h}(\bm{x}_{h}) the pullback of 𝒗\bm{v} from Γ\Gamma to Γh\Gamma_{h}. Then clearly

𝒗~Tm𝒗=Rm𝒗on Γh.\displaystyle\widetilde{\bm{v}}-T^{m}\bm{v}=R^{m}\bm{v}\qquad\textrm{on }\Gamma_{h}. (3.7)

Assume that (3.5) holds, we have the following two lemmas.

Lemma 3.1.

(cf. [23]). Given an integer j0j\geq 0. For K𝒯hbK\in\mathcal{T}_{h}^{b} and qPj(K)q\in P_{j}(K), one has

eKΓhq0,Ωhe2hKq0,K2.\displaystyle\sum_{e\subset\partial K\cap\Gamma_{h}}\|q\|_{0,\Omega_{h}^{e}}^{2}\lesssim h_{K}\|q\|_{0,K}^{2}.
Lemma 3.2.

(cf. [9]). For each ehbe\in\mathcal{E}_{h}^{b} and vH1(ΩΩh)v\in H^{1}(\Omega\cup\Omega_{h}), one has

v0,e\displaystyle\|v\|_{0,e} v0,e~+h|v|1,Ωhe,\displaystyle\lesssim\|v\|_{0,\widetilde{e}}+h|v|_{1,\Omega_{h}^{e}}, (3.8)
v0,e~\displaystyle\|v\|_{0,\widetilde{e}} v0,e+h|v|1,Ωhe,\displaystyle\lesssim\|v\|_{0,e}+h|v|_{1,\Omega_{h}^{e}}, (3.9)
v0,Ωhe\displaystyle\|v\|_{0,\Omega_{h}^{e}} δ1/2v0,e~+δv0,Ωhe.\displaystyle\lesssim\delta^{1/2}\|v\|_{0,\widetilde{e}}+\delta\|\nabla v\|_{0,\Omega_{h}^{e}}. (3.10)

Moreover, when v|Ω=0v|_{\partial\Omega}=0, one has

v0,ehv1,Ωhe.\displaystyle\|v\|_{0,e}\lesssim h\|v\|_{1,\Omega_{h}^{e}}. (3.11)

Lemmas 3.1-3.2 can be easily extended to vector-valued functions.

4 The finite element discretization

For a given integer k1k\geq 1, define the BDM finite element space on 𝒯h\mathcal{T}_{h} by

Vh\displaystyle V_{h} ={𝒗hH(div,Ωh):𝒗h|K𝑷k(K),K𝒯h},\displaystyle=\{\bm{v}_{h}\in H(\mathrm{div},\Omega_{h}):\bm{v}_{h}|_{K}\in\bm{P}_{k}(K),\,\forall\,K\in\mathcal{T}_{h}\},
Qh\displaystyle Q_{h} ={qhL2(Ωh):qh|KPk1(K),K𝒯h},\displaystyle=\{q_{h}\in L^{2}(\Omega_{h}):q_{h}|_{K}\in P_{k-1}(K),\,\forall\,K\in\mathcal{T}_{h}\},

and denote Q0h:=QhL02(Ωh)Q_{0h}:=Q_{h}\cap L_{0}^{2}(\Omega_{h}).

For K𝒯hK\in\mathcal{T}_{h}, define the local interpolation IK:𝑯s(K)BDMk(K),s>1/2I_{K}:\bm{H}^{s}(K)\rightarrow BDM_{k}(K),s>1/2, by

{(IK𝒗,ψk1)K=(𝒗,ψk1)K,ψk1Pk1(K),IK𝒗𝐧h,ψke=𝒗𝐧h,ψke,ψkPk(e), on all edge eK,(IK𝒗,curl(bKψk2))K=(𝒗,curl(bKψk2))K,ψk2Pk2(K), when k2,\begin{cases}(I_{K}\bm{v},\nabla\psi_{k-1})_{K}=(\bm{v},\nabla\psi_{k-1})_{K},\quad&\forall\,\psi_{k-1}\in P_{k-1}(K),\\ \langle I_{K}\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h},\psi_{k}\rangle_{e}=\langle\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h},\psi_{k}\rangle_{e},\quad&\forall\,\psi_{k}\in P_{k}(e),\text{ on all edge }e\subset\partial K,\\ (I_{K}\bm{v},\text{curl}(b_{K}\psi_{k-2}))_{K}=(\bm{v},\text{curl}(b_{K}\psi_{k-2}))_{K},\quad&\forall\,\psi_{k-2}\in P_{k-2}(K),\textrm{ when }k\geq 2,\end{cases} (4.1)

where bK=λ1λ2λ3b_{K}=\lambda_{1}\lambda_{2}\lambda_{3} is a bubble function, and λi,i=1,2,3\lambda_{i},\,i=1,2,3, are the barycentric coordinates associated with KK. Denote the global interpolation by Ih=K𝒯hIK:Hs(Ωh)VhI_{h}=\prod_{K\in\mathcal{T}_{h}}I_{K}:H^{s}(\Omega_{h})\to V_{h}.

Denote by Πk10,K\Pi_{k-1}^{0,K} and Πk0,e\Pi_{k}^{0,e} the L2L^{2} orthogonal projections onto Pk1(K)P_{k-1}(K) and Pk(e)P_{k}(e), respectively. Define the global projection Πh=(K𝒯hΠk10,K):L2(Ωh)Qh\Pi_{h}=\left(\prod_{K\in\mathcal{T}_{h}}\Pi_{k-1}^{0,K}\right):L^{2}(\Omega_{h})\to Q_{h}.

Lemma 4.1.

(cf. [11]). The following commutative diagram holds:

V\textstyle{V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}div\scriptstyle{\mathrm{div}}Ih\scriptstyle{I_{h}}L2(Ωh)\textstyle{L^{2}(\Omega_{h})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Πh\scriptstyle{\Pi_{h}}Vh\textstyle{V_{h}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}div\scriptstyle{\mathrm{div}}Qh\textstyle{Q_{h}}

where V:=H(div,Ωh)Hs(Ωh)V:=H(\mathrm{div},\Omega_{h})\cap H^{s}(\Omega_{h}) with s>1/2s>1/2. That is, for 𝐯V\bm{v}\in V one has

divIK𝒗=Πk10,Kdiv𝒗K𝒯h.\displaystyle\mathrm{div}\,I_{K}\bm{v}=\Pi_{k-1}^{0,K}\mathrm{div}\,\bm{v}\qquad\forall K\in\mathcal{T}_{h}. (4.2)
Lemma 4.2.

For K𝒯hK\in\mathcal{T}_{h}, the interpolation IKI_{K} satisfies, for 𝐯𝐇1(K)\bm{v}\in\bm{H}^{1}(K),

IK𝒗0,K\displaystyle\|I_{K}\bm{v}\|_{0,K} 𝒗0,K+hK|𝒗|1,K,\displaystyle\lesssim\|\bm{v}\|_{0,K}+h_{K}|\bm{v}|_{1,K}, (4.3)
|IK𝒗|1,K\displaystyle|I_{K}\bm{v}|_{1,K} |𝒗|1,K.\displaystyle\lesssim|\bm{v}|_{1,K}. (4.4)
Proof.

Inequality (4.3) is well known [19]. Inequality (4.4) follows immediately from applying (4.3) to 𝒗1|K|K𝒗dx\bm{v}-\frac{1}{|K|}\int_{K}\bm{v}\mathop{}\mathrm{d}x, the inverse inequality, and the Bramble-Hilbert lemma. More specifically, denoting 𝒗¯=1|K|K𝒗dx\bar{\bm{v}}=\frac{1}{|K|}\int_{K}\bm{v}\mathop{}\mathrm{d}x, one has

|IK𝒗|1,K\displaystyle|I_{K}\bm{v}|_{1,K} =|IK𝒗𝒗¯|1,K=|IK(𝒗𝒗¯)|1,K\displaystyle=|I_{K}\bm{v}-\bar{\bm{v}}|_{1,K}=|I_{K}(\bm{v}-\bar{\bm{v}})|_{1,K}
hK1IK(𝒗𝒗¯)0,K\displaystyle\lesssim h_{K}^{-1}\|I_{K}(\bm{v}-\bar{\bm{v}})\|_{0,K}
hK1(𝒗𝒗¯0,K+hK|𝒗|1,K)\displaystyle\lesssim h_{K}^{-1}(\|\bm{v}-\bar{\bm{v}}\|_{0,K}+h_{K}|\bm{v}|_{1,K})
|𝒗|1,K.\displaystyle\lesssim|\bm{v}|_{1,K}.

By the lemmas 2.1,2.2,2.4, it is not hard to see that

Lemma 4.3.

Given a positive integer kk. For any nonnegative integers jik+1j\leq i\leq k+1 and any K𝒯hK\in\mathcal{T}_{h}, one has

hKj|vΠk0,Kv|j,K\displaystyle h_{K}^{j}|v-\Pi_{k}^{0,K}v|_{j,K} hKij|v|i,K,\displaystyle\lesssim h_{K}^{i-j}|v|_{i,K}, vHi(K),\displaystyle\forall v\in H^{i}(K), (4.5)
hKj|𝒗IK𝒗|j,K\displaystyle h_{K}^{j}|\bm{v}-I_{K}\bm{v}|_{j,K} hKij|𝒗|i,K,\displaystyle\lesssim h_{K}^{i-j}|\bm{v}|_{i,K}, 𝒗𝑯i(K),\displaystyle\forall\bm{v}\in\bm{H}^{i}(K), (4.6)
div(𝒗IK𝒗)0,K\displaystyle\|\mathrm{div}\,(\bm{v}-I_{K}\bm{v})\|_{0,K} hKi|div𝒗|i,K,\displaystyle\lesssim h_{K}^{i}|\mathrm{div}\,\bm{v}|_{i,K}, 𝒗Hi(div,K).\displaystyle\forall\bm{v}\in H^{i}(\mathrm{div},K). (4.7)

Finally, we recall the extension property of Sobolev spaces:

Lemma 4.4.

(cf. [28]). Given a Lipschitz domain Ω\Omega in 2\mathbb{R}^{2} and s,s0s\in\mathbb{R},\,s\geq 0, there exists an extension operator E:Hs(Ω)Hs(2)E:H^{s}(\Omega)\to H^{s}(\mathbb{R}^{2}) such that

Ev|Ω=v,Evs,2vs,Ω,vHs(Ω).\displaystyle Ev|_{\Omega}=v,\quad\|Ev\|_{s,\mathbb{R}^{2}}\lesssim\|v\|_{s,\Omega},\qquad\forall v\in H^{s}(\Omega).

