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

Taylor-Hood like finite elements for nearly incompressible strain gradient elasticity problems

Yulei Liao and Pingbing Ming LSEC, Institute of Computational Mathematics and Scientific/Engineering Computing, AMSS, Chinese Academy of Sciences, No. 55, East Road Zhong-Guan-Cun, Beijing 100190, China School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China [email protected], [email protected]  and  Yun Xu Laboratory of Computational Physics, Institute of Applied Physics and Computational Mathematics, Beijing 100088, China [email protected]
Abstract.

We propose a family of mixed finite elements that are robust for the nearly incompressible strain gradient model, which is a fourth-order singular perturbed elliptic system. The element is similar to [C. Taylor and P. Hood, Comput. & Fluids, 1(1973), 73-100] in the Stokes flow. Using a uniform discrete B-B inequality for the mixed finite element pairs, we show the optimal rate of convergence that is robust in the incompressible limit. By a new regularity result that is uniform in both the materials parameter and the incompressibility, we prove the method converges with 1/21/2 order to the solution with strong boundary layer effects. Moreover, we estimate the convergence rate of the numerical solution to the unperturbed second-order elliptic system. Numerical results for both smooth solutions and the solutions with sharp layers confirm the theoretical prediction.

Key words and phrases:
Mixed finite elements; Nearly incompressible strain gradient elasticity; Uniform error estimate
2020 Mathematics Subject Classification:
Primary 65N15, 65N30; Secondary 74K20
The work of Liao and Ming were supported by National Natural Science Foundation of China through Grant No. 11971467. The work of Xu was supported by National Natural Science Foundation of China through Grant No. 11772067.

1. Introduction

The strain gradient models have drawn great attention recently because they capture the size effect of nano-materials for plasticity [FleckHutchinson:1997] as well as for mechanical meta-materials [Khakalo:2020] by incorporating the higher-order strain gradient and the intrinsic material length scale into the constitutive relations. Studies from the perspective of modeling, mechanics and mathematics may date back to 1960s [Koiter:19641, Mindlin:1964, Hlavacek:19692], while large-scale simulations are relatively recent [zervos:20092, Zybell:2012, Rud:2014, Phunpeng:2015]. Different methods such as H2 conforming finite element methods [zervos:20091, Fisher:2010], H1 conforming mixed finite element methods [Aravas:2002, Phunpeng:2015], nonconforming finite element methods [LiMingShi:2017, LiaoM:2019, LiMingWang:2021], discontinuous Galerkin methods [Engel:2002], isogeometric analysis [Klassen:2011, Niiranen:2016], and meshless methods [Askes:2002] have been used to simulate the strain gradient elastic models with different complexity, just to mention a few. One of the numerical difficulties is that the number of the materials parameters is too large [Mindlin:1964], another is that the materials parameters may cause boundary layer or numerical instability when they tend to certain critical values [Engel:2002].

The strain gradient elasticity model proposed by Altan and Aifantis [Altan:1992] seems the simplest one among them because it contains only one material parameter besides the Lamé constants, while it still models the size effect adequately [Aifantis:1999]. This model is described by the following boundary value problem:

(1.1) {(ι2Δ𝑰)(μΔ𝒖+(λ+μ)div𝒖)=𝒇in Ω,𝒖=𝒏𝒖=𝟎on Ω,\left\{\begin{aligned} (\iota^{2}\Delta-\boldsymbol{I})\left(\mu\Delta\boldsymbol{u}+(\lambda+\mu)\nabla\operatorname{div}\boldsymbol{u}\right)&=\boldsymbol{f}\quad&&\text{in\;}\Omega,\\ \boldsymbol{u}=\partial_{\boldsymbol{n}}\boldsymbol{u}&=\boldsymbol{0}\quad&&\text{on\;}\partial\Omega,\end{aligned}\right.

where Ω2\Omega\subset\mathbb{R}^{2} is a smooth domain, 𝒖:Ω2\boldsymbol{u}:\Omega\to\mathbb{R}^{2} is the displacement, 𝒏𝒖\partial_{\boldsymbol{n}}\boldsymbol{u} is the normal derivative of 𝒖\boldsymbol{u}, λ\lambda and μ\mu are the Lamé constants, and ι\iota is the microscopic parameter such that 0<ι10<\iota\leq 1, which stands for the intrinsic length scale. Besides modeling the strain gradient elasticity, the system (1.1) may also be regarded as a vector analog of the fourth-order singular perturbed problem, which usually models a clamped plate problem [John:1973, Schuss:1976, Semper:1992, Semper:1994, Tai:2001, Brenner:2011], or arises from a fourth-order perturbation of the fully nonlinear Monge-Ampère equation [Brenner:2011b, Feng:2014]. System of the form (1.1) may also come from the linearized model in MEMS [Laurencot:2017].

In the present work, we are interested in (1.1) for the nearly incompressible materials. Such materials are commonly used in industry and a typical example is natural rubber. To the best of our knowledge, the studies on the approximation of incompressible and nearly incompressible strain gradient elasticity have not been sufficiently addressed in the literature, although vast efforts have been devoted to finite element approximation of the incompressible and nearly incompressible elasticity problems; See e.g., [Hermann:1965, Vog:1983, Simo:1985, Babuska:1992, Brenner:1992, BraessMing:2005, Auricchio:2013]. In [FleckHutchinson:1997]*§III. C, the authors studied the incompressible limit of the strain gradient model. Mixed finite elements for the incompressible Fleck-Hutchinson strain gradient model have been designed and tested in [ShuKingFleck:1999]. A finite element method has been tested for the nearly incompressible strain gradient model in [Wei:2006]. A mixed finite element, which approximated the displacement with Bogner-Fox-Schmidt element [BFS:1965] and approximated the pressure with the 99-node quadrilateral element, was constructed for the coupled stress model in [Fisher:2011], and bore certain similarities with problem (1.1). Recently, Hu and Tian [Tian:2021] have proposed several robust elements for the two-dimensional strain gradient model in the framework of reduced integration. Unfortunately, none of the above work proved the robustness of the proposed elements rigorously in the incompressible limit.

Following the classical approach dealing with the nearly incompressible elasticity problem [Hermann:1965], we introduce an auxiliary variable “pressure” pp and recast (1.1) into a displacement-pressure mixed variational problem, i.e., the so-called (𝒖,p)(\boldsymbol{u},p)-formulation. We approximate the displacement by augmenting the finite element space in [GuzmanLeykekhmanNeilan:2012] with certain new bubble functions. The original motivation for the bubble functions is to design the stable finite element pair for the Stokes problem [Arnold:1984, Bernardi:1985]. The augmented bubble functions help out in dealing with the extra constraints such as the divergence stability in Stokes problem and the high order consistency error [Tai:2001, GuzmanLeykekhmanNeilan:2012, Zhang:2012]. Such idea has been exploited by one of the authors to design robust finite elements for the strain gradient elasticity model [LiMingShi:2017]. Besides, we employ the standard continuous Lagrangian finite element of one order lower than that for the displacement to approximate the pressure. Such a finite element pair is similar to the Taylor-Hood element in the Stokes flows [Hood:1973] which is rr1\mathbb{P}_{r}-\mathbb{P}_{r-1} scheme and continuous pressure approximation. For both smooth solutions and solutions with strong boundary layer effects, these mixed finite element pairs are robust in the incompressible limit, here the robustness is understood in the sense that the rate of convergence is uniform in both ι\iota and λ\lambda. The bubble function spaces in approximating the displacement are defined by certain orthogonal constraints, and the explicit representations of these spaces are desired for the sake of implementation. We achieve this with the aid of the Jacobi polynomial. In addition to perspicuous results in view of analytics, such representation lends itself to the construction of the analytical shape functions for the approximating space of the displacement. Though we focus on the two-dimensional problem, the element may be readily extended to the three-dimensional problem. cf., Remark 3.3.

By standard mixed finite element theory [BBF:2013], a discrete B-B inequality that is uniform in ι\iota is needed for the well-posedness of the mixed (𝒖,p)(\boldsymbol{u},p)-discretization problem. This B-B inequality reduces to the remarkable B-B inequality for the Stokes problem when ι\iota tends to zero. A natural way to prove the discrete B-B inequality is to construct a uniformly stable Fortin operator [Fortin:1977, Winther:2013, MardalTaiWinther:2002], which does not seem easy due to the complication of the constraints. To this end, we construct a quasi-Fortin operator that takes different forms for small ι\iota as well as large ι\iota. This quasi-Fortin operator is bounded in a weighted energy norm in the corresponding regimes of ι\iota. Besides the discrete B-B inequality, another ingredient in proving the robustness is a new regularity result for the solution of (1.1) that is uniform in both λ\lambda and ι\iota, which is crucial to prove the convergence rate for the layered solution. The proof combines the method dealing with the nearly incompressible linear elasticity [Vog:1983] and the regularity estimate for the fourth-order singular perturbed problem [Tai:2001, LiMingWang:2021].

The outline of the paper is as follows. In §2, we introduce Altan and Aifantis’ strain gradient model and its mixed variational formulation. We demonstrate the numerical difficulty caused by large λ\lambda, and prove the uniform regularity estimate for problem (1.1). In §3, we construct a family of nonconforming finite elements, and derive the explicit formulations for the bubble spaces. In §4, we use the nonconforming elements proposed in §3 together with the continuous Lagrangian finite elements to discretize the mixed variational problem and prove the optimal rate of convergence. In the last section, we report the numerical results, which are consistent with the theoretical prediction.

Throughout this paper, the constant CC may differ from line to line, while it is independent of the mesh size hh, the materials parameter ι\iota and the Lamé constant λ\lambda.

2. The Mixed Variational Formulation and Regularity Estimates

First we fix some notations. The space L2(Ω)L^{2}(\Omega) of the square-integrable functions defined on a smooth domain Ω\Omega is equipped with the inner product (,)(\cdot,\cdot) and the norm L2(Ω)\|\,\cdot\,\|_{L^{2}(\Omega)}, while L02(Ω)L_{0}^{2}(\Omega) is the subspace of L2(Ω)L^{2}(\Omega) with mean value zero. Let Hm(Ω)H^{m}(\Omega) be the standard Sobolev space [AdamsFournier:2003] with the norm Hm(Ω)\|\,\cdot\,\|_{H^{m}(\Omega)}, while H0m(Ω)H_{0}^{m}(\Omega) is the closure in Hm(Ω)H^{m}(\Omega) of C0(Ω)C_{0}^{\infty}(\Omega). We may drop Ω\Omega in Hm(Ω)\|\,\cdot\,\|_{H^{m}(\Omega)} when no confusion may occur. For any vector-valued function 𝒗\boldsymbol{v}, its gradient is a matrix-valued function with components (𝒗)ij=jvi(\nabla\boldsymbol{v})_{ij}=\partial_{j}v_{i}, and the symmetric part of 𝒗\nabla\boldsymbol{v} is defined by ϵ(𝒗)=(𝒗+[𝒗]T)/2.\boldsymbol{\epsilon}(\boldsymbol{v})=(\nabla\boldsymbol{v}+[\nabla\boldsymbol{v}]^{T})/2. The divergence operator is defined as div𝒗=1v1+2v2\operatorname{div}\boldsymbol{v}=\partial_{1}v_{1}+\partial_{2}v_{2}. The Sobolev spaces [Hm(Ω)]2[H^{m}(\Omega)]^{2}, [H0m(Ω)]2[H_{0}^{m}(\Omega)]^{2} and [L2(Ω)]2[L^{2}(\Omega)]^{2} of a vector-valued function may be defined similarly as their scalar counterpart. This rule equally applies to the inner products and the norms. The double inner product between tensors 𝑨=(Aij)i,j=12,𝑩=(Bij)i,j=12\boldsymbol{A}=(A_{ij})_{i,j=1}^{2},\boldsymbol{B}=(B_{ij})_{i,j=1}^{2} equals 𝑨:𝑩=i,j=12AijBij\boldsymbol{A}:\boldsymbol{B}=\sum_{i,j=1}^{2}A_{ij}B_{ij}.

We recast (1.1) into a variational problem: Find 𝒖V:=[H02(Ω)]2\boldsymbol{u}\in V{:}=[H^{2}_{0}(\Omega)]^{2} such that

(2.1) a(𝒖,𝒗)=(𝒇,𝒗)for all 𝒗V,a(\boldsymbol{u},\boldsymbol{v})=(\boldsymbol{f},\boldsymbol{v})\quad\text{for all\quad}\boldsymbol{v}\in V,

where a(𝒖,𝒗):=(ϵ(𝒖),ϵ(𝒗))+ι2(𝔻ϵ(𝒖),ϵ(𝒗)),a(\boldsymbol{u},\boldsymbol{v}){:}=(\mathbb{C}\boldsymbol{\epsilon}(\boldsymbol{u}),\boldsymbol{\epsilon}(\boldsymbol{v}))+\iota^{2}(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v})), and the fourth-order tensor \mathbb{C} and the sixth-order tensor 𝔻\mathbb{D} are defined as

ijkl=λδijδkl+2μδikδjland𝔻ijklmn=λδilδjkδmn+2μδilδjmδkn,\mathbb{C}_{ijkl}=\lambda\delta_{ij}\delta_{kl}+2\mu\delta_{ik}\delta_{jl}\quad\text{and}\quad\mathbb{D}_{ijklmn}=\lambda\delta_{il}\delta_{jk}\delta_{mn}+2\mu\delta_{il}\delta_{jm}\delta_{kn},

respectively. Here δij\delta_{ij} is the Kronecker delta function. The strain gradient ϵ(𝒗)\nabla\boldsymbol{\epsilon}(\boldsymbol{v}) is a third-order tensor that is defined by (ϵ(𝒗))ijk=i(ϵ(𝒗))jk(\nabla\boldsymbol{\epsilon}(\boldsymbol{v}))_{ijk}=\partial_{i}(\boldsymbol{\epsilon}(\boldsymbol{v}))_{jk}.

We are interested in the regime when λ\lambda\to\infty, which means that the material is nearly incompressible. Proceeding along the same line that leads to [LiMingWang:2021]*Theorem 5, the tensor product of the element (NTW) proposed in [Tai:2001] may be used to approximate (1.1), and the error estimate reads as

𝒖𝒖hCλ(h2+ιh)𝒖H3,\|\boldsymbol{u}-\boldsymbol{u}_{h}\|\leq C\lambda(h^{2}+\iota h)\|\,\boldsymbol{u}\,\|_{H^{3}},

where 𝒗2:=a(𝒗,𝒗)\|\boldsymbol{v}\|^{2}{:}=a(\boldsymbol{v},\boldsymbol{v}), and CC is independent of the mesh size hh, and ι\iota and λ\lambda. Therefore, the error bound degenerates when λ\lambda is large, and the NTW element does not seem a good candidate for the nearly incompressible material. The following numerical example confirms this observation.

Example 2.1.

Let Ω=(0,1)2\Omega=(0,1)^{2}, and 𝒖=(u1,u2)\boldsymbol{u}=(u_{1},u_{2}) with

u1=sin3(πx)sin(2πy)sin(πy),u2=sin(2πx)sin(πx)sin3(πy).u_{1}=-\sin^{3}(\pi x)\sin(2\pi y)\sin(\pi y),\quad u_{2}=\sin(2\pi x)\sin(\pi x)\sin^{3}(\pi y).

It is clear that div𝒖=0\operatorname{div}\boldsymbol{u}=0, hence the material is completely incompressible. The details of the numerical experiment such as the mesh generation, are the same as those in § 5. The relative error 𝒖𝒖h/𝒖\|\boldsymbol{u}-\boldsymbol{u}_{h}\|/\|\boldsymbol{u}\| in Table 1 shows that the rate of convergence deteriorates when λ\lambda is large.

Table 1. Relative errors and rate of convergence for NTW
ι\h\iota\backslash h 1/8 1/16 1/32 1/64
ν=0.3000,λ=0.5769,μ=0.3846\nu=0.3000,\lambda=0.5769,\mu=0.3846
1e+00 2.681e-01 1.373e-01 6.698e-02 3.334e-02
rate 0.97 1.04 1.01
1e-06 4.550e-02 1.244e-02 3.001e-03 7.467e-04
rate 1.87 2.05 2.01
ν=0.499999,λ=1.6667e5,μ=0.3333\nu=0.499999,\lambda=\text{1.6667e5},\mu=0.3333
1e+00 9.995e-01 9.979e-01 9.916e-01 9.682e-01
rate 0.00 0.01 0.03
1e-06 9.752e-01 7.233e-01 2.502e-01 6.561e-02
rate 0.43 1.53 1.93

2.1. The mixed variational formulation

We introduce an auxiliary variable p=λdiv𝒖,p=\lambda\operatorname{div}\boldsymbol{u}, and pP:=L02(Ω)H01(Ω)p\in P{:}=L_{0}^{2}(\Omega)\cap H_{0}^{1}(\Omega). We write Problem (2.1) into a mixed variational problem as

(2.2) {aι(𝒖,𝒗)+bι(𝒗,p)=(𝒇,𝒗)for all 𝒗V,bι(𝒖,q)λ1cι(p,q)=0for all qP,\left\{\begin{aligned} a_{\iota}(\boldsymbol{u},\boldsymbol{v})+b_{\iota}(\boldsymbol{v},p)&=(\boldsymbol{f},\boldsymbol{v})\qquad&&\text{for all\quad}\boldsymbol{v}\in V,\\ b_{\iota}(\boldsymbol{u},q)-\lambda^{-1}c_{\iota}(p,q)&=0\qquad&&\text{for all\quad}q\in P,\end{aligned}\right.

where

aι(𝒗,𝒘):\displaystyle a_{\iota}(\boldsymbol{v},\boldsymbol{w}){:} =2μ((ϵ(𝒗),ϵ(𝒘))+ι2(ϵ(𝒗),ϵ(𝒘))),\displaystyle=2\mu\Bigl{(}(\boldsymbol{\epsilon}(\boldsymbol{v}),\boldsymbol{\epsilon}(\boldsymbol{w}))+\iota^{2}(\nabla\boldsymbol{\epsilon}(\boldsymbol{v}),\nabla\boldsymbol{\epsilon}(\boldsymbol{w}))\Bigr{)},\qquad 𝒗,𝒘V,\displaystyle\boldsymbol{v},\boldsymbol{w}\in V,
bι(𝒗,q):\displaystyle b_{\iota}(\boldsymbol{v},q){:} =(div𝒗,q)+ι2(div𝒗,q),\displaystyle=(\operatorname{div}\boldsymbol{v},q)+\iota^{2}(\nabla\operatorname{div}\boldsymbol{v},\nabla q),\qquad 𝒗V,qP,\displaystyle\boldsymbol{v}\in V,q\in P,
cι(s,q):\displaystyle c_{\iota}(s,q){:} =(s,q)+ι2(s,q),\displaystyle=(s,q)+\iota^{2}(\nabla s,\nabla q),\qquad s,qP.\displaystyle s,q\in P.

It is convenient to define the weighted norm for all qPq\in P as

qι:=qL2+ιqL2.\|q\|_{\iota}{:}=\|\,q\,\|_{L^{2}}+\iota\|\,\nabla q\,\|_{L^{2}}.

qι\|q\|_{\iota} is a norm over PP for any qPq\in P and any finite ι\iota. By Poincaré’s inequality, 𝒗ι\|\nabla\boldsymbol{v}\|_{\iota} is a norm over VV for any 𝒗V\boldsymbol{v}\in V. To study the well-posedness of Problem (2.2), we start with the following B-B inequality that is uniform for any ι\iota.

Lemma 2.2.

For any qPq\in P, there exists 𝐯V\boldsymbol{v}\in V such that

(2.3) div𝒗=qand𝒗ιCqι,\operatorname{div}\boldsymbol{v}=q\qquad\text{and}\qquad\|\nabla\boldsymbol{v}\|_{\iota}\leq C\|q\|_{\iota},

where CC only depends on Ω\Omega but is independent of ι\iota.

Proof.

By [Galdi:2011]*Theorem III 3.3 and [Costabel:2010]*Proposition 4.1, for any qPq\in P, there exists 𝒗V\boldsymbol{v}\in V such that div𝒗=q\operatorname{div}\boldsymbol{v}=q and

(2.4) 𝒗H1CqL2and𝒗H2CqH1,\|\,\boldsymbol{v}\,\|_{H^{1}}\leq C\|\,q\,\|_{L^{2}}\qquad\text{and}\qquad\|\,\boldsymbol{v}\,\|_{H^{2}}\leq C\|\,q\,\|_{H^{1}},

where the constant CC only depends on Ω\Omega.

Because the mean of qq is zero for any qPq\in P, by Poincaré’s inequality, there exists CC such that

qH1CqL2.\|\,q\,\|_{H^{1}}\leq C\|\,\nabla q\,\|_{L^{2}}.

Combining the above two inequalities, we obtain

𝒗ι=𝒗L2+ι2𝒗L2Cqι.\|\nabla\boldsymbol{v}\|_{\iota}=\|\,\nabla\boldsymbol{v}\,\|_{L^{2}}+\iota\|\,\nabla^{2}\boldsymbol{v}\,\|_{L^{2}}\leq C\|q\|_{\iota}.

This gives (2.3). ∎

Lemma 2.3.

There exists a unique 𝐮V\boldsymbol{u}\in V and pPp\in P satisfying (2.2), and there exists CC independent of ι\iota and λ\lambda such that

(2.5) 𝒖ι+pιC𝒇H1.\|\nabla\boldsymbol{u}\|_{\iota}+\|p\|_{\iota}\leq C\|\,\boldsymbol{f}\,\|_{H^{-1}}.
Proof.

By the first Korn’s inequality [Korn:1908, Korn:1909],

(2.6) ϵ(𝒗)L2212𝒗L22for all 𝒗[H01(Ω)]2,\|\,\boldsymbol{\epsilon}(\boldsymbol{v})\,\|_{L^{2}}^{2}\geq\dfrac{1}{2}\|\,\nabla\boldsymbol{v}\,\|_{L^{2}}^{2}\qquad\text{for all\quad}\boldsymbol{v}\in[H_{0}^{1}(\Omega)]^{2},

and the H2 Korn’s inequality [LiMingWang:2021]*Theorem 1,

(2.7) ϵ(𝒗)L22(112)2𝒗L22for all 𝒗[H2(Ω)]2,\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{v})\,\|_{L^{2}}^{2}\geq\left(1-\dfrac{1}{\sqrt{2}}\right)\|\,\nabla^{2}\boldsymbol{v}\,\|_{L^{2}}^{2}\qquad\text{for all\quad}\boldsymbol{v}\in[H^{2}(\Omega)]^{2},

we obtain

aι(𝒗,𝒗)2μ(12𝒗L22+(112)ι22𝒗L22)μ2𝒗ι2.a_{\iota}(\boldsymbol{v},\boldsymbol{v})\geq 2\mu\left(\dfrac{1}{2}\|\,\nabla\boldsymbol{v}\,\|_{L^{2}}^{2}+\Bigl{(}1-\dfrac{1}{\sqrt{2}}\Bigr{)}\iota^{2}\|\,\nabla^{2}\boldsymbol{v}\,\|_{L^{2}}^{2}\right)\geq\dfrac{\mu}{2}\|\nabla\boldsymbol{v}\|_{\iota}^{2}.

Using (2.3), for any pPp\in P, there exists 𝒗0V\boldsymbol{v}_{0}\in V such that div𝒗0=p\operatorname{div}\boldsymbol{v}_{0}=p and 𝒗0ιCpι\|\nabla\boldsymbol{v}_{0}\|_{\iota}\leq C\|p\|_{\iota}. This implies

sup𝒗Vbι(𝒗,p)𝒗ιbι(𝒗0,p)𝒗0ι=pι2𝒗0ιCpι.\sup_{\boldsymbol{v}\in V}\dfrac{b_{\iota}(\boldsymbol{v},p)}{\|\nabla\boldsymbol{v}\|_{\iota}}\geq\dfrac{b_{\iota}(\boldsymbol{v}_{0},p)}{\|\nabla\boldsymbol{v}_{0}\|_{\iota}}=\dfrac{\|p\|_{\iota}^{2}}{\|\nabla\boldsymbol{v}_{0}\|_{\iota}}\geq C\|p\|_{\iota}.

By [Braess:1996]*Theorem 2, we immediately obtain the well-posedness of (2.2) and the estimate (2.5) by noting

|(𝒇,𝒗)|𝒇H1𝒗H1C𝒇H1𝒗L2C𝒇H1𝒗ι.\lvert(\boldsymbol{f},\boldsymbol{v})\rvert\leq\|\,\boldsymbol{f}\,\|_{H^{-1}}\|\,\boldsymbol{v}\,\|_{H^{1}}\leq C\|\,\boldsymbol{f}\,\|_{H^{-1}}\|\,\nabla\boldsymbol{v}\,\|_{L^{2}}\leq C\|\,\boldsymbol{f}\,\|_{H^{-1}}\|\nabla\boldsymbol{v}\|_{\iota}.

By the standard regularity theory for the elliptic system, we find 𝒖[H4(Ω)]2\boldsymbol{u}\in[H^{4}(\Omega)]^{2} and pH3(Ω)p\in H^{3}(\Omega) provided that 𝒇[L2(Ω)]2\boldsymbol{f}\in[L^{2}(\Omega)]^{2}, while we are interested in whether the shift estimate 2𝒖ι+pιC(ι)𝒇L2\|\,\nabla^{2}\boldsymbol{u}\,\|_{\iota}+\|\,\nabla p\,\|_{\iota}\leq C(\iota)\|\,\boldsymbol{f}\,\|_{L^{2}} is true with a λ\lambda-independent constant C(ι)C(\iota), this is the objective of the next part.

2.2. Regularity estimates

We aim to study the regularity of the solution of (1.1). Letting ι0\iota\to 0, we find 𝒖0[H01(Ω)]2\boldsymbol{u}_{0}\in[H_{0}^{1}(\Omega)]^{2} satisfying

(2.8) 𝒖0=𝒇in Ω,𝒖0=𝟎on Ω,-\mathcal{L}\boldsymbol{u}_{0}=\boldsymbol{f}\quad\text{in\;}\Omega,\qquad\boldsymbol{u}_{0}=\boldsymbol{0}\quad\text{on\;}\partial\Omega,

in the sense of distribution, where 𝒖0:=μΔ𝒖0+(λ+μ)div𝒖0\mathcal{L}\boldsymbol{u}_{0}{:}=\mu\Delta\boldsymbol{u}_{0}+(\lambda+\mu)\nabla\operatorname{div}\boldsymbol{u}_{0}. The H1{}^{1}-error for 𝒖𝒖0\boldsymbol{u}-\boldsymbol{u}_{0} will be given in Theorem 2.9, which is crucial for the regularity estimate of problem (1.1). We reshape (2.8) into a variational problem: Find 𝒖0[H01(Ω)]2\boldsymbol{u}_{0}\in[H_{0}^{1}(\Omega)]^{2} such that

(2.9) (ϵ(𝒖0),ϵ(𝒗))=(𝒇,𝒗)for all 𝒗[H01(Ω)]2.(\mathbb{C}\boldsymbol{\epsilon}(\boldsymbol{u}_{0}),\boldsymbol{\epsilon}(\boldsymbol{v}))=(\boldsymbol{f},\boldsymbol{v})\qquad\text{for all\;}\boldsymbol{v}\in[H_{0}^{1}(\Omega)]^{2}.

By [BacutaBramble:2003], we have the following shift estimate for 𝒖0\boldsymbol{u}_{0}: There exists CC independent of λ\lambda such that

(2.10) 𝒖0H2+λdiv𝒖0H1C𝒇L2.\|\,\boldsymbol{u}_{0}\,\|_{H^{2}}+\lambda\|\,\operatorname{div}\boldsymbol{u}_{0}\,\|_{H^{1}}\leq C\|\,\boldsymbol{f}\,\|_{L^{2}}.

Next we study an auxiliary boundary value problem:

(2.11) {Δ𝒘=𝑭,in Ω,𝒘=𝒏𝒘=𝟎,on Ω.\left\{\begin{aligned} \Delta\mathcal{L}\boldsymbol{w}&=\boldsymbol{F},\qquad\text{in\quad}\Omega,\\ \boldsymbol{w}=\partial_{\boldsymbol{n}}\boldsymbol{w}&=\boldsymbol{0},\qquad\text{on\quad}\partial\Omega.\end{aligned}\right.

The a-priori estimate for the solution of the above boundary value problem reads as

Lemma 2.4.

There exists a unique 𝐰V\boldsymbol{w}\in V satisfying (2.11), and there exists CC independent of λ\lambda such that

(2.12) 𝒘H2+λdiv𝒘H1C𝑭H2.\|\,\boldsymbol{w}\,\|_{H^{2}}+\lambda\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{1}}\leq C\|\,\boldsymbol{F}\,\|_{H^{-2}}.
Proof.