In Lemma 4.4, the hidden constant in \lesssim may depend on the shape of Ω\Omega. Throughout the paper, we shall only use Lemma 4.4 on Ω\Omega but not on Ωh\Omega_{h}, in order to ensure that the hidden constant does not depend on hh. For brevity, we also denote the extension by vE=Evv^{E}=Ev. The extension and its notation can be extended to vector-valued functions.

4.1 Discretization with boundary correction

Let (𝒖,p)(\bm{u},\,p) be the solution to (3.1) and (𝒖E,pE)(\bm{u}^{E},\,p^{E}) be their extension to 2\mathbb{R}^{2}. Testing with 𝒗hVh\bm{v}_{h}\in V_{h} in Ωh\Omega_{h} gives

(𝒖E+pE,𝒗h)Ωh=(𝒖E,𝒗h)Ωh(pE,div𝒗h)Ωh+ehb𝒗h𝐧h,pEe.\displaystyle(\bm{u}^{E}+\nabla p^{E},\bm{v}_{h})_{\Omega_{h}}=(\bm{u}^{E},\bm{v}_{h})_{\Omega_{h}}-(p^{E},\mathrm{div}\,\bm{v}_{h})_{\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h},p^{E}\rangle_{e}. (4.8)

We point out that although 𝒖E+pE=0\bm{u}^{E}+\nabla p^{E}=0 in Ω\Omega, the term 𝒖E+pE\bm{u}^{E}+\nabla p^{E} is not in general equal to zero in Ωh\Ω\Omega_{h}\backslash\Omega. Next, using (3.7), it holds

(Tm𝒖E+Rm𝒖E)𝐧~g~N=0,onΓh,\displaystyle(T^{m}\bm{u}^{E}+R^{m}\bm{u}^{E})\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}-\widetilde{g}_{N}=0,\quad\text{on}\,\Gamma_{h},

where g~N(𝒙h)=gN(ρ(𝒙h))\widetilde{g}_{N}(\bm{x}_{h})=g_{N}(\rho(\bm{x}_{h})) is a pull-back of the Neumann boundary data gNg_{N} from Γ\Gamma to Γh\Gamma_{h}. Equation (4.8) can then be rewritten into

(𝒖E,𝒗h)Ωh\displaystyle(\bm{u}^{E},\bm{v}_{h})_{\Omega_{h}} +(div𝒖E,div𝒗h)Ωh+ehbhK1Tm𝒖E𝐧~,Tm𝒗h𝐧~e(pE,div𝒗h)Ωh\displaystyle+(\mathrm{div}\,\bm{u}^{E},\mathrm{div}\,\bm{v}_{h})_{\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}\langle h_{K}^{-1}T^{m}\bm{u}^{E}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}},T^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\rangle_{e}-(p^{E},\mathrm{div}\,\bm{v}_{h})_{\Omega_{h}} (4.9)
+ehb𝒗h𝐧h,pEe=(𝒖E+pE,𝒗h)Ωh+(fE+(div𝒖EfE),div𝒗h)Ωh\displaystyle+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h},p^{E}\rangle_{e}=(\bm{u}^{E}+\nabla p^{E},\bm{v}_{h})_{\Omega_{h}}+(f^{E}+(\mathrm{div}\,\bm{u}^{E}-f^{E}),\mathrm{div}\,\bm{v}_{h})_{\Omega_{h}}
+ehbhK1(g~NRm𝒖E𝐧~),Tm𝒗h𝐧~e.\displaystyle\qquad\qquad\qquad\qquad\qquad\,+\sum_{e\in\mathcal{E}_{h}^{b}}\langle h_{K}^{-1}(\widetilde{g}_{N}-R^{m}\bm{u}^{E}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}),T^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\rangle_{e}.

Again, note that div𝒖EfE\mathrm{div}\,\bm{u}^{E}-f^{E} vanishes in Ω\Omega but not necessarily vanishes in Ωh\Ω\Omega_{h}\backslash\Omega.

Similarly, for any qhQhq_{h}\in Q_{h}, one has

(div𝒖E,qh)Ωh=(fE+(div𝒖EfE),qh)Ωh.\displaystyle(\mathrm{div}\,\bm{u}^{E},q_{h})_{\Omega_{h}}=(f^{E}+(\mathrm{div}\,\bm{u}^{E}-f^{E}),q_{h})_{\Omega_{h}}. (4.10)

We then introduce a discretization for (3.1), using (4.9)-(4.10). First, for 𝒘,𝒗H(div,ΩΩh)\bm{w},\,\bm{v}\in H(\mathrm{div},\Omega\cup\Omega_{h}) that are smooth enough so that TmT^{m} is well-defined, and for qL2(ΩΩh)q\in L^{2}(\Omega\cup\Omega_{h}), define bilinear forms and linear forms

ah(𝒘,𝒗):\displaystyle a_{h}(\bm{w},\bm{v}): =(𝒘,𝒗)Ωh+(div𝒘,div𝒗)Ωh+ehbhK1Tm𝒘𝐧~,Tm𝒗𝐧~e,\displaystyle=(\bm{w},\bm{v})_{\Omega_{h}}+(\mathrm{div}\,\bm{w},\mathrm{div}\,\bm{v})_{\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}\langle h_{K}^{-1}T^{m}\bm{w}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}},T^{m}\bm{v}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\rangle_{e}, (4.11)
bh1(𝒗,q):\displaystyle b_{h1}(\bm{v},q): =(div𝒗,q)Ωh+ehb𝒗𝐧h,qe,\displaystyle=-(\mathrm{div}\,\bm{v},q)_{\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h},q\rangle_{e},
bh0(𝒗,q):\displaystyle b_{h0}(\bm{v},q): =(div𝒗,q)Ωh,\displaystyle=-(\mathrm{div}\,\bm{v},q)_{\Omega_{h}},
lh(𝒗):\displaystyle l_{h}(\bm{v}): =(fE,div𝒗)Ωh+ehbhK1g~N,Tm𝒗𝐧~e.\displaystyle=(f^{E},\mathrm{div}\,\bm{v})_{\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}\langle h_{K}^{-1}\widetilde{g}_{N},T^{m}\bm{v}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\rangle_{e}.

Define the discrete formulation: Find (𝐮h,ph)Vh×Q0h(\bm{u}_{h},p_{h})\in V_{h}\times Q_{0h} such that

{ah(𝒖h,𝒗h)+bh1(𝒗h,ph)=lh(𝒗h),𝒗hVh,bh0(𝒖h,qh)=(fE,qh)Ωh,qhQ0h.\left\{\begin{aligned} a_{h}(\bm{u}_{h},\bm{v}_{h})+b_{h1}(\bm{v}_{h},p_{h})&=l_{h}(\bm{v}_{h}),\quad&\forall&\bm{v}_{h}\in V_{h},\\ b_{h0}(\bm{u}_{h},q_{h})&=-(f^{E},q_{h})_{\Omega_{h}},\quad&\forall&q_{h}\in Q_{0h}.\end{aligned}\right. (4.12)

In the rest of this section, we shall prove the well-posedness of (4.12), and discuss its “compatibility” condition.

4.2 Well-posedness of (4.12)

We first define a mesh-dependent norm in VhV_{h} as follows

𝒗0,h\displaystyle\|\bm{v}\|_{0,h} =(𝒗H(div,Ωh)2+ehbhK1/2Tm𝒗𝐧~0,e2)1/2.\displaystyle=\left(\|\bm{v}\|_{H(\mathrm{div},\Omega_{h})}^{2}+\sum_{e\in\mathcal{E}_{h}^{b}}\|h_{K}^{-1/2}T^{m}\bm{v}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}^{2}\right)^{1/2}.

Note that the norm 0,h\|\cdot\|_{0,h} is also meaningful for functions in Hm(ΩΩh)H^{m}(\Omega\cup\Omega_{h}) smooth enough so that TmT^{m} is well-defined.

Lemma 4.5.

Under the assumption (3.5), for 𝐯hVh\bm{v}_{h}\in V_{h} we have

hK1/2T1m𝒗h0,e\displaystyle\|h_{K}^{-1/2}T_{1}^{m}\bm{v}_{h}\|_{0,e} h|𝒗h|1,K,\displaystyle\leq h|\bm{v}_{h}|_{1,K}, (4.13)
hK1/2T1m𝒗h0,e\displaystyle\|h_{K}^{-1/2}T_{1}^{m}\bm{v}_{h}\|_{0,e} 𝒗h0,K.\displaystyle\lesssim\|\bm{v}_{h}\|_{0,K}. (4.14)
Proof.

By the definition of T1m𝒗hT_{1}^{m}\bm{v}_{h} in (3.6), and lemmas 2.1-2.2, we have

hK1/2T1m𝒗h0,e\displaystyle\|h_{K}^{-1/2}T_{1}^{m}\bm{v}_{h}\|_{0,e} =hK1/2j=1mδhjj!𝝂hj𝒗h0,e\displaystyle=h_{K}^{-1/2}\left\|\sum_{j=1}^{m}\frac{\delta_{h}^{j}}{j!}\partial_{\bm{\nu}_{h}}^{j}\bm{v}_{h}\right\|_{0,e}
j=1mh1/2δhj𝝂hj𝒗h0,e\displaystyle\lesssim\sum_{j=1}^{m}h^{-1/2}\delta_{h}^{j}\|\partial_{\bm{\nu}_{h}}^{j}\bm{v}_{h}\|_{0,e}
j=1m(δhh)j|𝒗h|1,K\displaystyle\lesssim\sum_{j=1}^{m}\left(\frac{\delta_{h}}{h}\right)^{j}|\bm{v}_{h}|_{1,K}
(δhh)|𝒗h|1,K\displaystyle\lesssim\left(\frac{\delta_{h}}{h}\right)|\bm{v}_{h}|_{1,K}
h|𝒗h|1,K,\displaystyle\lesssim h|\bm{v}_{h}|_{1,K},

where we have used (3.5) in the last inequality. This completes the proof of (4.13). Inequality (4.14) follows immediately from the inverse inequality in Lemma 2.2. ∎

Lemma 4.6.

Under the assumption (3.5), for 𝐯hVh\bm{v}_{h}\in V_{h}, we have

ehbhK1𝒗h𝐧h0,e2𝒗h0,h2.\displaystyle\sum_{e\in\mathcal{E}_{h}^{b}}h_{K}^{-1}\|\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e}^{2}\lesssim\|\bm{v}_{h}\|_{0,h}^{2}.
Proof.