We recast (2.11) into a variational problem: Find 𝒘V\boldsymbol{w}\in V such that

A(𝒘,𝒗)=(𝑭,𝒗)for all 𝒗V,A(\boldsymbol{w},\boldsymbol{v})=(\boldsymbol{F},\boldsymbol{v})\qquad\text{for all\quad}\boldsymbol{v}\in V,

where A(𝒗,𝒛):=2μ(ϵ(𝒗),ϵ(𝒛))+λ(div𝒗,div𝒛)A(\boldsymbol{v},\boldsymbol{z}){:}=2\mu(\nabla\boldsymbol{\epsilon}(\boldsymbol{v}),\nabla\boldsymbol{\epsilon}(\boldsymbol{z}))+\lambda(\nabla\operatorname{div}\boldsymbol{v},\nabla\operatorname{div}\boldsymbol{z}) for any 𝒗,𝒛V\boldsymbol{v},\boldsymbol{z}\in V.

For any 𝒗V\boldsymbol{v}\in V, by the H2{}^{2}-Korn’s inequality (2.7) and Poincaré’s inequality, there exists CC such that

A(𝒗,𝒗)2μϵ(𝒗)L22μ22𝒗L22C𝒗H22.A(\boldsymbol{v},\boldsymbol{v})\geq 2\mu\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{v})\,\|_{L^{2}}^{2}\geq\dfrac{\mu}{2}\|\,\nabla^{2}\boldsymbol{v}\,\|_{L^{2}}^{2}\geq C\|\,\boldsymbol{v}\,\|_{H^{2}}^{2}.

The existence and uniqueness of 𝒘V\boldsymbol{w}\in V follow from the Lax-Milgram theorem, and

(2.13) 2𝒘L2𝒘H2C𝑭H2.\|\,\nabla^{2}\boldsymbol{w}\,\|_{L^{2}}\leq\|\,\boldsymbol{w}\,\|_{H^{2}}\leq C\|\,\boldsymbol{F}\,\|_{H^{-2}}.

Noting that div𝒘P\operatorname{div}\boldsymbol{w}\in P, using (2.4), we obtain that, there exists 𝒗0V\boldsymbol{v}_{0}\in V such that div𝒗0=div𝒘,\operatorname{div}\boldsymbol{v}_{0}=\operatorname{div}\boldsymbol{w}, and

2𝒗0L2Cdiv𝒘H1Cdiv𝒘L2.\|\,\nabla^{2}\boldsymbol{v}_{0}\,\|_{L^{2}}\leq C\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{1}}\leq C\|\,\nabla\operatorname{div}\boldsymbol{w}\,\|_{L^{2}}.

A combination of the above two inequalities gives

λdiv𝒘L22\displaystyle\lambda\|\,\nabla\operatorname{div}\boldsymbol{w}\,\|_{L^{2}}^{2} =λ(div𝒘,div𝒗0)=A(𝒘,𝒗0)2μ(ϵ(𝒘),ϵ(𝒗0))\displaystyle=\lambda(\nabla\operatorname{div}\boldsymbol{w},\nabla\operatorname{div}\boldsymbol{v}_{0})=A(\boldsymbol{w},\boldsymbol{v}_{0})-2\mu(\nabla\boldsymbol{\epsilon}(\boldsymbol{w}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v}_{0}))
=(𝑭,𝒗0)2μ(ϵ(𝒘),ϵ(𝒗0))\displaystyle=(\boldsymbol{F},\boldsymbol{v}_{0})-2\mu(\nabla\boldsymbol{\epsilon}(\boldsymbol{w}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v}_{0}))
𝑭H2𝒗0H2+2μ2𝒘L22𝒗0L2\displaystyle\leq\|\,\boldsymbol{F}\,\|_{H^{-2}}\|\,\boldsymbol{v}_{0}\,\|_{H^{2}}+2\mu\|\,\nabla^{2}\boldsymbol{w}\,\|_{L^{2}}\|\,\nabla^{2}\boldsymbol{v}_{0}\,\|_{L^{2}}
C(𝑭H2+2μ2𝒘L2)2𝒗0L2\displaystyle\leq C\left(\|\,\boldsymbol{F}\,\|_{H^{-2}}+2\mu\|\,\nabla^{2}\boldsymbol{w}\,\|_{L^{2}}\right)\|\,\nabla^{2}\boldsymbol{v}_{0}\,\|_{L^{2}}
C𝑭H2div𝒘L2.\displaystyle\leq C\|\,\boldsymbol{F}\,\|_{H^{-2}}\|\,\nabla\operatorname{div}\boldsymbol{w}\,\|_{L^{2}}.

This implies λdiv𝒘L2C𝑭H2,\lambda\|\,\nabla\operatorname{div}\boldsymbol{w}\,\|_{L^{2}}\leq C\|\,\boldsymbol{F}\,\|_{H^{-2}}, which together with (2.13) and Poincaré’s inequality gives (2.12). ∎

Now we turn to prove the regularity estimate of problem (2.11). We consider an auxiliary elliptic system: For any 𝑭~[L2(Ω)]2\widetilde{\boldsymbol{F}}\in[L^{2}(\Omega)]^{2} and G~H1(Ω)\widetilde{G}\in H^{1}(\Omega), find 𝒛V\boldsymbol{z}\in V and qPq\in P such that the following boundary value problem is valid in the sense of distribution:

(2.14) {μΔ2𝒛+Δq=𝑭~in Ω,Δdiv𝒛=G~in Ω,𝒛=𝒏𝒛=𝟎on Ω,q=0on Ω.\left\{\begin{aligned} \mu\Delta^{2}\boldsymbol{z}+\nabla\Delta q&=\widetilde{\boldsymbol{F}}\qquad&&\text{in\quad}\Omega,\\ \Delta\operatorname{div}\boldsymbol{z}&=\widetilde{G}\qquad&&\text{in\quad}\Omega,\\ \boldsymbol{z}=\partial_{\boldsymbol{n}}\boldsymbol{z}&=\boldsymbol{0}\qquad&&\text{on\quad}\partial\Omega,\\ q&=0\qquad&&\text{on\quad}\partial\Omega.\end{aligned}\right.
Lemma 2.5.

Let 𝐳[H4(Ω)]2\boldsymbol{z}\in[H^{4}(\Omega)]^{2} and qH3(Ω)q\in H^{3}(\Omega) be the solution of (2.14). Assume that mm is a nonnegative integer, then there exists CC depending only on Ω\Omega and μ\mu such that

(2.15) 𝒛Hm+4+qHm+3C(𝑭~Hm+G~Hm+1+𝒛L2+qL2).\|\,\boldsymbol{z}\,\|_{H^{m+4}}+\|\,q\,\|_{H^{m+3}}\leq C\left(\|\,\widetilde{\boldsymbol{F}}\,\|_{H^{m}}+\|\,\widetilde{G}\,\|_{H^{m+1}}+\|\,\boldsymbol{z}\,\|_{L^{2}}+\|\,q\,\|_{L^{2}}\right).
Proof.

We write (2.14)1 and (2.14)2 as

(μΔ20xΔ0μΔ2yΔΔxΔy0)(z1z2q)=(F~1F~2G~).\begin{pmatrix}\mu\Delta^{2}&0&\partial_{x}\Delta\\ 0&\mu\Delta^{2}&\partial_{y}\Delta\\ \Delta\partial_{x}&\Delta\partial_{y}&0\end{pmatrix}\begin{pmatrix}z_{1}\\ z_{2}\\ q\end{pmatrix}=\begin{pmatrix}\widetilde{F}_{1}\\ \widetilde{F}_{2}\\ \widetilde{G}\end{pmatrix}.

The symbol of the above system is

(ξ)=(μ|ξ|40ξ1|ξ|20μ|ξ|4ξ2|ξ|2ξ1|ξ|2ξ2|ξ|20).\mathcal{L}(\xi)=\begin{pmatrix}\mu\lvert\xi\rvert^{4}&0&\xi_{1}\lvert\xi\rvert^{2}\\ 0&\mu\lvert\xi\rvert^{4}&\xi_{2}\lvert\xi\rvert^{2}\\ \xi_{1}\lvert\xi\rvert^{2}&\xi_{2}\lvert\xi\rvert^{2}&0\end{pmatrix}.

A direct calculation gives

|det(ξ)|=μ|ξ|10>0if ξ0.\lvert\det\mathcal{L}(\xi)\rvert=\mu\lvert\xi\rvert^{10}>0\qquad\text{if\quad}\xi\not=0.

This means that the boundary value problem (2.14) is elliptic in the sense of Agmon-Douglis-Nirenberg [Agmon:1964]. Moreover, the boundary condition is pure Dirichlet, and it is straightforward to verify that the boundary condition satisfies the complementing condition [Agmon:1964]. The regularity estimate (2.15) follows from [Agmon:1964]. ∎

A direct consequence of the above lemma is the following regularity estimate for problem (2.11).

Lemma 2.6.

Let 𝐰V\boldsymbol{w}\in V be the solution of (2.11), there exists CC independent of λ\lambda such that

(2.16) 𝒘H3+λdiv𝒘H2C𝑭H1.\|\,\boldsymbol{w}\,\|_{H^{3}}+\lambda\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{2}}\leq C\|\,\boldsymbol{F}\,\|_{H^{-1}}.
Proof.

Using the standard elliptic regularity estimate, there exists a unique solution 𝒘[H4(Ω)]2\boldsymbol{w}\in[H^{4}(\Omega)]^{2} when 𝑭[L2(Ω)]2\boldsymbol{F}\in[L^{2}(\Omega)]^{2}. We introduce 𝒛=𝒘\boldsymbol{z}=\boldsymbol{w} and q=(λ+μ)div𝒘q=(\lambda+\mu)\operatorname{div}\boldsymbol{w}, hence 𝒛[H4(Ω)]2\boldsymbol{z}\in[H^{4}(\Omega)]^{2} and qH3(Ω)q\in H^{3}(\Omega) satisfy (2.14) with 𝑭~=𝑭\widetilde{\boldsymbol{F}}=\boldsymbol{F} and G~=Δdiv𝒘\widetilde{G}=\Delta\operatorname{div}\boldsymbol{w}.

By (2.15) with m=0m=0, we obtain

𝒘H4+(λ+μ)div𝒘H3C(𝑭L2+div𝒘H3+𝒘L2+(λ+μ)div𝒘L2).\|\,\boldsymbol{w}\,\|_{H^{4}}+(\lambda+\mu)\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{3}}\leq C\left(\|\,\boldsymbol{F}\,\|_{L^{2}}+\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{3}}+\|\,\boldsymbol{w}\,\|_{L^{2}}+(\lambda+\mu)\|\,\operatorname{div}\boldsymbol{w}\,\|_{L^{2}}\right).

Using the a-priori estimate (2.12), we obtain

𝒘H4+(λ+μ)div𝒘H3C0(𝑭L2+div𝒘H3).\|\,\boldsymbol{w}\,\|_{H^{4}}+(\lambda+\mu)\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{3}}\leq C_{0}\left(\|\,\boldsymbol{F}\,\|_{L^{2}}+\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{3}}\right).

Now for λ+μ>2C0\lambda+\mu>2C_{0}, it follows from

𝒘H4+(λ+μ)div𝒘H3C0𝑭L2+λ+μ2div𝒘H3\|\,\boldsymbol{w}\,\|_{H^{4}}+(\lambda+\mu)\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{3}}\leq C_{0}\|\,\boldsymbol{F}\,\|_{L^{2}}+\dfrac{\lambda+\mu}{2}\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{3}}

that

𝒘H4+λdiv𝒘H32C0𝑭L2.\|\,\boldsymbol{w}\,\|_{H^{4}}+\lambda\|\,\operatorname{div}\boldsymbol{w}\,\|_{H^{3}}\leq 2C_{0}\|\,\boldsymbol{F}\,\|_{L^{2}}.

Interpolating the above inequality with (2.12), we obtain (2.16).

If λ+μ2C0\lambda+\mu\leq 2C_{0}, then (2.16) follows from the standard regularity estimates [Agmon:1964] for problem (2.11). ∎

We turn to prove the regularity of problem (1.1) when 𝒇[L2(Ω)]2\boldsymbol{f}\in[L^{2}(\Omega)]^{2}. Let 𝒖\boldsymbol{u} and 𝒖0\boldsymbol{u}_{0} be the solutions of (1.1) and (2.8), respectively. For any 𝒗[H01(Ω)H2(Ω)]2\boldsymbol{v}\in[H_{0}^{1}(\Omega)\cap H^{2}(\Omega)]^{2}, integration by parts gives

(2.17) (ϵ(𝒖),ϵ(𝒗))=(𝒖,𝒗),(\mathbb{C}\boldsymbol{\epsilon}(\boldsymbol{u}),\boldsymbol{\epsilon}(\boldsymbol{v}))=-(\mathcal{L}\boldsymbol{u},\boldsymbol{v}),

and

(2.18) ι2(𝔻ϵ(𝒖),ϵ(𝒗))=ι2(Δ𝒖,𝒗)+ι2Ω(𝒏𝝈𝒏)𝒏𝒗dσ(𝒙),\iota^{2}(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v}))=\iota^{2}(\Delta\mathcal{L}\boldsymbol{u},\boldsymbol{v})+\iota^{2}\int_{\partial\Omega}{(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})}\cdot\partial_{\boldsymbol{n}}\boldsymbol{v}\mathrm{d}\sigma(\boldsymbol{x}),

where 𝝈:=2μϵ(𝒖)+λdiv𝒖𝑰\boldsymbol{\sigma}{:}=2\mu\boldsymbol{\epsilon}(\boldsymbol{u})+\lambda\operatorname{div}\boldsymbol{u}\boldsymbol{I} and (𝒏𝝈)ij:=𝒏σij(\partial_{\boldsymbol{n}}\boldsymbol{\sigma})_{ij}{:}=\partial_{\boldsymbol{n}}\sigma_{ij}. The boundary term in (2.18) is derived by the fact jvi=nj𝒏vi+tj𝒕vi=nj𝒏vi\partial_{j}v_{i}=n_{j}\partial_{\boldsymbol{n}}v_{i}+t_{j}\partial_{\boldsymbol{t}}v_{i}=n_{j}\partial_{\boldsymbol{n}}v_{i} and

(2.19) 2μΩ𝒏ϵ(𝒖):ϵ(𝒗)dσ(𝒙)\displaystyle 2\mu\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u}):\boldsymbol{\epsilon}(\boldsymbol{v})\mathrm{d}\sigma(\boldsymbol{x}) +λΩ𝒏div𝒖div𝒗dσ(𝒙)=Ω𝒏𝝈:𝒗dσ(𝒙)\displaystyle+\lambda\int_{\partial\Omega}\partial_{\boldsymbol{n}}\operatorname{div}\boldsymbol{u}\operatorname{div}\boldsymbol{v}\mathrm{d}\sigma(\boldsymbol{x})=\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\sigma}:\nabla\boldsymbol{v}\mathrm{d}\sigma(\boldsymbol{x})
=\displaystyle= Ω𝒏σijjvidσ(𝒙)=Ω𝒏σijnj𝒏vidσ(𝒙)=Ω(𝒏𝝈𝒏)𝒏vdσ(𝒙).\displaystyle\int_{\partial\Omega}\partial_{\boldsymbol{n}}\sigma_{ij}\partial_{j}v_{i}\mathrm{d}\sigma(\boldsymbol{x})=\int_{\partial\Omega}\partial_{\boldsymbol{n}}\sigma_{ij}n_{j}\partial_{\boldsymbol{n}}v_{i}\mathrm{d}\sigma(\boldsymbol{x})=\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}v\mathrm{d}\sigma(\boldsymbol{x}).

A combination of (2.17) and (2.18) leads to

(ϵ(𝒖),ϵ(𝒗))+ι2(𝔻ϵ(𝒖),ϵ(𝒗))=ι2Ω(𝒏𝝈𝒏)𝒏𝒗dσ(𝒙)+(𝒇,𝒗),(\mathbb{C}\boldsymbol{\epsilon}(\boldsymbol{u}),\boldsymbol{\epsilon}(\boldsymbol{v}))+\iota^{2}(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v}))=\iota^{2}\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{v}\mathrm{d}\sigma(\boldsymbol{x})+(\boldsymbol{f},\boldsymbol{v}),

which together with (2.9) yields

(2.20) (ϵ(𝒖𝒖0),ϵ(𝒗))+ι2(𝔻ϵ(𝒖),ϵ(𝒗))=ι2Ω(𝒏𝝈𝒏)𝒏𝒗dσ(𝒙).(\mathbb{C}\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0}),\boldsymbol{\epsilon}(\boldsymbol{v}))+\iota^{2}(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v}))=\iota^{2}\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{v}\mathrm{d}\sigma(\boldsymbol{x}).

This identity is the starting point of the proof.

We shall frequently use the following multiplicative trace inequality. There exists CC that depends only on Ω\Omega such that

(2.21) ψL2(Ω)CψL2(Ω)1/2ψH1(Ω)1/2,ψH1(Ω).\|\,\psi\,\|_{L^{2}(\partial\Omega)}\leq C\|\,\psi\,\|_{L^{2}(\Omega)}^{1/2}\|\,\psi\,\|_{H^{1}(\Omega)}^{1/2},\qquad\psi\in H^{1}(\Omega).

Using Lemma 2.6, we condense the regularity of problem (1.1) to estimating 𝒖𝒖0\boldsymbol{u}-\boldsymbol{u}_{0} and pp0p-p_{0} in various norms, with p=λdiv𝒖p=\lambda\operatorname{div}\boldsymbol{u} and p0=λdiv𝒖0p_{0}=\lambda\operatorname{div}\boldsymbol{u}_{0}.

Lemma 2.7.

There exists CC independent of ι\iota and λ\lambda such that

(2.22) 𝒖H3+pH2Cι2(ϵ(𝒖𝒖0)L2+pp0L2).\|\,\boldsymbol{u}\,\|_{H^{3}}+\|\,p\,\|_{H^{2}}\leq C\iota^{-2}(\|\,\epsilon(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}}+\|\,p-p_{0}\,\|_{L^{2}}).
Proof.

We rewrite (1.1) as

{Δ𝒖=ι2(𝒖𝒖0)in Ω,𝒖=𝒏𝒖=𝟎on Ω.\left\{\begin{aligned} \Delta\mathcal{L}\boldsymbol{u}&=\iota^{-2}\mathcal{L}(\boldsymbol{u}-\boldsymbol{u}_{0})\qquad&&\text{in\quad}\Omega,\\ \boldsymbol{u}=\partial_{\boldsymbol{n}}\boldsymbol{u}&=\boldsymbol{0}\qquad&&\text{on\quad}\partial\Omega.\end{aligned}\right.

Applying the regularity estimate (2.16) to the elliptic system (2.11), and using (2.17), we obtain

𝒖H3+pH2Cι2(𝒖𝒖0)H1Cι2(ϵ(𝒖𝒖0)L2+pp0L2).\|\,\boldsymbol{u}\,\|_{H^{3}}+\|\,p\,\|_{H^{2}}\leq C\iota^{-2}\|\,\mathcal{L}(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{H^{-1}}\leq C\iota^{-2}(\|\,\epsilon(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}}+\|\,p-p_{0}\,\|_{L^{2}}).

The next lemma is crucial to prove Theorem 2.9, which transforms the estimate of pp0ι\|\,p-p_{0}\,\|_{\iota} in terms of (𝒖𝒖0)ι\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota} besides a term concerning 𝒇\boldsymbol{f}.

Lemma 2.8.

There exists CC independent of ι\iota and λ\lambda such that

(2.23) pp0ιC((𝒖𝒖0)ι+ι1/2𝒇L2).\|\,p-p_{0}\,\|_{\iota}\leq C(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}).
Proof.

By [Danchin:2013]*Theorem 3.1 and Poincaré’s inequality, there exists 𝒗0[H2(Ω)H01(Ω)]2\boldsymbol{v}_{0}\in[H^{2}(\Omega)\cap H_{0}^{1}(\Omega)]^{2} such that div𝒗0=div(𝒖𝒖0)\operatorname{div}\boldsymbol{v}_{0}=\operatorname{div}(\boldsymbol{u}-\boldsymbol{u}_{0}), and

(2.24) 𝒗0H1Cdiv(𝒖𝒖0)L2,𝒗0H2Cdiv(𝒖𝒖0)L2.\|\,\boldsymbol{v}_{0}\,\|_{H^{1}}\leq C\|\,\operatorname{div}(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}},\qquad\|\,\boldsymbol{v}_{0}\,\|_{H^{2}}\leq C\|\,\nabla\operatorname{div}(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}}.

Substituting 𝒗=𝒗0\boldsymbol{v}=\boldsymbol{v}_{0} into (2.20), multiplying the resulting identity by λ\lambda, we obtain

pp0ι2\displaystyle\|\,p-p_{0}\,\|_{\iota}^{2} =ι2(p0,(pp0))2λμ((ϵ(𝒖𝒖0),ϵ(𝒗0))+ι2(ϵ(𝒖),ϵ(𝒗0)))\displaystyle=-\iota^{2}(\nabla p_{0},\nabla(p-p_{0}))-2\lambda\mu\Bigl{(}(\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0}),\boldsymbol{\epsilon}(\boldsymbol{v}_{0}))+\iota^{2}(\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v}_{0}))\Bigr{)}
(2.25) +λι2Ω(𝒏𝝈𝒏)𝒏𝒗0dσ(𝒙).\displaystyle\quad+\lambda\iota^{2}\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{v}_{0}\mathrm{d}\sigma(\boldsymbol{x}).

By the regularity estimate (2.10), the first term may be bounded as

ι2|(p0,(pp0))|ι28(pp0)L22+2ι2p0L2218pp0ι2+Cι2𝒇L22.\iota^{2}\lvert(\nabla p_{0},\nabla(p-p_{0}))\rvert\leq\dfrac{\iota^{2}}{8}\|\,\nabla(p-p_{0})\,\|_{L^{2}}^{2}+2\iota^{2}\|\,\nabla p_{0}\,\|_{L^{2}}^{2}\leq\dfrac{1}{8}\|\,p-p_{0}\,\|_{\iota}^{2}+C\iota^{2}\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}.

Using the triangle inequality and (2.10) again, we obtain

(2.26) ιϵ(𝒖)L2ιϵ(𝒖𝒖0)L2+ιϵ(𝒖0)L2(𝒖𝒖0)ι+Cι𝒇L2.\iota\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}}\leq\iota\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}}+\iota\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u}_{0})\,\|_{L^{2}}\leq\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+C\iota\|\,\boldsymbol{f}\,\|_{L^{2}}.

Using (2.24) and the above inequality, we bound the second term as

2λμ|(ϵ(𝒖𝒖0),ϵ(𝒗0))+ι2(ϵ(𝒖),ϵ(𝒗0))|\displaystyle 2\lambda\mu\lvert(\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0}),\boldsymbol{\epsilon}(\boldsymbol{v}_{0}))+\iota^{2}(\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{v}_{0}))\rvert C(ϵ(𝒖𝒖0)L2+ιϵ(𝒖)L2)pp0ι\displaystyle\leq C(\|\,\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}}+\iota\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}})\|\,p-p_{0}\,\|_{\iota}
18pp0ι2+C((𝒖𝒖0)ι2+ι2𝒇L22).\displaystyle\leq\dfrac{1}{8}\|\,p-p_{0}\,\|_{\iota}^{2}+C\left(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}^{2}+\iota^{2}\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}\right).

Using (2.19) and the definition of 𝒗0\boldsymbol{v}_{0}, we rewrite the boundary term as

λι2Ω(𝒏𝝈𝒏)𝒏𝒗0dσ(𝒙)\displaystyle\lambda\iota^{2}\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{v}_{0}\mathrm{d}\sigma(\boldsymbol{x}) =2λμι2Ω𝒏ϵ(𝒖):ϵ(𝒗0)dσ(𝒙)+λ2ι2Ω𝒏div𝒖div𝒗0dσ(𝒙)\displaystyle=2\lambda\mu\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u}):\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\mathrm{d}\sigma(\boldsymbol{x})+\lambda^{2}\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\operatorname{div}\boldsymbol{u}\operatorname{div}\boldsymbol{v}_{0}\mathrm{d}\sigma(\boldsymbol{x})
=2λμι2Ω𝒏ϵ(𝒖):ϵ(𝒗0)dσ(𝒙)+ι2Ω𝒏p(pp0)dσ(𝒙)\displaystyle=2\lambda\mu\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u}):\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\mathrm{d}\sigma(\boldsymbol{x})+\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}p(p-p_{0})\mathrm{d}\sigma(\boldsymbol{x})
=2λμι2Ω𝒏ϵ(𝒖):ϵ(𝒗0)dσ(𝒙)ι2Ω𝒏pp0dσ(𝒙).\displaystyle=2\lambda\mu\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u}):\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\mathrm{d}\sigma(\boldsymbol{x})-\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}p\;p_{0}\mathrm{d}\sigma(\boldsymbol{x}).

Recalling the trace inequality (2.21), we estimate the first boundary term as

2λμι2|Ω𝒏ϵ(𝒖):ϵ(𝒗0)dσ(𝒙)|\displaystyle 2\lambda\mu\iota^{2}\biggl{|}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u}):\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\mathrm{d}\sigma(\boldsymbol{x})\biggr{|} 2λμι2𝒏ϵ(𝒖)L2(Ω)ϵ(𝒗0)L2(Ω)\displaystyle\leq 2\lambda\mu\iota^{2}\|\,\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}(\partial\Omega)}\|\,\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\,\|_{L^{2}(\partial\Omega)}
Cλι2ϵ(𝒖)L21/2ϵ(𝒖)H11/2ϵ(𝒗0)L21/2ϵ(𝒗0)H11/2.\displaystyle\leq C\lambda\iota^{2}\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}}^{1/2}\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{H^{1}}^{1/2}\|\,\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\,\|_{L^{2}}^{1/2}\|\,\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\,\|_{H^{1}}^{1/2}.

Using (2.24), there exists CC independent of λ\lambda and ι\iota such that

λ2ϵ(𝒗0)L2ϵ(𝒗0)H1Cpp0L2(pp0)L2Cι1pp0ι2.\lambda^{2}\|\,\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\,\|_{L^{2}}\|\,\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\,\|_{H^{1}}\leq C\|\,p-p_{0}\,\|_{L^{2}}\|\,\nabla(p-p_{0})\,\|_{L^{2}}\leq C\iota^{-1}\|\,p-p_{0}\,\|_{\iota}^{2}.

Using (2.26) to estimate ϵ(𝒖)L2\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}} and using (2.22) to bound ϵ(𝒖)H1\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{H^{1}}, we obtain

ϵ(𝒖)L2ϵ(𝒖)H1\displaystyle\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}}\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{H^{1}} Cι3((𝒖𝒖0)ι+ι𝒇L2)(ϵ(𝒖𝒖0)L2+pp0L2)\displaystyle\leq C\iota^{-3}\left(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota\|\,\boldsymbol{f}\,\|_{L^{2}}\right)\left(\|\,\epsilon(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}}+\|\,p-p_{0}\,\|_{L^{2}}\right)
Cι3((𝒖𝒖0)ι2+ι2𝒇L22+((𝒖𝒖0)ι+ι𝒇L2)pp0ι).\displaystyle\leq C\iota^{-3}\Bigl{(}\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}^{2}+\iota^{2}\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}+\left(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota\|\,\boldsymbol{f}\,\|_{L^{2}}\right)\|\,p-p_{0}\,\|_{\iota}\Bigr{)}.

A combination of the above three inequalities gives

2λμι2|Ω𝒏ϵ(𝒖):ϵ(𝒗0)dσ(𝒙)|\displaystyle 2\lambda\mu\iota^{2}\biggl{|}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u}):\boldsymbol{\epsilon}(\boldsymbol{v}_{0})\mathrm{d}\sigma(\boldsymbol{x})\biggr{|}\leq C(((𝒖𝒖0)ι+ι𝒇L2)pp0ι\displaystyle C\Bigl{(}(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota\|\,\boldsymbol{f}\,\|_{L^{2}})\|\,p-p_{0}\,\|_{\iota}
+((𝒖𝒖0)ι+ι𝒇L2)1/2pp0ι3/2)\displaystyle\qquad+(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota\|\,\boldsymbol{f}\,\|_{L^{2}})^{1/2}\|\,p-p_{0}\,\|_{\iota}^{3/2}\Bigr{)}
\displaystyle\leq 18pp0ι2+C((𝒖𝒖0)ι2+ι2𝒇L22).\displaystyle\dfrac{1}{8}\|\,p-p_{0}\,\|_{\iota}^{2}+C\left(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}^{2}+\iota^{2}\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}\right).