For any ehbe\in\mathcal{E}_{h}^{b}, by the definition of TmT^{m} and lemmas 4.5, 2.1, and 2.2, one gets

𝒗h𝐧h0,e\displaystyle\|\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e} =𝒗h𝐧~+𝒗h(𝐧~𝐧h)0,e\displaystyle=\|\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}+\bm{v}_{h}\cdot(\widetilde{\bm{\mathop{}\mathrm{n}}}-\bm{\mathop{}\mathrm{n}}_{h})\|_{0,e}
=Tm𝒗h𝐧~T1m𝒗h𝐧~+𝒗h(𝐧~𝐧h)0,e\displaystyle=\|T^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}-T_{1}^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}+\bm{v}_{h}\cdot(\widetilde{\bm{\mathop{}\mathrm{n}}}-\bm{\mathop{}\mathrm{n}}_{h})\|_{0,e}
Tm𝒗h𝐧~0,e+T1m𝒗h𝐧~0,e+𝒗h(𝐧~𝐧h)0,e\displaystyle\leq\|T^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}+\|T_{1}^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}+\|\bm{v}_{h}\cdot(\widetilde{\bm{\mathop{}\mathrm{n}}}-\bm{\mathop{}\mathrm{n}}_{h})\|_{0,e}
Tm𝒗h𝐧~0,e+hK1/2𝒗h0,K.\displaystyle\lesssim\|T^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}+h_{K}^{1/2}\|\bm{v}_{h}\|_{0,K}.

Summing up over ehbe\in\mathcal{E}_{h}^{b}, this completes the proof of the lemma. ∎

Lemma 4.7.

For 𝐮h\bm{u}_{h}, 𝐯hVh\bm{v}_{h}\in V_{h} and qhQ0hq_{h}\in Q_{0h}, we have

ah(𝒖h,𝒗h)\displaystyle a_{h}(\bm{u}_{h},\bm{v}_{h}) 𝒖h0,h𝒗h0,h,\displaystyle\leq\|\bm{u}_{h}\|_{0,h}\|\bm{v}_{h}\|_{0,h},\qquad ah(𝒗h,𝒗h)\displaystyle a_{h}(\bm{v}_{h},\bm{v}_{h}) =𝒗h0,h2,\displaystyle=\|\bm{v}_{h}\|_{0,h}^{2},
bh1(𝒗h,qh)\displaystyle b_{h1}(\bm{v}_{h},q_{h}) 𝒗h0,hqh0,Ωh,\displaystyle\lesssim\|\bm{v}_{h}\|_{0,h}\|q_{h}\|_{0,\Omega_{h}},\qquad bh0(𝒗h,qh)\displaystyle b_{h0}(\bm{v}_{h},q_{h}) 𝒗h0,hqh0,Ωh.\displaystyle\lesssim\|\bm{v}_{h}\|_{0,h}\|q_{h}\|_{0,\Omega_{h}}.
Proof.

From the definition, it is obvious that ah(𝒗h,𝒗h)=𝒗h0,h2a_{h}(\bm{v}_{h},\bm{v}_{h})=\|\bm{v}_{h}\|_{0,h}^{2}. The upper bounds of aha_{h} and bh0b_{h0} follow immediately from the Schwarz inequality.

Using the Schwarz inequality and lemmas 4.6, 2.3, we obtain

|bh1(𝒗h,qh)|\displaystyle|b_{h1}(\bm{v}_{h},q_{h})| div𝒗h0,Ωhqh0,Ωh+ehbhK1/2𝒗h𝐧h0,ehK1/2qh0,e\displaystyle\leq\|\mathrm{div}\,\bm{v}_{h}\|_{0,\Omega_{h}}\|q_{h}\|_{0,\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}h_{K}^{-1/2}\|\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e}h_{K}^{1/2}\|q_{h}\|_{0,e}
div𝒗h0,Ωhqh0,Ωh+𝒗h0,hqh0,Ωh\displaystyle\lesssim\|\mathrm{div}\,\bm{v}_{h}\|_{0,\Omega_{h}}\|q_{h}\|_{0,\Omega_{h}}+\|\bm{v}_{h}\|_{0,h}\|q_{h}\|_{0,\Omega_{h}}
𝒗h0,hqh0,Ωh.\displaystyle\lesssim\|\bm{v}_{h}\|_{0,h}\|q_{h}\|_{0,\Omega_{h}}.

Lemma 4.7 gives the boundedness of ah(,)a_{h}(\cdot,\cdot), bh1(,)b_{h1}(\cdot,\cdot), bh0(,)b_{h0}(\cdot,\cdot), and the coercivity of ah(,)a_{h}(\cdot,\cdot). We then prove the inf\inf-sup\sup condition for bh1(,)b_{h1}(\cdot,\cdot) and bh0(,)b_{h0}(\cdot,\cdot). Note that a function qhQ0hq_{h}\in Q_{0h} is mean-value free on Ωh\Omega_{h} but not on Ω\Omega. This poses additional difficulty in the analysis. To this end, we first introduce the following notation to distinguish between two types of crescent-shaped regions Ωhe\Omega_{h}^{e}, for each ehbe\in\mathcal{E}_{h}^{b}, as shown in Fig. 3,

Ωhe={Ωhe,+,inΩ\Ωh,Ωhe,,inΩh\Ω.σ={1,inΩhe,+,1,inΩhe,.\Omega_{h}^{e}=\begin{cases}\Omega_{h}^{e,+},\quad\text{in}\,\Omega\backslash\Omega_{h},\\ \Omega_{h}^{e,-},\quad\text{in}\,\Omega_{h}\backslash\Omega.\end{cases}\qquad\sigma=\begin{cases}1,\quad&\text{in}\,\Omega_{h}^{e,+},\\ -1,\quad&\text{in}\,\Omega_{h}^{e,-}.\end{cases}
Refer to caption
Figure 3: The true boundary Γ\Gamma (blue curve), the approximated boundary Γh\Gamma_{h} (red lines) and two typical regions Ωhe,+,Ωhe,\Omega_{h}^{e,+},\Omega_{h}^{e,-} bounded by Γ\Gamma and Γh\Gamma_{h}.
Lemma 4.8.

(Inf-Sup condition). When hh is sufficiently small, there exist positive constants β0\beta_{0} and β1\beta_{1} independent of hh such that, for qhQ0hq_{h}\in Q_{0h},

sup𝒗hVhbh1(𝒗h,qh)𝒗h0,hβ1qh0,Ωh,sup𝒗hVhbh0(𝒗h,qh)𝒗h0,hβ0qh0,Ωh.\displaystyle\sup_{\bm{v}_{h}\in V_{h}}\frac{b_{h1}(\bm{v}_{h},q_{h})}{\|\bm{v}_{h}\|_{0,h}}\geq\beta_{1}\|q_{h}\|_{0,\Omega_{h}},\qquad\sup_{\bm{v}_{h}\in V_{h}}\frac{b_{h0}(\bm{v}_{h},q_{h})}{\|\bm{v}_{h}\|_{0,h}}\geq\beta_{0}\|q_{h}\|_{0,\Omega_{h}}.
Proof.

Each function qhQ0hL02(Ωh)q_{h}\in Q_{0h}\subset L_{0}^{2}(\Omega_{h}) can be naturally extended to ΩΩh\Omega\cup\Omega_{h}, by filling the gap region Ω\Ωh\Omega\backslash\Omega_{h} with the same polynomial on neighboring mesh element. For simplicity, we still denote the extension by qhq_{h}. Define q¯h=qh1|Ω|ΩqhdxL02(Ω)\bar{q}_{h}=q_{h}-\frac{1}{|\Omega|}\int_{\Omega}q_{h}\mathop{}\mathrm{d}x\in L_{0}^{2}(\Omega). Since qhq_{h} is mean value free in Ωh\Omega_{h}, we have

q¯h=qh1|Ω|(ΩqhdxΩhqhdx)=qh+(1)σ1|Ω|ehbΩheqhdx.\displaystyle\bar{q}_{h}=q_{h}-\frac{1}{|\Omega|}\bigg{(}\int_{\Omega}q_{h}\mathop{}\mathrm{d}x-\int_{\Omega_{h}}q_{h}\mathop{}\mathrm{d}x\bigg{)}=q_{h}+(-1)^{\sigma}\frac{1}{|\Omega|}\sum_{e\in\mathcal{E}_{h}^{b}}\int_{\Omega_{h}^{e}}q_{h}\mathop{}\mathrm{d}x. (4.15)

Similar to [23], one gets

q¯h0,Ωqh0,Ωh,(q¯h,qh)Ωh=qh0,Ωh2.\displaystyle\|\bar{q}_{h}\|_{0,\Omega}\lesssim\|q_{h}\|_{0,\Omega_{h}},\qquad(\bar{q}_{h},q_{h})_{\Omega_{h}}=\|q_{h}\|_{0,\Omega_{h}}^{2}. (4.16)