Using the trace inequality (2.21) and the regularity estimate (2.10) again, we bound

ι2|Ω𝒏pp0dσ(𝒙)|Cι2pL21/2pH11/2p0H1Cι2pL21/2pH11/2𝒇L2.\iota^{2}\biggl{|}\int_{\partial\Omega}\partial_{\boldsymbol{n}}pp_{0}\mathrm{d}\sigma(\boldsymbol{x})\biggr{|}\leq C\iota^{2}\|\,\nabla p\,\|_{L^{2}}^{1/2}\|\,\nabla p\,\|_{H^{1}}^{1/2}\|\,p_{0}\,\|_{H^{1}}\leq C\iota^{2}\|\,\nabla p\,\|_{L^{2}}^{1/2}\|\,\nabla p\,\|_{H^{1}}^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}.

Using the triangle inequality and (2.10) again, we obtain

ιpL2pp0ι+ιp0L2pp0ι+Cι𝒇L2,\iota\|\,\nabla p\,\|_{L^{2}}\leq\|\,p-p_{0}\,\|_{\iota}+\iota\|\,\nabla p_{0}\,\|_{L^{2}}\leq\|\,p-p_{0}\,\|_{\iota}+C\iota\|\,\boldsymbol{f}\,\|_{L^{2}},

which together with (2.22) implies

pL2pH1Cι3(pp0ι2+((𝒖𝒖0)ι+ι𝒇L2)pp0ι+ι(𝒖𝒖0)ι𝒇L2).\|\,\nabla p\,\|_{L^{2}}\|\,\nabla p\,\|_{H^{1}}\leq C\iota^{-3}\Bigl{(}\|\,p-p_{0}\,\|_{\iota}^{2}+(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota\|\,\boldsymbol{f}\,\|_{L^{2}})\|\,p-p_{0}\,\|_{\iota}+\iota\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}\|\,\boldsymbol{f}\,\|_{L^{2}}\Bigr{)}.

Combining the above three inequalities, we bound the second boundary term as

ι2|Ω𝒏pp0dσ(𝒙)|\displaystyle\iota^{2}\biggl{|}\int_{\partial\Omega}\partial_{\boldsymbol{n}}pp_{0}\mathrm{d}\sigma(\boldsymbol{x})\biggr{|}\leq C(ι1/2pp0ι𝒇L2+((𝒖𝒖0)ι+ι1/2𝒇L2)3/2pp0ι1/2\displaystyle C\Bigl{(}\iota^{1/2}\|\,p-p_{0}\,\|_{\iota}\|\,\boldsymbol{f}\,\|_{L^{2}}+(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}})^{3/2}\|\,p-p_{0}\,\|_{\iota}^{1/2}
+ι(𝒖𝒖0)ι1/2𝒇L23/2)\displaystyle\qquad+\iota\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}^{3/2}\Bigr{)}
18pp0ι2+C((𝒖𝒖0)ι2+ι𝒇L22).\displaystyle\leq\dfrac{1}{8}\|\,p-p_{0}\,\|_{\iota}^{2}+C\left(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}^{2}+\iota\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}\right).

Substituting the above inequalities into (2.25), we obtain

pp0ι212pp0ι2+C((𝒖𝒖0)ι2+ι𝒇L22).\|\,p-p_{0}\,\|_{\iota}^{2}\leq\dfrac{1}{2}\|\,p-p_{0}\,\|_{\iota}^{2}+C\left(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}^{2}+\iota\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}\right).

This immediately gives (2.23). ∎

We are ready to prove the regularity of problem (1.1).

Theorem 2.9.

There exists CC independent of ι\iota and λ\lambda such that

(2.27) 𝒖𝒖0H1+pp0L2Cι1/2𝒇L2,\|\,\boldsymbol{u}-\boldsymbol{u}_{0}\,\|_{H^{1}}+\|\,p-p_{0}\,\|_{L^{2}}\leq C\iota^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}},

and

(2.28) 𝒖H2+k+pH1+kCι1/2k𝒇L2,k=0,1.\|\,\boldsymbol{u}\,\|_{H^{2+k}}+\|\,p\,\|_{H^{1+k}}\leq C\iota^{-1/2-k}\|\,\boldsymbol{f}\,\|_{L^{2}},\qquad k=0,1.

The estimate (2.27) improves the known results [LiMingWang:2021]*Lemma 1 in two aspects. It clarifies the fact that the estimate is λ\lambda-independent and it gives the convergence rate for the pressure. The rate ι1/2\iota^{1/2} is optimal even for the scalar counterpart; cf., [Tai:2001]*Lemma 5.1.

Proof.

Substituting 𝒗=𝒖𝒖0\boldsymbol{v}=\boldsymbol{u}-\boldsymbol{u}_{0} into (2.20), we get

a(𝒖𝒖0,𝒖𝒖0)=ι2(𝔻ϵ(𝒖),ϵ(𝒖𝒖0))ι2Ω(𝒏𝝈𝒏)𝒏𝒖0dσ(𝒙).a(\boldsymbol{u}-\boldsymbol{u}_{0},\boldsymbol{u}-\boldsymbol{u}_{0})=-\iota^{2}(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0}))-\iota^{2}\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{u}_{0}\mathrm{d}\sigma(\boldsymbol{x}).

Using the regularity estimate (2.10), we bound the first term as

ι2|(𝔻ϵ(𝒖),ϵ(𝒖𝒖0))|ι24(𝔻ϵ(𝒖𝒖0),ϵ(𝒖𝒖0))+ι2(𝔻ϵ(𝒖0),ϵ(𝒖0))14a(𝒖𝒖0,𝒖𝒖0)+Cι2𝒇L22.\iota^{2}\lvert(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}),\nabla\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0}))\rvert\leq\dfrac{\iota^{2}}{4}(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0}),\nabla\boldsymbol{\epsilon}(\boldsymbol{u}-\boldsymbol{u}_{0}))+\iota^{2}(\mathbb{D}\nabla\boldsymbol{\epsilon}(\boldsymbol{u}_{0}),\nabla\boldsymbol{\epsilon}(\boldsymbol{u}_{0}))\leq\dfrac{1}{4}a(\boldsymbol{u}-\boldsymbol{u}_{0},\boldsymbol{u}-\boldsymbol{u}_{0})+C\iota^{2}\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}.

To bound the second term, we let 𝒗=𝒖0\boldsymbol{v}=\boldsymbol{u}_{0} in (2.19) and obtain

ι2Ω(𝒏𝝈𝒏)𝒏𝒖0dσ(𝒙)=2μι2Ω𝒏ϵ(𝒖):ϵ(𝒖0)dσ(𝒙)+ι2Ω𝒏div𝒖p0dσ(𝒙).\iota^{2}\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{u}_{0}\mathrm{d}\sigma(\boldsymbol{x})=2\mu\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\boldsymbol{\epsilon}(\boldsymbol{u}):\boldsymbol{\epsilon}(\boldsymbol{u}_{0})\mathrm{d}\sigma(\boldsymbol{x})+\iota^{2}\int_{\partial\Omega}\partial_{\boldsymbol{n}}\operatorname{div}\boldsymbol{u}\;p_{0}\mathrm{d}\sigma(\boldsymbol{x}).

Invoking the trace inequality, using the fact div𝒖L2ϵ(𝒖)L2\|\,\nabla\operatorname{div}\boldsymbol{u}\,\|_{L^{2}}\leq\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}} and (2.10), we obtain

ι2|Ω(𝒏𝝈𝒏)𝒏𝒖0dσ(𝒙)|\displaystyle\iota^{2}\left\lvert\,\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{u}_{0}\mathrm{d}\sigma(\boldsymbol{x})\,\right\rvert Cι2(ϵ(𝒖)L21/2ϵ(𝒖)H11/2ϵ(𝒖0)H1+div𝒖L21/2div𝒖H11/2p0H1)\displaystyle\leq C\iota^{2}\left(\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}}^{1/2}\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{H^{1}}^{1/2}\|\,\boldsymbol{\epsilon}(\boldsymbol{u}_{0})\,\|_{H^{1}}+\|\,\nabla\operatorname{div}\boldsymbol{u}\,\|_{L^{2}}^{1/2}\|\,\nabla\operatorname{div}\boldsymbol{u}\,\|_{H^{1}}^{1/2}\|\,p_{0}\,\|_{H^{1}}\right)
Cι2ϵ(𝒖)L21/2𝒖H31/2(𝒖0H2+p0H1)\displaystyle\leq C\iota^{2}\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}}^{1/2}\|\,\boldsymbol{u}\,\|_{H^{3}}^{1/2}\left(\|\,\boldsymbol{u}_{0}\,\|_{H^{2}}+\|\,p_{0}\,\|_{H^{1}}\right)
Cι2ϵ(𝒖)L21/2𝒖H31/2𝒇L2.\displaystyle\leq C\iota^{2}\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}}^{1/2}\|\,\boldsymbol{u}\,\|_{H^{3}}^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}.

Substituting (2.23) into (2.22), we obtain,

𝒖H3+pH2Cι2((𝒖𝒖0)ι+ι1/2𝒇L2).\|\,\boldsymbol{u}\,\|_{H^{3}}+\|\,p\,\|_{H^{2}}\leq C\iota^{-2}(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}).

Invoking (2.26) again, we get

ι2ϵ(𝒖)L21/2𝒖H31/2Cι1/2((𝒖𝒖0)ι+ι1/2𝒇L2).\iota^{2}\|\,\nabla\boldsymbol{\epsilon}(\boldsymbol{u})\,\|_{L^{2}}^{1/2}\|\,\boldsymbol{u}\,\|_{H^{3}}^{1/2}\leq C\iota^{1/2}(\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}+\iota^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}).

A combination of the above three inequalities gives

ι2|Ω(𝒏𝝈𝒏)𝒏𝒖0dσ(𝒙)|14(𝒖𝒖0)ι2+Cι𝒇L22.\iota^{2}\left\lvert\,\int_{\partial\Omega}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot\partial_{\boldsymbol{n}}\boldsymbol{u}_{0}\mathrm{d}\sigma(\boldsymbol{x})\,\right\rvert\leq\dfrac{1}{4}\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}^{2}+C\iota\|\,\boldsymbol{f}\,\|_{L^{2}}^{2}.

Combining the above inequalities, we obtain

(𝒖𝒖0)ιCι1/2𝒇L2,\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{\iota}\leq C\iota^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}},

which together with (2.23) leads to

pp0ιCι1/2𝒇L2.\|\,p-p_{0}\,\|_{\iota}\leq C\iota^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}.

The above two estimates immediately give (2.27) and

2(𝒖𝒖0)L2+(pp0)L2Cι1/2𝒇L2,\|\,\nabla^{2}(\boldsymbol{u}-\boldsymbol{u}_{0})\,\|_{L^{2}}+\|\,\nabla(p-p_{0})\,\|_{L^{2}}\leq C\iota^{-1/2}\|\,\boldsymbol{f}\,\|_{L^{2}},

which together with (2.10) gives (2.28) with k=0k=0.

Substituting (2.27) into (2.22), we obtain the higher regularity estimate (2.28) with k=1k=1. ∎

A direct consequence of the above theorem is

Corollary 2.10.

There exists CC independent of ι\iota and λ\lambda such that

(2.29) 𝒖H3/2+pH1/2C𝒇L2,\|\,\boldsymbol{u}\,\|_{H^{3/2}}+\|\,p\,\|_{H^{1/2}}\leq C\|\,\boldsymbol{f}\,\|_{L^{2}},

and

(2.30) 𝒖H5/2+pH3/2Cι1𝒇L2.\|\,\boldsymbol{u}\,\|_{H^{5/2}}+\|\,p\,\|_{H^{3/2}}\leq C\iota^{-1}\|\,\boldsymbol{f}\,\|_{L^{2}}.
Proof.

Using triangle inequality, (2.10) and (2.28), we obtain

𝒖𝒖0H2+pp0H1𝒖H2+pH1+𝒖0H2+p0H1Cι1/2𝒇L2.\|\,\boldsymbol{u}-\boldsymbol{u}_{0}\,\|_{H^{2}}+\|\,p-p_{0}\,\|_{H^{1}}\leq\|\,\boldsymbol{u}\,\|_{H^{2}}+\|\,p\,\|_{H^{1}}+\|\,\boldsymbol{u}_{0}\,\|_{H^{2}}+\|\,p_{0}\,\|_{H^{1}}\leq C\iota^{-1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}.

Interpolating the above inequality and (2.27), we obtain

𝒖𝒖0H3/2+pp0H1/2C𝒇L2.\|\,\boldsymbol{u}-\boldsymbol{u}_{0}\,\|_{H^{3/2}}+\|\,p-p_{0}\,\|_{H^{1/2}}\leq C\|\,\boldsymbol{f}\,\|_{L^{2}}.

Using (2.10), we get

𝒖0H3/2+p0H1/2C𝒖0H2+p0H1C𝒇L2.\|\,\boldsymbol{u}_{0}\,\|_{H^{3/2}}+\|\,p_{0}\,\|_{H^{1/2}}\leq C\|\,\boldsymbol{u}_{0}\,\|_{H^{2}}+\|\,p_{0}\,\|_{H^{1}}\leq C\|\,\boldsymbol{f}\,\|_{L^{2}}.

A combination of the above two inequalities and the triangle inequality leads to (2.29).

Interpolating (2.28) with k=0k=0 and (2.28) with k=1k=1, we obtain (2.30). ∎

3. A Family of Nonconforming Finite Elements

We introduce a family of finite elements to approximate the mixed variational problem (2.2). Let 𝒯h\mathcal{T}_{h} be a triangulation of Ω\Omega with maximum mesh size hh. We assume all elements in 𝒯h\mathcal{T}_{h} are shape-regular in the sense of Ciarlet and Raviart [Ciarlet:1978]. We also assume that 𝒯h\mathcal{T}_{h} satisfies the inverse assumption: there exists σ0\sigma_{0} such that h/hKσ0h/h_{K}\leq\sigma_{0} for all K𝒯hK\in\mathcal{T}_{h}. The space of piecewise vector fields is defined by

[Hm(Ω,𝒯h)]2:={𝒗[L2(Ω)]2𝒗|K[Hm(T)]2for all K𝒯h},[H^{m}(\Omega,\mathcal{T}_{h})]^{2}{:}=\left\{\,\boldsymbol{v}\in[L^{2}(\Omega)]^{2}\,\mid\,\boldsymbol{v}|_{K}\in[H^{m}(T)]^{2}\quad\text{for all\quad}K\in\mathcal{T}_{h}\,\right\},

which is equipped with the broken norm

𝒗Hm(Ω,𝒯h):=𝒗L2+k=1mhk𝒗L2,\|\,\boldsymbol{v}\,\|_{H^{m}(\Omega,\mathcal{T}_{h})}{:}=\|\,\boldsymbol{v}\,\|_{L^{2}}+\sum_{k=1}^{m}\|\,\nabla^{k}_{h}\boldsymbol{v}\,\|_{L^{2}},

where hk𝒗L22=K𝒯hk𝒗L2(K)2.\|\,\nabla^{k}_{h}\boldsymbol{v}\,\|_{L^{2}}^{2}=\sum_{K\in\mathcal{T}_{h}}\|\,\nabla^{k}\boldsymbol{v}\,\|_{L^{2}(K)}^{2}. For an interior edge ee shared by the triangles K+K^{+} and KK^{-}, we define the jump of 𝒗\boldsymbol{v} across ee as

[[𝒗]]:=𝒗+𝒏++𝒗𝒏with𝒗±=v|K±,[\![\boldsymbol{v}]\!]{:}=\boldsymbol{v}^{+}\boldsymbol{n}^{+}+\boldsymbol{v}^{-}\boldsymbol{n}^{-}\qquad\text{with}\quad\boldsymbol{v}^{\pm}=v|_{K^{\pm}},

where 𝒏±\boldsymbol{n}^{\pm} is the unit normal vector of ee towards the outside of K±K^{\pm}. For eΩe\cap\partial\Omega\not=\emptyset, we set [[𝒗]]=𝒗𝒏[\![\boldsymbol{v}]\!]=\boldsymbol{v}\boldsymbol{n}.

3.1. A family of finite elements

Our construction is motivated by the element proposed in [GuzmanLeykekhmanNeilan:2012]. Define the element with a triple (K,PK,ΣK)(K,P_{K},\Sigma_{K}) by specifying KK as a triangle, and

(3.1) PK:=r(K)+bKi=13biQir2(K)+bK2Rr2(K),P_{K}{:}=\mathbb{P}_{r}(K)+b_{K}\sum_{i=1}^{3}b_{i}Q_{i}^{r-2}(K)+b_{K}^{2}R^{r-2}(K),

where bK=i=13λib_{K}=\prod_{i=1}^{3}\lambda_{i} and bi=bK/λib_{i}=b_{K}/\lambda_{i} with {λi}i=13\{\lambda_{i}\}_{i=1}^{3} the barycentric coordinates of KK.

Define

(3.2) Qir2(K):={vr2(K)KbKbivqd𝒙=0 for all qr3(K)},Q_{i}^{r-2}(K){:}=\left\{\,v\in\mathbb{P}_{r-2}(K)\,\mid\,\int_{K}b_{K}b_{i}vq\,\mathrm{d}\boldsymbol{x}=0\textup{ for all }q\in\mathbb{P}_{r-3}(K)\,\right\},

and

(3.3) Rr2(K):={vr2(K)KbK2vqd𝒙=0 for all qr3(K)}.R^{r-2}(K){:}=\left\{\,v\in\mathbb{P}_{r-2}(K)\,\mid\,\int_{K}b_{K}^{2}vq\,\mathrm{d}\boldsymbol{x}=0\textup{ for all }q\in\mathbb{P}_{r-3}(K)\,\right\}.

The degrees of freedom (DoFs) for PKP_{K} are given by

(3.4) v{v(a)for all verticesa,evqdσ(𝒙)for all edges e and qr2(e),e𝒏vqdσ(𝒙)for all edges e and qr2(e),Kvqd𝒙for all qr2(K).v\mapsto\left\{\begin{aligned} v(a)&\quad\text{for all vertices}\,a,\\ \int_{e}vq\,\mathrm{d}\sigma(\boldsymbol{x})&\quad\text{for all edges $e$ and\,}q\in\mathbb{P}_{r-2}(e),\\ \int_{e}\partial_{\boldsymbol{n}}vq\,\mathrm{d}\sigma(\boldsymbol{x})&\quad\text{for all edges $e$ and\,}q\in\mathbb{P}_{r-2}(e),\\ \int_{K}vq\,\mathrm{d}\boldsymbol{x}&\quad\text{for all\,}q\in\mathbb{P}_{r-2}(K).\end{aligned}\right.

We plot the DoFs for r=2,3r=2,3 in Figure 1.

a1a_{1}a2a_{2}a3a_{3}e1e_{1}e2e_{2}e3e_{3}a1a_{1}a2a_{2}a3a_{3}e1e_{1}e2e_{2}e3e_{3}
Figure 1. Diagram for DoFs. Left: DoFs for r=2r=2 are point evaluations of the function values at the vertex, the mean of the function along each edge, the mean of the normal derivative along each edge, and the mean of the function over the element; Right: DoFs for r=3r=3 are point evaluations of the function values at the vertex, the means of the function against 1\mathbb{P}_{1} along each edge, the means of the normal derivative against 1\mathbb{P}_{1} along each edge, and the means of the function against 1\mathbb{P}_{1} over the element
Lemma 3.1.

The set (K,PK,ΣK)(K,P_{K},\Sigma_{K}) is unisolvent.

Proof.

Firstly we show that if the DoFs (3.4) can determine an element in PKP_{K}, then the element is unique. Suppose vPKv\in P_{K} vanishes at the DoFs listed in (3.4), it suffices to show v0v\equiv 0. Assume that

v=pr+bKi=13biqi+bK2qr,v=p_{r}+b_{K}\sum_{i=1}^{3}b_{i}q_{i}+b_{K}^{2}q_{r},

where prr(K)p_{r}\in\mathbb{P}_{r}(K), and qiQir2(K)q_{i}\in Q_{i}^{r-2}(K), and qrRr2(K)q_{r}\in R^{r-2}(K). DoFs associated with r(K)\mathbb{P}_{r}(K) are

v{v(a)for all verticesa,evqdσ(𝒙)for all edges e and qr2(e),Kvqd𝒙for all qr3(K).v\mapsto\left\{\begin{aligned} v(a)&\quad\text{for all vertices}\,a,\\ \int_{e}vq\,\mathrm{d}\sigma(\boldsymbol{x})&\quad\text{for all edges $e$ and\,}q\in\mathbb{P}_{r-2}(e),\\ \int_{K}vq\,\mathrm{d}\boldsymbol{x}&\quad\text{for all\,}q\in\mathbb{P}_{r-3}(K).\end{aligned}\right.

The bubble space vanishes on this subset of DoFs by (3.2) and (3.3). The number of the DoFs is 3+3(r1)+(r1)(r2)/2=(r+1)(r+2)/23+3(r-1)+(r-1)(r-2)/2=(r+1)(r+2)/2, which equals to the cardinality of r(K)\mathbb{P}_{r}(K). Hence pr0p_{r}\equiv 0.

A direct calculation gives

ei𝒏vqdσ(𝒙)=|λi|eibi2qiqdσ(𝒙)=0for all qr2(ei).\int_{e_{i}}\partial_{\boldsymbol{n}}vq\,\mathrm{d}\sigma(\boldsymbol{x})=-\lvert\nabla\lambda_{i}\rvert\int_{e_{i}}b_{i}^{2}q_{i}q\,\mathrm{d}\sigma(\boldsymbol{x})=0\qquad\text{for all }q\in\mathbb{P}_{r-2}(e_{i}).

Taking q=qiq=q_{i} in the above identity, we obtain qi=0q_{i}=0 on eie_{i}. Therefore, we write qi=λipr3q_{i}=\lambda_{i}p_{r-3} for certain pr3r3(K)p_{r-3}\in\mathbb{P}_{r-3}(K). Using (3.2), we get

KbKbiqiqd𝒙=KbK2pr3qd𝒙=0for all qr3(K).\int_{K}b_{K}b_{i}q_{i}q\,\mathrm{d}\boldsymbol{x}=\int_{K}b_{K}^{2}p_{r-3}q\,\mathrm{d}\boldsymbol{x}=0\qquad\text{for all }q\in\mathbb{P}_{r-3}(K).

Taking q=pr3q=p_{r-3} in the above identity, we obtain pr30p_{r-3}\equiv 0. Therefore v=bK2qrv=b_{K}^{2}q_{r} for certain qrRr2(K)q_{r}\in R^{r-2}(K). The last set of DoFs equals zero, i.e.,

KbK2qrqd𝒙=0for all qr2(K).\int_{K}b_{K}^{2}q_{r}q\,\mathrm{d}\boldsymbol{x}=0\qquad\text{for all }q\in\mathbb{P}_{r-2}(K).

Taking q=qrq=q_{r} in the above identity, we obtain qr0q_{r}\equiv 0. So does vv.

It remains to show the dimension of PKP_{K} equals the number of DoFs (3.4). Proceeding along the same line as above, the element v0v\equiv 0 has a unique representation. Therefore (3.1) is a direct sum, and

dimPK=dimr(K)+4(dimr2(K)dimr3(K))=12(r2+11r6),\text{dim}P_{K}=\text{dim}\mathbb{P}_{r}(K)+4\left(\text{dim}\mathbb{P}_{r-2}(K)-\text{dim}\mathbb{P}_{r-3}(K)\right)=\dfrac{1}{2}(r^{2}+11r-6),

which equals to the number of DoFs (3.4) exactly. ∎

We define a local interpolation operator πK:H2(K)PK\pi_{K}:H^{2}(K)\mapsto P_{K} as:

(3.5) {πKv(a)=v(a)for all vertices a,eπKvqdσ(𝒙)=evqdσ(𝒙)for all edges e and qr2(e),e𝒏πKvqdσ(𝒙)=e𝒏vqdσ(𝒙)for all edges e and qr2(e),KπKvqd𝒙=Kvqd𝒙for all qr2(K).\left\{\begin{aligned} \pi_{K}v(a)&=v(a)&&\quad\text{for all vertices\,}a,\\ \int_{e}\pi_{K}vq\,\mathrm{d}\sigma(\boldsymbol{x})&=\int_{e}vq\,\mathrm{d}\sigma(\boldsymbol{x})&&\quad\text{for all edges $e$ and\,}q\in\mathbb{P}_{r-2}(e),\\ \int_{e}\partial_{\boldsymbol{n}}\pi_{K}vq\,\mathrm{d}\sigma(\boldsymbol{x})&=\int_{e}\partial_{\boldsymbol{n}}vq\,\mathrm{d}\sigma(\boldsymbol{x})&&\quad\text{for all edges $e$ and\,}q\in\mathbb{P}_{r-2}(e),\\ \int_{K}\pi_{K}vq\,\mathrm{d}\boldsymbol{x}&=\int_{K}vq\,\mathrm{d}\boldsymbol{x}&&\quad\text{for all\,}q\in\mathbb{P}_{r-2}(K).\end{aligned}\right.
Lemma 3.2.

There exists CC independent of hKh_{K} such that for vHk(K)v\in H^{k}(K) with 2kr+12\leq k\leq r+1, there holds

(3.6) j(vπKv)L2(K)ChKkjkvL2(K),\|\,\nabla^{j}(v-\pi_{K}v)\,\|_{L^{2}(K)}\leq Ch_{K}^{k-j}\|\,\nabla^{k}v\,\|_{L^{2}(K)},

where 0jk0\leq j\leq k.

Proof.

For any vr(K)PKv\in\mathbb{P}_{r}(K)\subset P_{K}, the definition (3.5) shows that vπKvPKv-\pi_{K}v\in P_{K} and all DoFs of vPKvv-P_{K}v vanish, then v=πKvv=\pi_{K}v. The estimate (3.6) immediately follows from the r(K)\mathbb{P}_{r}(K)-invariance of the local interpolation operator πK\pi_{K} [CiarletRaviart:1972]. ∎

Remark 3.3.

The element has a natural extension to three-dimensions by specifying KK as a tetrahedron, and

PK:=r(K)+bKi=14biQir2(K)+bK2Rr2(K),P_{K}{:}=\mathbb{P}_{r}(K)+b_{K}\sum_{i=1}^{4}b_{i}Q_{i}^{r-2}(K)+b_{K}^{2}R^{r-2}(K),

where bK=i=14λib_{K}=\prod_{i=1}^{4}\lambda_{i} is the element bubble function with λi\lambda_{i} the barycentric coordinates associated with the vertices aia_{i} for i=1,,4i=1,\cdots,4. bi=bK/λib_{i}=b_{K}/\lambda_{i} is the face bubble function associated with the face fif_{i}.

Define

Qir2(K):={vr2(K)KbKbivqd𝒙=0 for all qr3(K)},Q_{i}^{r-2}(K){:}=\left\{\,v\in\mathbb{P}_{r-2}(K)\,\mid\,\int_{K}b_{K}b_{i}vq\,\mathrm{d}\boldsymbol{x}=0\textup{ for all }q\in\mathbb{P}_{r-3}(K)\,\right\},

and

Rr2(K):={vr2(K)KbK2vqd𝒙=0 for all qr4(K)}.R^{r-2}(K){:}=\left\{\,v\in\mathbb{P}_{r-2}(K)\,\mid\,\int_{K}b_{K}^{2}vq\,\mathrm{d}\boldsymbol{x}=0\textup{ for all }q\in\mathbb{P}_{r-4}(K)\,\right\}.

The DoFs for PKP_{K} are given by

v{v(a)for all vertices,evqdσ(𝒙)for all edges e and qr2(e),fvqdσ(𝒙)for all faces f and qr3(f),f𝒏vqdσ(𝒙)for all faces f and qr2(f),Kvqd𝒙for all qr2(K).v\mapsto\left\{\begin{aligned} v(a)&\quad\text{for all vertices},\\ \int_{e}vq\,\mathrm{d}\sigma(\boldsymbol{x})&\quad\text{for all edges $e$ and\,}q\in\mathbb{P}_{r-2}(e),\\ \int_{f}vq\,\mathrm{d}\sigma(\boldsymbol{x})&\quad\text{for all faces $f$ and\,}q\in\mathbb{P}_{r-3}(f),\\ \int_{f}\partial_{\boldsymbol{n}}vq\,\mathrm{d}\sigma(\boldsymbol{x})&\quad\text{for all faces $f$ and\,}q\in\mathbb{P}_{r-2}(f),\\ \int_{K}vq\,\mathrm{d}\boldsymbol{x}&\quad\text{for all\,}q\in\mathbb{P}_{r-2}(K).\end{aligned}\right.

Similar to Lemma 3.1, the set (K,PK,ΣK)(K,P_{K},\Sigma_{K}) is also unisolvent.