It is well-known (see, e.g., [10]) that there exists a 𝒗𝑯01(Ω)\bm{v}\in\bm{H}_{0}^{1}(\Omega) satisfying div𝒗=q¯h\mathrm{div}\,\bm{v}=-\bar{q}_{h} on Ω\Omega and 𝒗1,Ωq¯h0,Ωqh0,Ωh\|\bm{v}\|_{1,\Omega}\lesssim\|\bar{q}_{h}\|_{0,\Omega}\lesssim\|q_{h}\|_{0,\Omega_{h}}, where the hidden constant in \lesssim may depend on the shape of Ω\Omega but not on hh. Extend 𝒗\bm{v} to 2\mathbb{R}^{2} by setting its value to be 0 outside Ω\Omega. The extension is still denoted by 𝒗\bm{v}, which shall not be mistaken for the extension E𝒗=𝒗EE\bm{v}=\bm{v}^{E} defined in Lemma 4.4. It is clear that, after the zero-extension, one has 𝒗H01(ΩΩ)\bm{v}\in H_{0}^{1}(\Omega\cup\Omega) and hence the interpolation Ih𝒗VhI_{h}\bm{v}\in V_{h} is well-defined. By the commutative property (4.2), the definition of IhI_{h} and (4.16), one has

bh1(Ih𝒗,qh)\displaystyle b_{h1}(I_{h}\bm{v},q_{h}) =(divIh𝒗,qh)Ωh+ehbIh𝒗𝐧h,qhe\displaystyle=-(\mathrm{div}\,I_{h}\bm{v},q_{h})_{\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}\langle I_{h}\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h},q_{h}\rangle_{e} (4.17)
=(Πk10div𝒗,qh)Ωh+ehb𝒗𝐧h,qhe\displaystyle=-(\Pi_{k-1}^{0}\mathrm{div}\,\bm{v},q_{h})_{\Omega_{h}}+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h},q_{h}\rangle_{e}
(div𝒗,qh)Ωhehbh1/2𝒗𝐧h0,eh1/2qh0,e\displaystyle\geq-(\mathrm{div}\,\bm{v},q_{h})_{\Omega_{h}}-\sum_{e\in\mathcal{E}_{h}^{b}}h^{-1/2}\|\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e}h^{1/2}\|q_{h}\|_{0,e}
(div𝒗,qh)Ωh(ehbh1𝒗𝐧h0,e2)1/2(ehbhqh0,e2)1/2\displaystyle\geq-(\mathrm{div}\,\bm{v},q_{h})_{\Omega_{h}}-\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}h^{-1}\|\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e}^{2}\bigg{)}^{1/2}\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}h\|q_{h}\|_{0,e}^{2}\bigg{)}^{1/2}
((div𝒗,qh)Ω+(div𝒗,qh)Ω\Ωh)\displaystyle\geq-\bigg{(}(\mathrm{div}\,\bm{v},q_{h})_{\Omega}+(\mathrm{div}\,\bm{v},q_{h})_{\Omega\backslash\Omega_{h}}\bigg{)}
(ehbh1𝒗𝐧h0,e2)1/2(ehbhqh0,e2)1/2.\displaystyle\quad-\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}h^{-1}\|\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e}^{2}\bigg{)}^{1/2}\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}h\|q_{h}\|_{0,e}^{2}\bigg{)}^{1/2}.

We then estimate the right-hand side of (4.17) one by one. Note that div𝒗=q¯h\mathrm{div}\,\bm{v}=-\bar{q}_{h} on Ω\Omega. By (4.16), (4.15) and Lemma 3.1, one has

(div𝒗,qh)Ω\displaystyle-(\mathrm{div}\,\bm{v},q_{h})_{\Omega} =(q¯h,qh)Ω=(q¯h,qh)Ωh(q¯h,qh)Ωh\Ω+(q¯h,qh)Ω\Ωh\displaystyle=(\bar{q}_{h},q_{h})_{\Omega}=(\bar{q}_{h},q_{h})_{\Omega_{h}}-(\bar{q}_{h},q_{h})_{\Omega_{h}\backslash\Omega}+(\bar{q}_{h},q_{h})_{\Omega\backslash\Omega_{h}}
=qh0,Ωh2(1)σehbΩheq¯hqhdx\displaystyle=\|q_{h}\|_{0,\Omega_{h}}^{2}-(-1)^{\sigma}\sum_{e\in\mathcal{E}_{h}^{b}}\int_{\Omega_{h}^{e}}\bar{q}_{h}\,q_{h}\mathop{}\mathrm{d}x
qh0,Ωh2ehbqh0,Ωhe21|Ω|(ehbΩheqhdx)2\displaystyle\geq\|q_{h}\|_{0,\Omega_{h}}^{2}-\sum_{e\in\mathcal{E}_{h}^{b}}\|q_{h}\|_{0,\Omega_{h}^{e}}^{2}-\frac{1}{|\Omega|}\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}\int_{\Omega_{h}^{e}}q_{h}\mathop{}\mathrm{d}x\bigg{)}^{2}
qh0,Ωh2hqh0,Ωh21|Ω|ehb|Ωhe|qh0,Ωhe2\displaystyle\geq\|q_{h}\|_{0,\Omega_{h}}^{2}-h\|q_{h}\|_{0,\Omega_{h}}^{2}-\frac{1}{|\Omega|}\sum_{e\in\mathcal{E}_{h}^{b}}|\Omega_{h}^{e}|\|q_{h}\|_{0,\Omega_{h}^{e}}^{2}
(1O(h))qh0,Ωh2.\displaystyle\geq(1-O(h))\|q_{h}\|_{0,\Omega_{h}}^{2}.

Using lemmas 4.4 and 3.1, we have

(div𝒗,qh)Ω\Ωh\displaystyle(\mathrm{div}\,\bm{v},q_{h})_{\Omega\backslash\Omega_{h}} ehbdiv𝒗0,Ωheqh0,Ωhediv𝒗0,Ω(ehbqh0,Ωhe2)1/2\displaystyle\leq\sum_{e\in\mathcal{E}_{h}^{b}}\|\mathrm{div}\,\bm{v}\|_{0,\Omega_{h}^{e}}\|q_{h}\|_{0,\Omega_{h}^{e}}\leq\|\mathrm{div}\,\bm{v}\|_{0,\Omega}\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}\|q_{h}\|_{0,\Omega_{h}^{e}}^{2}\bigg{)}^{1/2}
h1/2𝒗1,Ωqh0,Ωhh1/2qh0,Ωh2.\displaystyle\lesssim h^{1/2}\|\bm{v}\|_{1,\Omega}\|q_{h}\|_{0,\Omega_{h}}\lesssim h^{1/2}\|q_{h}\|_{0,\Omega_{h}}^{2}.

By the trace-inverse inequality in Lemma 2.3, one gets

(ehbhqh0,e2)1/2qh0,Ωh.\displaystyle\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}h\|q_{h}\|_{0,e}^{2}\bigg{)}^{1/2}\lesssim\|q_{h}\|_{0,\Omega_{h}}.

Note that 𝒗|Γ=0\bm{v}|_{\Gamma}=0 and 𝒗=0\bm{v}=0 on Ωh\Ω\Omega_{h}\backslash\Omega. By (3.11) and Lemma 4.4, one has

(ehbh1𝒗𝐧h0,e2)1/2\displaystyle\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}h^{-1}\|\bm{v}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e}^{2}\bigg{)}^{1/2} (ehbh1𝒗0,e2)1/2h1/2(ehb𝒗1,Ωhe2)1/2\displaystyle\lesssim\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}h^{-1}\|\bm{v}\|_{0,e}^{2}\bigg{)}^{1/2}\lesssim h^{1/2}\bigg{(}\sum_{e\in\mathcal{E}_{h}^{b}}\|\bm{v}\|_{1,\Omega_{h}^{e}}^{2}\bigg{)}^{1/2}
h1/2𝒗1,Ωh1/2qh0,Ωh.\displaystyle\lesssim h^{1/2}\|\bm{v}\|_{1,\Omega}\lesssim h^{1/2}\|q_{h}\|_{0,\Omega_{h}}.

Combining the above, for hh sufficiently small, one gets

bh1(Ih𝒗,qh)(1O(h1/2))qh0,Ωh2qh0,Ωh2.\displaystyle b_{h1}(I_{h}\bm{v},q_{h})\geq(1-O(h^{1/2}))\|q_{h}\|_{0,\Omega_{h}}^{2}\gtrsim\|q_{h}\|_{0,\Omega_{h}}^{2}.

A similar argument gives

bh0(Ih𝒗,qh)qh0,Ωh2.b_{h0}(I_{h}\bm{v},q_{h})\gtrsim\|q_{h}\|_{0,\Omega_{h}}^{2}.

Then, for i=0,1i=0,1, one obtains

sup𝒗hVhbhi(𝒗h,qh)𝒗h0,hbhi(Ih𝒗,qh)Ih𝒗0,hqh0,Ωh2Ih𝒗0,h,\displaystyle\sup_{\bm{v}_{h}\in V_{h}}\frac{b_{hi}(\bm{v}_{h},q_{h})}{\|\bm{v}_{h}\|_{0,h}}\geq\frac{b_{hi}(I_{h}\bm{v},q_{h})}{\|I_{h}\bm{v}\|_{0,h}}\gtrsim\frac{\|q_{h}\|_{0,\Omega_{h}}^{2}}{\|I_{h}\bm{v}\|_{0,h}},

where Ih𝒗I_{h}\bm{v} is the interpolation of the zero-extension of 𝒗𝑯01(Ω)\bm{v}\in\bm{H}_{0}^{1}(\Omega), as explained before.

Finally, we prove Ih𝒗0,hqh0,Ωh\|I_{h}\bm{v}\|_{0,h}\lesssim\|q_{h}\|_{0,\Omega_{h}}, which combined with the above inequality will give the desired inf-sup condition. By Lemma 4.3 and the property (4.2), we have

Ih𝒗0,K𝒗1,K,\|I_{h}\bm{v}\|_{0,K}\lesssim\|\bm{v}\|_{1,K},

and

divIh𝒗0,K=Πk10,Kdiv𝒗0,Kdiv𝒗0,K|𝒗|1,K.\|\mathrm{div}\,I_{h}\bm{v}\|_{0,K}=\|\Pi_{k-1}^{0,K}\mathrm{div}\,\bm{v}\|_{0,K}\leq\|\mathrm{div}\,\bm{v}\|_{0,K}\lesssim|\bm{v}|_{1,K}.

By Lemma 4.5, 4.3 and definitions of TmT^{m} and T1mT_{1}^{m}, it follows that

hK1/2Tm(Ih𝒗)𝐧~0,e\displaystyle\|h_{K}^{-1/2}T^{m}(I_{h}\bm{v})\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e} hK1/2Ih𝒗𝐧~0,e+hK1/2T1m(Ih𝒗)𝐧~0,e\displaystyle\leq\|h_{K}^{-1/2}I_{h}\bm{v}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}+\|h_{K}^{-1/2}T_{1}^{m}(I_{h}\bm{v})\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}
hK1/2Ih𝒗𝐧~0,e+Ih𝒗0,K\displaystyle\lesssim\|h_{K}^{-1/2}I_{h}\bm{v}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}+\|I_{h}\bm{v}\|_{0,K}
hK1/2Ih𝒗0,e+𝒗1,K.\displaystyle\lesssim\|h_{K}^{-1/2}I_{h}\bm{v}\|_{0,e}+\|\bm{v}\|_{1,K}.

By (3.8) and Lemma 3.1 and Inequality (4.4), it holds

hK1/2Ih𝒗0,e\displaystyle\|h_{K}^{-1/2}I_{h}\bm{v}\|_{0,e} h1/2(Ih𝒗0,e~+h|Ih𝒗|1,Ωhe)\displaystyle\lesssim h^{-1/2}(\|I_{h}\bm{v}\|_{0,\widetilde{e}}+h|I_{h}\bm{v}|_{1,\Omega_{h}^{e}})
h1/2(Ih𝒗0,e~+h3/2|𝒗|1,K).\displaystyle\lesssim h^{-1/2}(\|I_{h}\bm{v}\|_{0,\widetilde{e}}+h^{3/2}|\bm{v}|_{1,K}).

Note that 𝒗𝑯01(Ω)\bm{v}\in\bm{H}_{0}^{1}(\Omega), using lemmas 2.1, 3.1 and inequalities (3.9), (4.6), one gets

hK1/2Ih𝒗0,e~\displaystyle h_{K}^{-1/2}\|I_{h}\bm{v}\|_{0,\widetilde{e}} =hK1/2Ih𝒗𝒗0,e~\displaystyle=h_{K}^{-1/2}\|I_{h}\bm{v}-\bm{v}\|_{0,\widetilde{e}}
h1/2(Ih𝒗𝒗0,e+h|Ih𝒗𝒗|1,Ωhe)\displaystyle\lesssim h^{-1/2}(\|I_{h}\bm{v}-\bm{v}\|_{0,e}+h|I_{h}\bm{v}-\bm{v}|_{1,\Omega_{h}^{e}})
h1Ih𝒗𝒗0,K+|Ih𝒗𝒗|1,K+h1/2|Ih𝒗|1,Ωhe+h1/2|𝒗|1,Ωhe\displaystyle\lesssim h^{-1}\|I_{h}\bm{v}-\bm{v}\|_{0,K}+|I_{h}\bm{v}-\bm{v}|_{1,K}+h^{1/2}|I_{h}\bm{v}|_{1,\Omega_{h}^{e}}+h^{1/2}|\bm{v}|_{1,\Omega_{h}^{e}}
|𝒗|1,KΩhe.\displaystyle\lesssim|\bm{v}|_{1,K\cup\Omega_{h}^{e}}.

Combining the above, we obtain Ih𝒗0,h𝒗1,Ωqh0,Ωh\|I_{h}\bm{v}\|_{0,h}\lesssim\|\bm{v}\|_{1,\Omega}\lesssim\|q_{h}\|_{0,\Omega_{h}}. This completes the proof of the lemma. ∎

From Theorem 3.1 in [25] (also see [6]), lemmas 4.7 and 4.8 immediately imply that

Theorem 4.1.

The discrete problem (4.12) is well-posed, i.e. it admits a unique solution and

𝝈h,ζhHsup(𝒗h,qh)Vh×Q0hBh((𝝈h,ζh),(𝒗h,qh))𝒗h,qhH,(𝝈h,ζh)Vh×Q0h,\displaystyle\|\bm{\sigma}_{h},\zeta_{h}\|_{H}\lesssim\sup_{(\bm{v}_{h},q_{h})\in V_{h}\times Q_{0h}}\frac{B_{h}((\bm{\sigma}_{h},\zeta_{h}),(\bm{v}_{h},q_{h}))}{\|\bm{v}_{h},q_{h}\|_{H}},\qquad\forall\,(\bm{\sigma}_{h},\zeta_{h})\in V_{h}\times Q_{0h}, (4.18)

where

Bh((𝝈h,ζh),(𝒗h,qh))\displaystyle B_{h}((\bm{\sigma}_{h},\zeta_{h}),(\bm{v}_{h},q_{h})) :=ah(𝝈h,𝒗h)+bh1(𝒗h,ζh)+bh0(𝝈h,qh),\displaystyle:=a_{h}(\bm{\sigma}_{h},\bm{v}_{h})+b_{h1}(\bm{v}_{h},\zeta_{h})+b_{h0}(\bm{\sigma}_{h},q_{h}),
𝝈h,ζhH\displaystyle\|\bm{\sigma}_{h},\zeta_{h}\|_{H} :=(𝝈h0,h2+ζh0,Ωh2)1/2.\displaystyle:=(\|\bm{\sigma}_{h}\|_{0,h}^{2}+\|\zeta_{h}\|_{0,\Omega_{h}}^{2})^{1/2}.

4.3 Compatibility of (4.12) and an equivalent discrete form

Recall that a compatibility condition (3.2) is needed for the well-posedness of the continuous problem (3.1), as well as its weak formulation (3.3). We point out that the compatibility mechanism for the discrete problem works differently. In fact, the essential boundary condition 𝒖𝐧=gN\bm{u}\cdot\bm{\mathop{}\mathrm{n}}=g_{N} on Γ\Gamma has been weakened in (4.12), and thus no relation is required between g~N\widetilde{g}_{N} and fEf^{E}. In Section 4.2, we have proved that the discrete system (4.12) admits a unique solution, and hence is “compatible”.

However, in practical implementation, the space Q0hQ_{0h} is not commonly used. It is more convenient to construct a set of basis for QhQ_{h} than for Q0hQ_{0h}. Hence we introduce an equivalent discrete problem: Find (𝐮h,ph)Vh×Qh(\bm{u}_{h},p_{h})\in V_{h}\times Q_{h} such that

{ah(𝒖h,𝒗h)+bh1(𝒗h,ph)=lh(𝒗h),𝒗hVh,bh0(𝒖h,qh)+ehb𝒖h𝐧h,q¯he=(fEfE¯,qh)Ωh,qhQh,\left\{\begin{aligned} a_{h}(\bm{u}_{h},\bm{v}_{h})+b_{h1}(\bm{v}_{h},p_{h})&=l_{h}(\bm{v}_{h}),\quad&\forall&\bm{v}_{h}\in V_{h},\\ b_{h0}(\bm{u}_{h},q_{h})+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{u}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h},\bar{q}_{h}\rangle_{e}&=-(f^{E}-\overline{f^{E}},q_{h})_{\Omega_{h}},\quad&\forall&q_{h}\in Q_{h},\end{aligned}\right. (4.19)

where q¯h=1|Ωh|Ωhqhdx\bar{q}_{h}=\frac{1}{|\Omega_{h}|}\int_{\Omega_{h}}q_{h}\mathop{}\mathrm{d}x and fE¯=1|Ωh|ΩhfEdx\overline{f^{E}}=\frac{1}{|\Omega_{h}|}\int_{\Omega_{h}}f^{E}\mathop{}\mathrm{d}x.

Lemma 4.9.

The systems (4.12) and (4.19) are equivalent in the sense that they admit the same solution in Vh×Q0hV_{h}\times Q_{0h}.

Proof.

The space QhQ_{h} can be decomposed into Q0hQ_{0h}\oplus\mathbb{R}, and the decomposition is orthogonal in the L2(Ωh)L^{2}(\Omega_{h}) norm. Using integration by parts, it is obvious that bh1(𝒗h,1)=0b_{h1}(\bm{v}_{h},1)=0 for all 𝒗hVh\bm{v}_{h}\in V_{h}. Since qhq¯hQ0hq_{h}-\bar{q}_{h}\in Q_{0h}, the second equation of (4.12) implies that, for all qhQhq_{h}\in Q_{h},

bh0(𝒖h,qhq¯h)\displaystyle b_{h0}(\bm{u}_{h},q_{h}-\bar{q}_{h}) =(fE,qhq¯h)Ωh,\displaystyle=-(f^{E},q_{h}-\bar{q}_{h})_{\Omega_{h}},
bh0(𝒖h,qh)+ehb𝒖h𝐧h,q¯he\displaystyle\Rightarrow\quad b_{h0}(\bm{u}_{h},q_{h})+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{u}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h},\bar{q}_{h}\rangle_{e} =(fE,qhq¯h)Ωh,\displaystyle=-(f^{E},q_{h}-\bar{q}_{h})_{\Omega_{h}},

which is just the second equation of (4.19). ∎

As long as (4.12) is well-posed, problem (4.19) is solvable, i.e., the underlying discrete linear system is “compatible”. However, its solution is not unique since (𝒖h,ph+constant)(\bm{u}_{h},\,p_{h}+constant) is also a solution. There are various ways to deal with this case, for example, using a Krylov subspace iterative solver, or adding a constraint to the linear system to ensure that the solution stays in Vh×Q0hV_{h}\times Q_{0h}.

5 Error analysis

We first obtain an error equation by subtracting equations (4.9)-(4.10) from the discrete problem (4.12)

ER:=\displaystyle E_{R}:= ah(𝒖h𝒖E,𝒗h)+bh1(𝒗h,phpE)+bh0(𝒖h𝒖E,qh)\displaystyle~{}a_{h}(\bm{u}_{h}-\bm{u}^{E},\bm{v}_{h})+b_{h1}(\bm{v}_{h},p_{h}-p^{E})+b_{h0}(\bm{u}_{h}-\bm{u}^{E},q_{h}) (5.1)
=\displaystyle= (𝒖E+pE,𝒗h)Ωh(div𝒖EfE,div𝒗h)Ωh+(div𝒖EfE,qh)Ωh\displaystyle~{}-(\bm{u}^{E}+\nabla p^{E},\bm{v}_{h})_{\Omega_{h}}-(\mathrm{div}\,\bm{u}^{E}-f^{E},\mathrm{div}\,\bm{v}_{h})_{\Omega_{h}}+(\mathrm{div}\,\bm{u}^{E}-f^{E},q_{h})_{\Omega_{h}}
+ehbhK1g~NTm𝒖E𝐧~,Tm𝒗h𝐧~e.\displaystyle\quad+\sum_{e\in\mathcal{E}_{h}^{b}}h_{K}^{-1}\langle\widetilde{g}_{N}-T^{m}\bm{u}^{E}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}},T^{m}\bm{v}_{h}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\rangle_{e}.
Lemma 5.1.

(cf. [12]). Assume that 𝐮𝐇max{l+1,m+2}(Ω),pHl+1(Ω)\bm{u}\in\bm{H}^{\max\{l+1,m+2\}}(\Omega),\,p\in H^{l+1}(\Omega) and fHl(Ω)f\in H^{l}(\Omega) satisfy Equation (3.1), then

𝒖E+pE0,Ωh\displaystyle\|\bm{u}^{E}+\nabla p^{E}\|_{0,\Omega_{h}} δlDl(𝒖E+pE)0,Ωh\Ω,\displaystyle\lesssim\delta^{l}\|D^{l}(\bm{u}^{E}+\nabla p^{E})\|_{0,\Omega_{h}\backslash\Omega}, (5.2)
div𝒖EfE0,Ωh\displaystyle\|\mathrm{div}\,\bm{u}^{E}-f^{E}\|_{0,\Omega_{h}} δlDl(div𝒖EfE)0,Ωh\Ω,\displaystyle\lesssim\delta^{l}\|D^{l}(\mathrm{div}\,\bm{u}^{E}-f^{E})\|_{0,\Omega_{h}\backslash\Omega}, (5.3)
ehbhK1g~NTm𝒖E𝐧~0,e2\displaystyle\sum_{e\in\mathcal{E}_{h}^{b}}h_{K}^{-1}\|\widetilde{g}_{N}-T^{m}\bm{u}^{E}\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e}^{2} δ2m+2h1Dm+2𝒖0,Ω2.\displaystyle\lesssim\delta^{2m+2}h^{-1}\|D^{m+2}\bm{u}\|_{0,\Omega}^{2}. (5.4)
Proof.