3.2. Explicit representation for the bubble space

We clarify the structures of (3.2) and (3.3) associated with the set of DoFs (3.4)3 and the subset of (3.4)4 respectively, and derive the explicit formulations of the corresponding shape functions, which seems missing in the literature, while such explicit representations are useful for implementation. We firstly recall the following facts about the Jacobi polynomials [Szego:1975]. For any α,β>1\alpha,\beta>-1 and nonnegative integers n,mn,m, there holds

(3.7) 11(1t)α(1+t)βPn(α,β)(t)Pm(α,β)(t)dt=hn(α,β)δnm,\int_{-1}^{1}(1-t)^{\alpha}(1+t)^{\beta}P_{n}^{(\alpha,\beta)}(t)P_{m}^{(\alpha,\beta)}(t)\mathrm{d}t=h_{n}^{(\alpha,\beta)}\delta_{nm},

where

hn(α,β)=11(1t)α(1+t)β[Pn(α,β)(t)]2dt.h_{n}^{(\alpha,\beta)}=\int_{-1}^{1}(1-t)^{\alpha}(1+t)^{\beta}\bigl{[}P_{n}^{(\alpha,\beta)}(t)\bigr{]}^{2}\mathrm{d}t.

By [Szego:1975]*Eq. (4.3.3), we may write

(3.8) hn(α,β)=2α+β+12n+α+β+1Γ(n+α+1)Γ(n+β+1)Γ(n+α+β+1)Γ(n+1),h_{n}^{(\alpha,\beta)}=\dfrac{2^{\alpha+\beta+1}}{2n+\alpha+\beta+1}\dfrac{\Gamma(n+\alpha+1)\Gamma(n+\beta+1)}{\Gamma(n+\alpha+\beta+1)\Gamma(n+1)},

where Γ\Gamma is the Gamma function.

One of the explicit form for Pn(α,β)P_{n}^{(\alpha,\beta)} is

(1t)α(1+t)βPn(α,β)(t)=(1)n2nn!dndtn((1t)n+α(1+t)n+β).(1-t)^{\alpha}(1+t)^{\beta}P_{n}^{(\alpha,\beta)}(t)=\dfrac{(-1)^{n}}{2^{n}n!}\dfrac{\mathrm{d}^{n}}{\mathrm{d}t^{n}}\left((1-t)^{n+\alpha}(1+t)^{n+\beta}\right).

In particular,

P0(α,β)(t)=1,P1(α,β)(t)=12(α+β+2)t+12(αβ).P_{0}^{(\alpha,\beta)}(t)=1,\quad P_{1}^{(\alpha,\beta)}(t)=\dfrac{1}{2}(\alpha+\beta+2)t+\dfrac{1}{2}(\alpha-\beta).

Next we list certain facts about the Jacobi polynomials on the triangle [Dunkl:2014]*Section 2.4. For a triangle KK with vertices a1,a2,a3a_{1},a_{2},a_{3}, any point xKx\in K is uniquely expressed as

x=λ1a1+λ2a2+λ3a3,λi0 and λ1+λ2+λ3=1.x=\lambda_{1}a_{1}+\lambda_{2}a_{2}+\lambda_{3}a_{3},\quad\lambda_{i}\geq 0\text{ and }\lambda_{1}+\lambda_{2}+\lambda_{3}=1.

Then (λ1,λ2,λ3)(\lambda_{1},\lambda_{2},\lambda_{3}) is the barycentric coordinates of the point xx with respect to KK. For nonnegative integers k,nk,n such that knk\leq n, we define

(3.9) Pk,n(α,β,γ)(λ1,λ2,λ3):=(λ2+λ3)kPnk(2k+β+γ+1,α)(λ1λ2λ3)Pk(γ,β)((λ2λ3)/(λ2+λ3)).P_{k,n}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},\lambda_{3}){:}=(\lambda_{2}+\lambda_{3})^{k}P_{n-k}^{(2k+\beta+\gamma+1,\alpha)}(\lambda_{1}-\lambda_{2}-\lambda_{3})P_{k}^{(\gamma,\beta)}((\lambda_{2}-\lambda_{3})/(\lambda_{2}+\lambda_{3})).

It is straightforward to verify Pk,n(α,β,γ)n(K)P_{k,n}^{(\alpha,\beta,\gamma)}\in\mathbb{P}_{n}(K). In particular,

P0,0(α,β,γ)(λ1,λ2,λ3)=1,P0,1(α,β,γ)(λ1,λ2,λ3)=(β+γ+2)λ1(α+1)(λ2+λ3),P_{0,0}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},\lambda_{3})=1,\quad P_{0,1}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},\lambda_{3})=(\beta+\gamma+2)\lambda_{1}-(\alpha+1)(\lambda_{2}+\lambda_{3}),

and

P1,1(α,β,γ)(λ1,λ2,λ3)=(γ+1)λ2(β+1)λ3.P_{1,1}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},\lambda_{3})=(\gamma+1)\lambda_{2}-(\beta+1)\lambda_{3}.

For all α,β,γ>1\alpha,\beta,\gamma>-1, and nonnegative integers j,k,m,nj,k,m,n such that jmj\leq m and knk\leq n, there holds

(3.10) Kλ1αλ2βλ3γPk,n(α,β,γ)(λ1,λ2,λ3)Pj,m(α,β,γ)(λ1,λ2,λ3)d𝒙\displaystyle{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}\lambda_{1}^{\alpha}\lambda_{2}^{\beta}\lambda_{3}^{\gamma}P_{k,n}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},\lambda_{3})P_{j,m}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},\lambda_{3})\,\mathrm{d}\boldsymbol{x}
=\displaystyle= 2K^λ1αλ2β(1λ1λ2)γPk,n(α,β,γ)(λ1,λ2,1λ1λ2)Pj,m(α,β,γ)(λ1,λ2,1λ1λ2)dλ1dλ2\displaystyle 2\int_{\widehat{K}}\lambda_{1}^{\alpha}\lambda_{2}^{\beta}(1-\lambda_{1}-\lambda_{2})^{\gamma}P_{k,n}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},1-\lambda_{1}-\lambda_{2})P_{j,m}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},1-\lambda_{1}-\lambda_{2})\mathrm{d}\lambda_{1}\mathrm{d}\lambda_{2}
=\displaystyle= 2hk,n(α,β,γ)δjkδmn,\displaystyle 2h_{k,n}^{(\alpha,\beta,\gamma)}\delta_{jk}\delta_{mn},

where K^:={(λ1,λ2)λ10,λ20,λ1+λ21}\widehat{K}:=\left\{\,(\lambda_{1},\lambda_{2})\,\mid\,\lambda_{1}\geq 0,\lambda_{2}\geq 0,\lambda_{1}+\lambda_{2}\leq 1\,\right\} is the standard reference triangle and

(3.11) hk,n(α,β,γ)=\displaystyle h_{k,n}^{(\alpha,\beta,\gamma)}= 2(2k+α+2β+2γ+3)hnk(2k+β+γ+1,α)hk(γ,β)\displaystyle 2^{-(2k+\alpha+2\beta+2\gamma+3)}h_{n-k}^{(2k+\beta+\gamma+1,\alpha)}h_{k}^{(\gamma,\beta)}
=\displaystyle= 1(2n+α+β+γ+2)(2k+β+γ+1)\displaystyle\dfrac{1}{(2n+\alpha+\beta+\gamma+2)(2k+\beta+\gamma+1)}
×Γ(nk+α+1)Γ(n+k+β+γ+2)Γ(k+β+1)Γ(k+γ+1)Γ(nk+1)Γ(n+k+α+β+γ+2)Γ(k+1)Γ(k+β+γ+1).\displaystyle\times\dfrac{\Gamma(n-k+\alpha+1)\Gamma(n+k+\beta+\gamma+2)\Gamma(k+\beta+1)\Gamma(k+\gamma+1)}{\Gamma(n-k+1)\Gamma(n+k+\alpha+\beta+\gamma+2)\Gamma(k+1)\Gamma(k+\beta+\gamma+1)}.

Using the notation (x)n=Γ(x+n)/Γ(n)(x)_{n}=\Gamma(x+n)/\Gamma(n), we may find that the expression (3.11) is equivalent to the one in [Dunkl:2014]*Eq. (2.4.3). The identity (3.10) illustrates that {Pk,n(α,β,γ)(λ1,λ2,λ3)0knr}\{P_{k,n}^{(\alpha,\beta,\gamma)}(\lambda_{1},\lambda_{2},\lambda_{3})\mid 0\leq k\leq n\leq r\} are mutually orthogonal bases of r(K)\mathbb{P}_{r}(K) with respect to the weight λ1αλ2βλ3γ\lambda_{1}^{\alpha}\lambda_{2}^{\beta}\lambda_{3}^{\gamma}.

Next we study the structure of the bubble spaces. For the barycentric coordinate function λi\lambda_{i} such that λi0\lambda_{i}\equiv 0 on eie_{i}, let λi+\lambda_{i}^{+} and λi\lambda_{i}^{-} be the two other barycentric coordinates associated with the edges ei+e_{i}^{+} and eie_{i}^{-}, respectively. (ei,ei+,ei)(e_{i},e_{i}^{+},e_{i}^{-}) are chosen in a counterclockwise manner. The space Qir2(K)Q_{i}^{r-2}(K) can be clarified by the Jacobi polynomials with respect to the weight bKbib_{K}b_{i}, while Rr2(K)R^{r-2}(K) can be clarified by the Jacobi polynomials with respect to the weight bK2b_{K}^{2}, which are formulated in the following lemmas.

Lemma 3.4.

The space Qir2(K)Q_{i}^{r-2}(K) takes the form

Qir2(K)=span{Pk,r2(1,2,2)(λi,λi+,λi) 0kr2}.Q_{i}^{r-2}(K)=\textup{span}\left\{\,P_{k,r-2}^{(1,2,2)}(\lambda_{i},\lambda_{i}^{+},\lambda_{i}^{-})\,\mid\,0\leq k\leq r-2\,\right\}.
Proof.

For any vQir2(K)r2(K)v\in Q_{i}^{r-2}(K)\subset\mathbb{P}_{r-2}(K), vv may be expanded into

v=0knr2aknPk,n(1,2,2)(λi,λi+,λi)v=\sum_{0\leq k\leq n\leq r-2}a_{kn}P_{k,n}^{(1,2,2)}(\lambda_{i},\lambda_{i}^{+},\lambda_{i}^{-})

with unknown parameters akna_{kn}. Using the above representation, we may write the constraint in the definition (3.2) as

0knr2aknKbKbiPk,n(1,2,2)(λi,λi+,λi)qdx=0for all qr3(K).\sum_{0\leq k\leq n\leq r-2}a_{kn}{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}b_{K}b_{i}P_{k,n}^{(1,2,2)}(\lambda_{i},\lambda_{i}^{+},\lambda_{i}^{-})q\mathrm{d}x=0\qquad\text{for all\quad}q\in\mathbb{P}_{r-3}(K).

Substituting q=Pj,m(1,2,2)(λi,λi+,λi)q=P_{j,m}^{(1,2,2)}(\lambda_{i},\lambda_{i}^{+},\lambda_{i}^{-}) for 0jmr30\leq j\leq m\leq r-3 into the above equation, and using the orthogonal relation (3.10), we obtain ajm=0a_{jm}=0 for 0jmr30\leq j\leq m\leq r-3. This concludes the lemma. ∎

Motivated by the above lemma, we change the definition of DoFs for the bubble space bKi=13biQir2(K)b_{K}\sum_{i=1}^{3}b_{i}Q_{i}^{r-2}(K) from ei𝒏vqdσ(𝒙)\int_{e_{i}}\partial_{\boldsymbol{n}}vq\,\mathrm{d}\sigma(\boldsymbol{x}) for any qr2(ei)q\in\mathbb{P}_{r-2}(e_{i}) to

eivnPk(2,2)(λi+λ)dσ(𝒙),k=0,,r2.{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}\dfrac{\partial v}{\partial n}P_{k}^{(2,2)}(\lambda_{i}^{+}-\lambda^{-})\,\mathrm{d}\sigma(\boldsymbol{x}),\quad k=0,\cdots,r-2.
Lemma 3.5.

The shape functions for the bubble space bKi=13biQir2(K)b_{K}\sum_{i=1}^{3}b_{i}Q_{i}^{r-2}(K) associated with the above definition of DoFs are

ak,r2bKbiPk,r2(1,2,2)(λi,λi+,λi)a_{k,r-2}b_{K}b_{i}P_{k,r-2}^{(1,2,2)}(\lambda_{i},\lambda_{i}^{+},\lambda_{i}^{-})

with

(3.12) ak,r2=(1)rk1|λi|(k+3)(k+4)(2k+5)(rk1)(k+1)(k+2)k=0,,r2.a_{k,r-2}=\dfrac{(-1)^{r-k-1}}{\lvert\nabla\lambda_{i}\rvert}\dfrac{(k+3)(k+4)(2k+5)}{(r-k-1)(k+1)(k+2)}\qquad k=0,\cdots,r-2.
Proof.

A direct calculation gives

𝒏(bKbiPk,r2(1,2,2)(λi,λi+,λi))|ei=|λi|bi2Pr2k(2k+5,1)(1)Pk(2,2)(λi+λ),\dfrac{\partial}{\partial\boldsymbol{n}}\left(b_{K}b_{i}P_{k,r-2}^{(1,2,2)}(\lambda_{i},\lambda_{i}^{+},\lambda_{i}^{-})\right)|_{e_{i}}=-\lvert\nabla\lambda_{i}\rvert b_{i}^{2}P_{r-2-k}^{(2k+5,1)}(-1)P_{k}^{(2,2)}(\lambda_{i}^{+}-\lambda^{-}),

For j=1,,r2j=1,\cdots,r-2. Using the relation (3.7), and noting that bi=λi+λi=λi+(1λi+)b_{i}=\lambda_{i}^{+}\lambda_{i}^{-}=\lambda_{i}^{+}(1-\lambda_{i}^{+}) on eie_{i}, we obtain

ei𝒏(bKbiPk,r2(1,2,2)(λi,λi+,λi))Pj(2,2)(λi+λ)dσ(𝒙)\displaystyle{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}\dfrac{\partial}{\partial\boldsymbol{n}}\left(b_{K}b_{i}P_{k,r-2}^{(1,2,2)}(\lambda_{i},\lambda_{i}^{+},\lambda_{i}^{-})\right)P_{j}^{(2,2)}(\lambda_{i}^{+}-\lambda^{-})\,\mathrm{d}\sigma(\boldsymbol{x})
=\displaystyle= |λi|Pr2k(2k+5,1)(1)01(λi+(1λi+))2Pk(2,2)(2λi+1)Pj(2,2)(2λi+1)dλi+\displaystyle-\lvert\nabla\lambda_{i}\rvert P_{r-2-k}^{(2k+5,1)}(-1)\int_{0}^{1}\left(\lambda_{i}^{+}(1-\lambda_{i}^{+})\right)^{2}P_{k}^{(2,2)}(2\lambda_{i}^{+}-1)P_{j}^{(2,2)}(2\lambda_{i}^{+}-1)\mathrm{d}\lambda_{i}^{+}
=\displaystyle= |λi|32Pr2k(2k+5,1)(1)11(1t)2(1+t)2Pk(2,2)(t)Pj(2,2)(t)dt\displaystyle-\dfrac{\lvert\nabla\lambda_{i}\rvert}{32}P_{r-2-k}^{(2k+5,1)}(-1)\int_{-1}^{1}(1-t)^{2}(1+t)^{2}P_{k}^{(2,2)}(t)P_{j}^{(2,2)}(t)\mathrm{d}t
=\displaystyle= |λi|32Pr2k(2k+5,1)(1)hk(2,2)δjk.\displaystyle-\dfrac{\lvert\nabla\lambda_{i}\rvert}{32}P_{r-2-k}^{(2k+5,1)}(-1)h_{k}^{(2,2)}\delta_{jk}.