Inequalities (5.2) and (5.3) follow immediately from (3.14) in [12]. The estimate of the residual (5.4) follows from (3.12) in [12] and Lemma 4.4. ∎

Combining (5.1) and using Lemma 5.1, one gets an estimate of the consistency error:

Lemma 5.2.

(Consistency Error). Assume that 𝐮𝐇max{l+1,m+2}(Ω),pHl+1(Ω)\bm{u}\in\bm{H}^{\max\{l+1,m+2\}}(\Omega),\,p\in H^{l+1}(\Omega) and fHl(Ω)f\in H^{l}(\Omega) satisfy Equation (3.1), we have

|ER|[δl(Dl(𝒖E+pE)0,Ωh\Ω\displaystyle|E_{R}|\lesssim\bigg{[}\delta^{l}\bigg{(}\|D^{l}(\bm{u}^{E}+\nabla p^{E})\|_{0,\Omega_{h}\backslash\Omega} +Dl(div𝒖EfE)0,Ωh\Ω)\displaystyle+\|D^{l}(\mathrm{div}\,\bm{u}^{E}-f^{E})\|_{0,\Omega_{h}\backslash\Omega}\bigg{)}
+δm+1h1/2Dm+2𝒖0,Ω]𝒗h,qhH.\displaystyle\qquad+\delta^{m+1}h^{-1/2}\|D^{m+2}\bm{u}\|_{0,\Omega}\bigg{]}\|\bm{v}_{h},q_{h}\|_{H}.
Lemma 5.3.

(Interpolation Error). Let (𝐯,p)𝐇r+1(Ω)×Ht(Ω)(\bm{v},p)\in\bm{H}^{r+1}(\Omega)\times H^{t}(\Omega) with max{m,1}r+1k+1\max\{m,1\}\leq r+1\leq k+1 and 0tk0\leq t\leq k. Let 𝐯I=Ih𝐯EVh\bm{v}_{I}=I_{h}\bm{v}^{E}\in V_{h} be the interpolation of 𝐯E\bm{v}^{E}, and pIQ0hp_{I}\in Q_{0h} be the L2L^{2} orthogonal projection of pEp^{E}. Then

𝒗E𝒗I0,h+pEpI0,Ωhhmin{r,t,k}(|𝒗|r,Ω+|𝒗|r+1,Ω+|p|t,Ω).\displaystyle\|\bm{v}^{E}-\bm{v}_{I}\|_{0,h}+\|p^{E}-p_{I}\|_{0,\Omega_{h}}\lesssim h^{\min\{r,t,k\}}(|\bm{v}|_{r,\Omega}+|\bm{v}|_{r+1,\Omega}+|p|_{t,\Omega}).
Proof.

By Lemmas 4.3 and 4.4, we obtain

𝒗E𝒗I0,Ωh\displaystyle\|\bm{v}^{E}-\bm{v}_{I}\|_{0,\Omega_{h}} hr|𝒗E|r,Ωhhr|𝒗|r,Ω,\displaystyle\lesssim h^{r}|\bm{v}^{E}|_{r,\Omega_{h}}\lesssim h^{r}|\bm{v}|_{r,\Omega},
div(𝒗E𝒗I)0,Ωh\displaystyle\|\mathrm{div}(\bm{v}^{E}-\bm{v}_{I})\|_{0,\Omega_{h}} hr|div𝒗E|r,Ωhhr|𝒗E|r+1,Ωhhr|𝒗|r+1,Ω,\displaystyle\lesssim h^{r}|\mathrm{div}\,\bm{v}^{E}|_{r,\Omega_{h}}\lesssim h^{r}|\bm{v}^{E}|_{r+1,\Omega_{h}}\lesssim h^{r}|\bm{v}|_{r+1,\Omega},
pEpI0,Ωh\displaystyle\|p^{E}-p_{I}\|_{0,\Omega_{h}} ht|pE|t,Ωhht|p|t,Ω,\displaystyle\lesssim h^{t}|p^{E}|_{t,\Omega_{h}}\lesssim h^{t}|p|_{t,\Omega},

By Lemma 2.1, inequality (3.5) and interpolation error (4.6), one has

hK1/2Tm(𝒗E𝒗I)𝐧~0,e\displaystyle\|h_{K}^{-1/2}T^{m}(\bm{v}^{E}-\bm{v}_{I})\cdot\widetilde{\bm{\mathop{}\mathrm{n}}}\|_{0,e} j=0mδhjhK1/2(hK1/2|𝒗E𝒗I|j,K+hK1/2|𝒗E𝒗I|j+1,K)\displaystyle\lesssim\sum_{j=0}^{m}\delta_{h}^{j}h_{K}^{-1/2}(h_{K}^{-1/2}|\bm{v}^{E}-\bm{v}_{I}|_{j,K}+h_{K}^{1/2}|\bm{v}^{E}-\bm{v}_{I}|_{j+1,K})
j=0mδhjhK1hr+1j|𝒗E|r+1,K\displaystyle\lesssim\sum_{j=0}^{m}\delta_{h}^{j}h_{K}^{-1}h^{r+1-j}|\bm{v}^{E}|_{r+1,K}
j=0m(δhh)jhrj|𝒗E|r+1,K\displaystyle\lesssim\sum_{j=0}^{m}\bigg{(}\frac{\delta_{h}}{h}\bigg{)}^{j}h^{r-j}|\bm{v}^{E}|_{r+1,K}
hr|𝒗E|r+1,K.\displaystyle\lesssim h^{r}|\bm{v}^{E}|_{r+1,K}.

Combining the above and using the definitions of norms, we complete the proof. ∎

Lemma 5.4.

Let (𝐮,p)𝐇max{r+1,l+1,m+2}(Ω)×Hmax{t,l+1}(Ω)(\bm{u},p)\in\bm{H}^{\max\{r+1,l+1,m+2\}}(\Omega)\times H^{\max\{t,l+1\}}(\Omega) be the solution to Problem (3.1), fHl(Ω)f\in H^{l}(\Omega), and (𝐮h,ph)Vh×Q0h(\bm{u}_{h},p_{h})\in V_{h}\times Q_{0h} be the discrete solution of (4.12). Then

𝒖h𝒖I0,h+phpI0,Ωh\displaystyle\|\bm{u}_{h}-\bm{u}_{I}\|_{0,h}+\|p_{h}-p_{I}\|_{0,\Omega_{h}}
\displaystyle\lesssim hmin{r,t,k}(|𝒖|r,Ω+|𝒖|r+1,Ω+|p|t,Ω)+δm+1h1/2Dm+2𝒖0,Ω\displaystyle~{}h^{\min\{r,t,k\}}(|\bm{u}|_{r,\Omega}+|\bm{u}|_{r+1,\Omega}+|p|_{t,\Omega})+\delta^{m+1}h^{-1/2}\|D^{m+2}\bm{u}\|_{0,\Omega}
+δl(Dl(𝒖E+pE)0,Ωh\Ω+Dl(div𝒖EfE)0,Ωh\Ω).\displaystyle\quad+\delta^{l}\bigg{(}\|D^{l}(\bm{u}^{E}+\nabla p^{E})\|_{0,\Omega_{h}\backslash\Omega}+\|D^{l}(\mathrm{div}\,\bm{u}^{E}-f^{E})\|_{0,\Omega_{h}\backslash\Omega}\bigg{)}.

where 𝐮I=Ih𝐮EVh\bm{u}_{I}=I_{h}\bm{u}^{E}\in V_{h} is the interpolation of 𝐮E\bm{u}^{E}, and pIQ0hp_{I}\in Q_{0h} is the L2L^{2} orthogonal projection of pEp^{E}

Proof.

From (4.18), one has

𝒖h𝒖I,phpIHsup(𝒗h,qh)Vh×Q0hBh((𝒖h𝒖I,phpI),(𝒗h,qh))𝒗h,qhH.\displaystyle\|\bm{u}_{h}-\bm{u}_{I},p_{h}-p_{I}\|_{H}\lesssim\sup_{(\bm{v}_{h},q_{h})\in V_{h}\times Q_{0h}}\frac{B_{h}((\bm{u}_{h}-\bm{u}_{I},p_{h}-p_{I}),(\bm{v}_{h},q_{h}))}{\|\bm{v}_{h},q_{h}\|_{H}}.

Note that

Bh((𝒖h𝒖I,phpI),(𝒗h,qh))=\displaystyle B_{h}((\bm{u}_{h}-\bm{u}_{I},p_{h}-p_{I}),(\bm{v}_{h},q_{h}))= ah(𝒖h𝒖E,𝒗h)+bh1(𝒗h,phpE)+bh0(𝒖h𝒖E,qh)\displaystyle~{}a_{h}(\bm{u}_{h}-\bm{u}^{E},\bm{v}_{h})+b_{h1}(\bm{v}_{h},p_{h}-p^{E})+b_{h0}(\bm{u}_{h}-\bm{u}^{E},q_{h})
+ah(𝒖E𝒖I,𝒗h)+bh1(𝒗h,pEpI)+bh0(𝒖E𝒖I,qh)\displaystyle+a_{h}(\bm{u}^{E}-\bm{u}_{I},\bm{v}_{h})+b_{h1}(\bm{v}_{h},p^{E}-p_{I})+b_{h0}(\bm{u}^{E}-\bm{u}_{I},q_{h})
=:\displaystyle=: ER+Eh.\displaystyle~{}E_{R}+E_{h}.

By the definition of 𝒖I\bm{u}_{I} and pIp_{I}, one gets

bh0(𝒖E𝒖I,qh)\displaystyle b_{h0}(\bm{u}^{E}-\bm{u}_{I},q_{h}) =(div(𝒖I𝒖E),qh)Ωh\displaystyle=(\mathrm{div}\,(\bm{u}_{I}-\bm{u}^{E}),q_{h})_{\Omega_{h}}
=K𝒯h(𝒖E𝒖I,qh)KK𝒯h(𝒖E𝒖I)𝐧h,qhK=0,\displaystyle=\sum_{K\in\mathcal{T}_{h}}(\bm{u}^{E}-\bm{u}_{I},\nabla q_{h})_{K}-\sum_{K\in\mathcal{T}_{h}}\langle(\bm{u}^{E}-\bm{u}_{I})\cdot\bm{\mathop{}\mathrm{n}}_{h},q_{h}\rangle_{\partial K}=0,

and

bh1(𝒗h,pEpI)\displaystyle b_{h1}(\bm{v}_{h},p^{E}-p_{I}) =bh0(𝒗h,pEpI)+ehb𝒗h𝐧h,pEpIe\displaystyle=b_{h0}(\bm{v}_{h},p^{E}-p_{I})+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h},p^{E}-p_{I}\rangle_{e}
=ehb𝒗h𝐧h,pEpIe.\displaystyle=\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h},p^{E}-p_{I}\rangle_{e}.

Therefore

Eh\displaystyle E_{h} =ah(𝒖E𝒖I,𝒗h)+bh1(𝒗h,pEpI)+bh0(𝒖E𝒖I,qh)\displaystyle=a_{h}(\bm{u}^{E}-\bm{u}_{I},\bm{v}_{h})+b_{h1}(\bm{v}_{h},p^{E}-p_{I})+b_{h0}(\bm{u}^{E}-\bm{u}_{I},q_{h})
=ah(𝒖E𝒖I,𝒗h)+ehb𝒗h𝐧h,pEpIe.\displaystyle=a_{h}(\bm{u}^{E}-\bm{u}_{I},\bm{v}_{h})+\sum_{e\in\mathcal{E}_{h}^{b}}\langle\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h},p^{E}-p_{I}\rangle_{e}.

By the Cauchy-Schwarz inequality, lemmas 5.3, 4.6, 2.1, 4.4 and Inequality (4.5), one gets

Eh\displaystyle E_{h} 𝒖E𝒖I0,h𝒗h0,h+ehbhK1/2𝒗h𝐧h0,ehK1/2pEpI0,e\displaystyle\leq\|\bm{u}^{E}-\bm{u}_{I}\|_{0,h}\|\bm{v}_{h}\|_{0,h}+\sum_{e\in\mathcal{E}_{h}^{b}}h_{K}^{-1/2}\|\bm{v}_{h}\cdot\bm{\mathop{}\mathrm{n}}_{h}\|_{0,e}h_{K}^{1/2}\|p^{E}-p_{I}\|_{0,e}
(𝒖E𝒖I0,h+pEpI0,Ωh+h|pEpI|1,Ωh)𝒗h0,h\displaystyle\lesssim(\|\bm{u}^{E}-\bm{u}_{I}\|_{0,h}+\|p^{E}-p_{I}\|_{0,\Omega_{h}}+h|p^{E}-p_{I}|_{1,\Omega_{h}})\|\bm{v}_{h}\|_{0,h}
hmin{r,t,k}(|𝒖|r,Ω+|𝒖|r+1,Ω+|p|t,Ω)𝒗h0,h.\displaystyle\lesssim h^{\min\{r,t,k\}}(|\bm{u}|_{r,\Omega}+|\bm{u}|_{r+1,\Omega}+|p|_{t,\Omega})\|\bm{v}_{h}\|_{0,h}.

Combining the above estimate of EhE_{h} above and the estimate of ERE_{R} in Lemma 5.2, we obtain the proof of the lemma. ∎

Using the above lemmas and the triangle inequality, we get the main theorem of this section:

Theorem 5.1.

Under the same assumption of Lemma 5.4, one has

𝒖E𝒖h0,h+pEph0,Ωh\displaystyle\|\bm{u}^{E}-\bm{u}_{h}\|_{0,h}+\|p^{E}-p_{h}\|_{0,\Omega_{h}} (5.5)
\displaystyle\lesssim hmin{r,t,k}(|𝒖|r,Ω+|𝒖|r+1,Ω+|p|t,Ω)+δm+1h1/2Dm+2𝒖0,Ω\displaystyle~{}h^{\min\{r,t,k\}}(|\bm{u}|_{r,\Omega}+|\bm{u}|_{r+1,\Omega}+|p|_{t,\Omega})+\delta^{m+1}h^{-1/2}\|D^{m+2}\bm{u}\|_{0,\Omega}
+δl(Dl(𝒖E+pE)0,Ωh\Ω+Dl(div𝒖EfE)0,Ωh\Ω).\displaystyle\qquad+\delta^{l}\bigg{(}\|D^{l}(\bm{u}^{E}+\nabla p^{E})\|_{0,\Omega_{h}\backslash\Omega}+\|D^{l}(\mathrm{div}\,\bm{u}^{E}-f^{E})\|_{0,\Omega_{h}\backslash\Omega}\bigg{)}.

Moreover, the terms Dl(𝐮E+pE)0,Ωh\Ω\|D^{l}(\bm{u}^{E}+\nabla p^{E})\|_{0,\Omega_{h}\backslash\Omega} and Dl(div𝐮EfE)0,Ωh\Ω\|D^{l}(\mathrm{div}\,\bm{u}^{E}-f^{E})\|_{0,\Omega_{h}\backslash\Omega} vanish when Ω\Omega is convex.

Remark 5.1.

(An implementation trick). When m=km=k, one has Tm𝐯h=𝐯hT^{m}\bm{v}_{h}=\bm{v}_{h} for all 𝐯hVh\bm{v}_{h}\in V_{h}, which greatly simplifies the implementation as one no longer needs to compute the Taylor expansion.

For a given polynomial order k1k\geq 1, we want to find when the proposed discretization (4.12) reaches the optimal convergence rate O(hk)O(h^{k}). Suppose that Inequality (3.5) holds, that is, δh2\delta\lesssim h^{2}. Then Theorem 5.1 indicates that one has to have

rk,tk,2lk,2m+3/2k,r\geq k,\quad t\geq k,\quad 2l\geq k,\quad 2m+3/2\geq k,

to obtain O(hk)O(h^{k}) convergence. By taking r=t=kr=t=k and l=max{1,k1}l=\max\{1,k-1\}, we get the following corollary:

Corollary 5.1.

Assume (3.5) holds, max{0,k/23/4}mk\max\{0,k/2-3/4\}\leq m\leq k, and l=max{1,k1}l=\max\{1,k-1\}. Let (𝐮,p)𝐇k+2(Ω)×Hl+1(Ω)(\bm{u},p)\in\bm{H}^{k+2}(\Omega)\times H^{l+1}(\Omega) be the solution to Problem (3.1), fHl(Ω)f\in H^{l}(\Omega), and (𝐮h,ph)Vh×Q0h(\bm{u}_{h},p_{h})\in V_{h}\times Q_{0h} be the discrete solution of (4.12), then

𝒖E𝒖h0,h+pEph0,Ωhhk(𝒖k+2,Ω+pl+1,Ω+fl,Ω).\|\bm{u}^{E}-\bm{u}_{h}\|_{0,h}+\|p^{E}-p_{h}\|_{0,\Omega_{h}}\lesssim h^{k}\left(\|\bm{u}\|_{k+2,\Omega}+\|p\|_{l+1,\Omega}+\|f\|_{l,\Omega}\right). (5.6)

When Ω\Omega is convex, the above result holds for l=k1l=k-1.

6 Numerical experiment

In this section, we present numerical results for problem (3.1) on two types of curved domains, as shown in Fig. 4. Body-fitted, unstructured triangular meshes are used in the computation. We always set m=km=k, which simplifies the implementation according to Remark 5.1. For convenience, denote

Eu,p=𝒖E𝒖h0,h+pEph0,Ωh.E_{u,p}=\|\bm{u}^{E}-\bm{u}_{h}\|_{0,h}+\|p^{E}-p_{h}\|_{0,\Omega_{h}}.

If no boundary correction technique is used, then 0,h\|\cdot\|_{0,h} is just the H(div,Ωh)H(\textrm{div},\Omega_{h}) norm.

Refer to caption
Refer to caption
Figure 4: (a). The mesh with h=1/2h=1/2 on circular domain. (b). The mesh with h=1/4h=1/4 on ring domain.

𝐄𝐱𝐚𝐦𝐩𝐥𝐞𝟏.𝐂𝐢𝐫𝐜𝐮𝐥𝐚𝐫𝐝𝐨𝐦𝐚𝐢𝐧{\bf{Example~{}1.~{}Circular~{}domain}}.

The domain is a unit disk {(x,y):x2+y2<1}\{(x,y):x^{2}+y^{2}<1\}. The exact solution is

𝒖(x,y)=(3x2y2+32xy),p(x,y)=3x+x(x2+y2),\displaystyle\bm{u}(x,y)=\dbinom{-3x^{2}-y^{2}+3}{-2xy},\qquad p(x,y)=-3x+x(x^{2}+y^{2}),

which satisfies a homogeneous Neumann boundary condition 𝒖𝐧=0\bm{u}\cdot\bm{\mathop{}\mathrm{n}}=0 on Γ\Gamma, and Ωpdx=0\int_{\Omega}p\mathop{}\mathrm{d}x=0. Note that 𝒖\bm{u}, pp and ff are smooth, i.e. they satisfy the regularity requirement of Cor. 5.1. We solve Example 11 on meshes illustrated in Fig. 4 (a), for various size of hh.

We first test Example 11 using the primal mixed formulation [7] with BDMkBDM_{k} discretization, without boundary correction. A homogeneous Neumann (essential) boundary condition is imposed on Γh\Gamma_{h}. We report Eu,pE_{u,p} in Tab. 1, which only reaches O(h1/2)O(h^{1/2}) for k=1,2,3k=1,2,3, i.e. there is a loss of accuracy as expected.

Table 1: Errors of Example 1 without boundary correction on circular domain.
hh k=1k=1 k=2k=2 k=3k=3
Eu,pE_{u,p} order Eu,pE_{u,p} order Eu,pE_{u,p} order
1/8 4.89e-01 3.48e-01 3.44e-01
1/16 3.06e-01 0.68 2.53e-01 0.46 2.51e-01 0.45
1/32 1.96e-01 0.65 1.77e-01 0.52 1.76e-01 0.51
1/64 1.29e-01 0.60 1.22e-02 0.54 1.22e-02 0.53

We then solve the same problem using the proposed discretization (4.12) with boundary correction. We first set m=km=k and present the result in Tab. 2. An O(hk)O(h^{k}) optimal convergence is observed, which agrees well with the theoretical result in Cor. 5.1.

Cor. 5.1 also gives a lower bound mmax{0,k/23/4}m\geq\max\{0,k/2-3/4\} in order for the scheme to reach optimal convergence. For k=1,2,3k=1,2,3, the smallest mm that satisfies the lower bound is 0,1,10,1,1, respectively. We test the scheme (4.12) with these values, and present the results in Tab. 3. Again, optimal O(hk)O(h^{k}) convergence is observed, as predicted in Cor. 5.1.

Table 2: Errors of Example 1 with boundary correction for m=km=k on circular domain.
hh k=1k=1 k=2k=2 k=3k=3
Eu,pE_{u,p} order Eu,pE_{u,p} order Eu,pE_{u,p} order
1/8 3.84e-01 3.54e-03 6.54e-05
1/16 1.94e-01 0.98 7.61e-04 2.22 7.45e-06 3.13
1/32 9.47e-02 1.03 1.67e-04 2.18 7.85e-07 3.25
1/64 4.64e-02 1.03 3.86e-05 2.12 8.73e-08 3.17
Table 3: Errors of Example 1 with boundary correction for suitable mm on circular domain.
hh k=1,m=0k=1,m=0 k=2,m=1k=2,m=1 k=3,m=1k=3,m=1
Eu,pE_{u,p} order Eu,pE_{u,p} order Eu,pE_{u,p} order
1/8 3.86e-01 3.57e-03 6.47e-05
1/16 1.95e-01 0.99 7.71e-04 2.21 7.42e-06 3.13
1/32 9.48e-02 1.04 1.69e-04 2.19 7.82e-07 3.25
1/64 4.65e-02 1.03 3.89e-05 2.12 8.68e-08 3.17

𝐄𝐱𝐚𝐦𝐩𝐥𝐞𝟐.𝐑𝐢𝐧𝐠𝐝𝐨𝐦𝐚𝐢𝐧.\bf{Example~{}2.~{}Ring~{}domain.}

We repeat the above tests in a ring domain {(x,y)|0.25<x2+y2<1}\{(x,y)|0.25<x^{2}+y^{2}<1\}, which is not convex. The exact solution is

𝒖(x,y)=(2πcos(2πx)sin(2πy)2πsin(2πx)cos(2πy)),p(x,y)=sin(2πx)sin(2πy),\displaystyle\bm{u}(x,y)=\dbinom{2\pi\cos(2\pi x)\sin(2\pi y)}{2\pi\sin(2\pi x)\cos(2\pi y)},\qquad p(x,y)=-\sin(2\pi x)\sin(2\pi y),

which satisfies a non-homogeneous Neumann boundary condition 𝒖𝐧0\bm{u}\cdot\bm{\mathop{}\mathrm{n}}\neq 0 on Γ\Gamma and Ωpdx=0\int_{\Omega}p\mathop{}\mathrm{d}x=0. This time, we did not perform the test without boundary correction, since there is no effective way to impose the nonhomogeneous Neumann boundary condition on Γh\Gamma_{h}.

We test Example 22 on meshes illustrated in Fig. 4 (b), for various sizes of hh, using the scheme (4.12) with boundary correction. The results are reported in Tabs. 4 and 5, both indicating optimal convergence. This verifies that Cor. 5.1 holds on nonconvex domains.

Table 4: Errors of Example 2 with boundary correction for m=km=k on ring domain.
hh k=1k=1 k=2k=2 k=3k=3
Eu,pE_{u,p} order Eu,pE_{u,p} order Eu,pE_{u,p} order
1/8 1.36e+01 1.66e+00 1.50e-01
1/16 6.86e+00 0.99 4.20e-01 1.98 1.89e-02 2.99
1/32 3.44e+00 1.00 1.05e-01 2.00 2.37e-03 3.00
1/64 1.72e+00 1.00 2.64e-02 2.00 2.96e-04 3.00
Table 5: Errors of Example 2 with boundary correction for suitable mm on ring domain.
hh k=1,m=0k=1,m=0 k=2,m=1k=2,m=1 k=3,m=1k=3,m=1
Eu,pE_{u,p} order Eu,pE_{u,p} order Eu,pE_{u,p} order
1/8 1.36e+01 1.66e-00 1.50e-01
1/16 6.86e+00 0.99 4.20e-01 1.98 1.89e-02 2.99
1/32 3.44e+00 1.00 1.05e-01 2.00 2.37e-03 3.00
1/64 1.72e+00 1.00 2.64e-02 2.00 2.96e-04 3.00

7 Conclusion

In this paper, we consider the mixed finite element method for second-order elliptic equations on domains with curved boundaries. A boundary value correction technique is proposed to compensate for the loss of accuracy in the discretization. The proposed scheme achieves optimal convergence. Numerical experiments further validate the theoretical results.

References

  • [1] N. M. Atallah, C. Canuto, and G. Scovazzi, The high-order shifted boundary method and its analysis, Computer Methods in Applied Mechanics and Engineering, 394:114885, 2022.
  • [2] S. Bertoluzza, M. Pennacchio, and D. Prada, Weakly imposed Dirichlet boundary conditions for 2D and 3D virtual elements, Computer Methods in Applied Mechanics and Engineering, 400:115454, 2022.
  • [3] F. Bertrand, S. Munzenmaier, and G. Starke. First-order system least squares on curved boundaries: Lowest-order Raviart-Thomas elements, SIAM Journal on Numerical Analysis, 52:880-894, 2014.
  • [4] F. Bertrand, S. Munzenmaier, and G. Starke, First-order system least squares on curved boundaries: Higher-order Raviart-Thomas elements, SIAM Journal on Numerical Analysis, 52:3165-3180, 2014.
  • [5] F. Bertrand and G. Starke, Parametric Raviart-Thomas elements for mixed methods on domains with curved surfaces, SIAM Journal on Numerical Analysis, 54:3648-3667, 2016.
  • [6] D. Boffi, F. Brezzi, and M. Fortin, Mixed finite element methods and applications. Springer, 2013.
  • [7] D. Braess, Finite elements: theory, fast solvers, and applications in solid mechanics, Cambridge University Press: Cambridge, UK, 2007.
  • [8] J. H. Bramble, T. Dupont, and V. Thomée, Projection methods for Dirichlet’s problem in approximating polygonal domains with boundary-value corrections, Mathematics of Computation, 26:869-879, 1972.
  • [9] J. H. Bramble and J. T. King, A robust finite element method for nonhomogeneous Dirichlet problems in domains with curved boundaries, Mathematics of Computation, 63:1-17, 1994.
  • [10] S. C. Brenner and L. R. Scott, The mathematical theory of finite element methods, Springer, New York, 2008.
  • [11] F. Brezzi, J. Douglas, Jr., and L. D. Marini, Two families of mixed finite elements for second order elliptic problems, Numerische Mathematik, 47(2):217-235, 1985.
  • [12] E. Burman, P. Hansbo, and M. G. Larson, A cut finite element method with boundary value correction, Mathematics of Computation, 87:633-657, 2018.
  • [13] E. Burman and R. Puppi, Two mixed finite element formulations for the weak imposition of the Neumann boundary conditions for the Darcy flow, Journal of Numerical Mathematics, 30:141-162, 2022.
  • [14] P. Cao, J. Chen, and F. Wang, An extended mixed finite element method for elliptic interface problems, Computers Mathematics with Applications, 113:148-159, 2022.
  • [15] B. Cockburn, D. Gupta, and F. Reitich, Boundary-conforming discontinuous Galerkin methods via extensions from subdomains, Journal of Scientific Computing, 42:144-184, 2009.
  • [16] B. Cockburn and M. Solano, Solving Dirichlet boundary-value problems on curved domains by extensions from subdomains, SIAM Journal on Scientific Computing, 34:A497-A519, 2012.
  • [17] J. A. Cottrell, T. J. R. Hughes, and Y. Bazilevs, Isogeometric Analysis: Toward Integration of CAD and FEA, John Wiley &\& Sons, Ltd., Chichester, 2009.
  • [18] C. D’Angelo and A. Scotti, A mixed finite element method for Darcy flow in fractured porous media with non-matching grids, ESAIM: Mathematical Modelling and Numerical Analysis, 46: 465-489, 2012.
  • [19] R. Durán, Mixed finite element methods, vol. 1939 of Lecture Notes in Mathematics, Springer-Verlag, Berlin, 1-44, 2008, lectures given at the C.I.M.E. Summer School held in Cetraro, June 26-July 1, 2006. Edited by D. Boffi and L. Gastaldi.
  • [20] I. Ergatoudis, B. Irons, and O. Zienkiewicz, Curved, isoparametric, quadrilateral elements for finite element analysis, International Journal of Solids and Structures, 4:31-42, 1968.
  • [21] T. J. R. Hughes, J. A. Cottrell, and Y. Bazilevs, Isogeometric analysis: CAD, finite elements, NURBS, exact geometry and mesh refinement, Computer Methods in Applied Mechanics and Engineering, 194:4135-4195, 2005.
  • [22] M. Lenoir, Optimal isoparametric finite elements and error estimates for domains involving curved boundaries, SIAM Journal on Numerical Analysis, 23(3):562-580, 1986.
  • [23] Y. Liu, W. Chen, and Y. Wang, A weak Galerkin mixed finite element method for second order elliptic equations on 2D curved domains, Communications in Computational Physics, 32:1094-1128, 2022.
  • [24] A. Main and G. Scovazzi, The shifted boundary method for embedded domain computations. Part I: Poisson and Stokes problems, Journal of Computational Physics, 372:972-995, 2018.
  • [25] R. A. Nicolaides, Existence, uniqueness and approximation for generalized saddle point problems, SIAM ournal on Numerical Analysis, 19:349-357, 1982.
  • [26] R. Oyarzúa, M. Solano, and P. Zúñiga, A high order mixed-FEM for diffusion problems on curved domains, Journal of Scientific Computing, 79:49-78, 2019.
  • [27] R. Puppi, A cut finite element method for the Darcy problem, Preprint arXiv:2111.09922 [math. NA].
  • [28] E. M. Stein, Singular integrals and differentiability properties of functions, Princeton University Press, 2, 1970.
  • [29] G. Strang and A. E. Berger, The change in solution due to change in domain, Partial Differential Equations (D. C. Spencer, ed.), Proc. Sympos. Pure Math., vol. 23, Amer. Math. Soc, Providence, RI, 199-205, 1973.
  • [30] V. Thomée, Polygonal domain approximation in Dirichlet’s problem, Ima Journal of Applied Mathematics, 11:33-44, 1973.