This gives

ak,r2=32|λi|Pr2k(2k+5,1)(1)hk(2,2).a_{k,r-2}=-\dfrac{32}{\lvert\nabla\lambda_{i}\rvert P_{r-2-k}^{(2k+5,1)}(-1)h_{k}^{(2,2)}}.

By [Szego:1975]*Eq. (4.1.1),(4.1.3), we obtain Pr2k(2k+5,1)(1)=(1)rk(rk1).P_{r-2-k}^{(2k+5,1)}(-1)=(-1)^{r-k}(r-k-1). Using (3.8), we obtain

hk(2,2)=32(k+1)(k+2)(k+3)(k+4)(2k+5).h_{k}^{(2,2)}=\dfrac{32(k+1)(k+2)}{(k+3)(k+4)(2k+5)}.

A combination of the above three identities leads to (3.12). ∎

Next we list the shape functions for the elements of low-order.

Example 3.6.

The bubble space bKi=13biQir2(K)b_{K}\sum_{i=1}^{3}b_{i}Q_{i}^{r-2}(K) for the lowest-order r=2r=2 is

bKspan{bii=1,2,3}.b_{K}\textup{span}\left\{\,b_{i}\,\mid\,i=1,2,3\,\right\}.

The shape functions associated with ei𝒏vdσ(𝒙){\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}\partial_{\boldsymbol{n}}v\,\mathrm{d}\sigma(\boldsymbol{x}) is 30|λi|bKbi.-\dfrac{30}{\lvert\nabla\lambda_{i}\rvert}b_{K}b_{i}.

The bubble space bKi=13biQir2(K)b_{K}\sum_{i=1}^{3}b_{i}Q_{i}^{r-2}(K) for the case r=3r=3 is

bKspan{bi(3λiλi+λi),bi(λi+λi)i=1,2,3}.b_{K}\textup{span}\left\{\,b_{i}(3\lambda_{i}-\lambda_{i}^{+}-\lambda_{i}^{-}),b_{i}(\lambda_{i}^{+}-\lambda_{i}^{-})\,\mid\,i=1,2,3\,\right\}.

The shape functions associated with ei𝒏vdσ(𝒙){\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}\partial_{\boldsymbol{n}}v\,\mathrm{d}\sigma(\boldsymbol{x}) is

30|λi|bKbi(3λiλi+λi).\dfrac{30}{\lvert\nabla\lambda_{i}\rvert}b_{K}b_{i}(3\lambda_{i}-\lambda_{i}^{+}-\lambda_{i}^{-}).

The shape functions associated with 3ei𝒏v(λi+λi)dσ(𝒙)3{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}\partial_{\boldsymbol{n}}v(\lambda_{i}^{+}-\lambda_{i}^{-})\,\mathrm{d}\sigma(\boldsymbol{x}) is

70|λi|bKbi(λi+λi).-\dfrac{70}{\lvert\nabla\lambda_{i}\rvert}b_{K}b_{i}(\lambda_{i}^{+}-\lambda_{i}^{-}).
Lemma 3.7.

The space Rr2(K)R^{r-2}(K) takes the forms

Rr2(K)=span{Pk,r2(2,2,2)(λ1,λ2,λ3) 0kr2}.R^{r-2}(K)=\textup{span}\left\{\,P_{k,r-2}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3})\,\mid\,0\leq k\leq r-2\,\right\}.
Proof.

For any vRr2(K)r2(K)v\in R^{r-2}(K)\subset\mathbb{P}_{r-2}(K), we expand vv into

v=0knr2aknPk,n(2,2,2)(λ1,λ2,λ3)v=\sum_{0\leq k\leq n\leq r-2}a_{kn}P_{k,n}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3})

with akna_{kn} to be determined later on. Using the above representation, we may write the constraint in the definition of the space (3.3) as: For all qr3(K)q\in\mathbb{P}_{r-3}(K),

0knr2aknKbK2Pk,n(2,2,2)(λ1,λ2,λ3)qdx=0.\sum_{0\leq k\leq n\leq r-2}a_{kn}{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}b_{K}^{2}P_{k,n}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3})q\mathrm{d}x=0.

We substitute q=Pj,m(2,2,2)(λ1,λ2,λ3)q=P_{j,m}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3}) for 0jmr30\leq j\leq m\leq r-3 into the above equation. Using the orthogonal relation (3.10), we may obtain ajm=0a_{jm}=0 for 0jmr30\leq j\leq m\leq r-3. This completes the proof. ∎

Motivated by the above lemma, we change the definition of DoFs for bK2Rr2(K)b_{K}^{2}R^{r-2}(K) from Kvqd𝒙\int_{K}vq\,\mathrm{d}\boldsymbol{x} for any qr2(K)r3(K)q\in\mathbb{P}_{r-2}(K)\setminus\mathbb{P}_{r-3}(K) to

KvPk,r2(2,2,2)(λ1,λ2,λ3)d𝒙,k=0,,r2.{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}vP_{k,r-2}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3})\,\mathrm{d}\boldsymbol{x},\quad k=0,\cdots,r-2.
Lemma 3.8.

The shape functions for the bubble space bK2Rr2(K)b_{K}^{2}R^{r-2}(K) associated with the above definition of DoFs are

ak,r2bK2Pk,r2(2,2,2)(λ1,λ2,λ3)a_{k,r-2}b_{K}^{2}P_{k,r-2}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3})

with

ak,r2=(2r+4)(2k+5)(r+k+4)(r+k+5)(k+3)(k+4)2(rk)(r1k)(k+1)(k+2)k=0,,r2.a_{k,r-2}=\dfrac{(2r+4)(2k+5)(r+k+4)(r+k+5)(k+3)(k+4)}{2(r-k)(r-1-k)(k+1)(k+2)}\qquad k=0,\cdots,r-2.
Proof.

For j=1,,r2j=1,\cdots,r-2, we obtain

KbK2Pk,r2(2,2,2)(λ1,λ2,λ3)Pj,r2(2,2,2)(λ1,λ2,λ3)d𝒙=2hk,r2(2,2,2)δjk,{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}b_{K}^{2}P_{k,r-2}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3})P_{j,r-2}^{(2,2,2)}(\lambda_{1},\lambda_{2},\lambda_{3})\,\mathrm{d}\boldsymbol{x}=2h_{k,r-2}^{(2,2,2)}\delta_{jk},

which gives

ak,r2=12hk,r2(2,2,2).a_{k,r-2}=\dfrac{1}{2h_{k,r-2}^{(2,2,2)}}.

Using (3.11), we obtain

hk,r2(2,2,2)=(rk)(r1k)(k+1)(k+2)(2r+4)(2k+5)(r+k+4)(r+k+5)(k+3)(k+4).h_{k,r-2}^{(2,2,2)}=\dfrac{(r-k)(r-1-k)(k+1)(k+2)}{(2r+4)(2k+5)(r+k+4)(r+k+5)(k+3)(k+4)}.

These give the simplified expression of ak,r2a_{k,r-2}. ∎

According to the definition of bubble space, we may have

Example 3.9.

The bubble space bK2Rr2(K)b_{K}^{2}R^{r-2}(K) for the lowest-order case r=2r=2 is span{bK2}\{b_{K}^{2}\}.

The shape functions associated with Kvd𝒙{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}v\,\mathrm{d}\boldsymbol{x} is 2520bK22520b_{K}^{2}.

The bubble space bK2Rr2(K)b_{K}^{2}R^{r-2}(K) for the case r=3r=3 is bK2b_{K}^{2}span{2λ1λ2λ3,λ2λ3}\{2\lambda_{1}-\lambda_{2}-\lambda_{3},\lambda_{2}-\lambda_{3}\}.

The shape functions associated with 3Kv(2λ1λ2λ3)d𝒙3\,{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}v(2\lambda_{1}-\lambda_{2}-\lambda_{3})\,\mathrm{d}\boldsymbol{x} is 4200bK2(2λ1λ2λ3)4200b_{K}^{2}(2\lambda_{1}-\lambda_{2}-\lambda_{3}).

The shape functions associated with 3Kv(λ2λ3)d𝒙3\,{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}v(\lambda_{2}-\lambda_{3})\,\mathrm{d}\boldsymbol{x} is 12600bK2(λ2λ3)12600b_{K}^{2}(\lambda_{2}-\lambda_{3}).

Remark 3.10.

Based on the above results, we give the details for the element of the lowest-order, i.e., r=2r=2, which have been used in the numerical examples. The local finite element space

PK=2(K)+bKi=13span{bi}+span{bK2},P_{K}=\mathbb{P}_{2}(K)+b_{K}\sum_{i=1}^{3}\text{span}\{b_{i}\}+\text{span}\{b_{K}^{2}\},

and DoFs

ΣK={v(ai),eivdσ(𝒙),ei𝒏vdσ(𝒙),Kvd𝒙i=1,2,3}.\Sigma_{K}=\left\{\,v(a_{i}),{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}v\,\mathrm{d}\sigma(\boldsymbol{x}),{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}\partial_{\boldsymbol{n}}v\,\mathrm{d}\sigma(\boldsymbol{x}),{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}v\,\mathrm{d}\boldsymbol{x}\,\mid\,i=1,2,3\,\right\}.

The shape functions associated with {v(ai)}i=1,2,3\{v(a_{i})\}_{i=1,2,3} are

ϕi=λi(3λi2)+30bK(2bi+jiλiλj|λj|2bj(4λj1)+6bK).\phi_{i}=\lambda_{i}(3\lambda_{i}-2)+30b_{K}\left(2b_{i}+\sum_{j\neq i}\dfrac{\nabla\lambda_{i}\cdot\nabla\lambda_{j}}{\lvert\nabla\lambda_{j}\rvert^{2}}b_{j}(4\lambda_{j}-1)+6b_{K}\right).

The shape functions associated with {eivdσ(𝒙)}i=1,2,3\{{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}v\mathrm{d}\sigma(\boldsymbol{x})\}_{i=1,2,3} are

φi=6bi+90bK(bijibj10bK).\varphi_{i}=6b_{i}+90b_{K}\left(b_{i}-\sum_{j\neq i}b_{j}-10b_{K}\right).

The shape functions associated with {ei𝒏vdσ(𝒙)}i=1,2,3\{{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{e_{i}}\partial_{\boldsymbol{n}}v\mathrm{d}\sigma(\boldsymbol{x})\}_{i=1,2,3} are

ψi=30|λi|bKbi(4λi1).\psi_{i}=\dfrac{30}{\lvert\nabla\lambda_{i}\rvert}b_{K}b_{i}(4\lambda_{i}-1).

The shape functions associated with Kvd𝒙{\int\negthickspace\negthickspace\negthickspace\negthinspace-}_{K}v\,\mathrm{d}\boldsymbol{x} is ϕ0=2520bK2\phi_{0}=2520b_{K}^{2}.

4. The Mixed Finite Elements Approximation

In this part, we construct stable finite element pairs to approximate (2.2). We ignore the influence of the curved boundary for error estimates for brevity. Define

Xh:={vH01(Ω)v|KPK for all K𝒯h,e[[𝒏v]]qdσ(𝒙)=0 for all eh,qr2(e)},X_{h}{:}=\left\{\,v\in H_{0}^{1}(\Omega)\,\mid\,v|_{K}\in P_{K}\text{ for all }K\in\mathcal{T}_{h},\quad\int_{e}[\![\partial_{\boldsymbol{n}}v]\!]q\,\mathrm{d}\sigma(\boldsymbol{x})=0\text{ for all }e\in\mathcal{E}_{h},q\in\mathbb{P}_{r-2}(e)\,\right\},

and Vh:=[Xh]2V_{h}{:}=[X_{h}]^{2}. Let PhPP_{h}\subset P be the continuous Lagrangian finite element of order r1r-1. We shall prove a uniform discrete B-B inequality for the pair (Vh,Ph)(V_{h},P_{h}).

The following rescaled trace inequality will be used later on: There exists CC independent of hKh_{K} such that

(4.1) vL2(K)C(hK1/2vL2(K)+vL2(K)1/2𝒗L2(K)1/2).\|\,v\,\|_{L^{2}(\partial K)}\leq C\left(h_{K}^{-1/2}\|\,v\,\|_{L^{2}(K)}+\|\,v\,\|_{L^{2}(K)}^{1/2}\|\,\nabla\boldsymbol{v}\,\|_{L^{2}(K)}^{1/2}\right).

This inequality may be found in [BrennerScott:2008].

4.1. The mixed finite element approximation

We define the mixed finite element approximation problem as follows. Find 𝒖hVh\boldsymbol{u}_{h}\in V_{h} and phPhp_{h}\in P_{h} such that

(4.2) {aι,h(𝒖h,𝒗)+bι,h(𝒗,ph)=(𝒇,𝒗)for all 𝒗Vh,bι,h(𝒖h,q)λ1cι(ph,q)=0for all qPh,\left\{\begin{aligned} a_{\iota,h}(\boldsymbol{u}_{h},\boldsymbol{v})+b_{\iota,h}(\boldsymbol{v},p_{h})&=(\boldsymbol{f},\boldsymbol{v})\qquad&&\text{for all\quad}\boldsymbol{v}\in V_{h},\\ b_{\iota,h}(\boldsymbol{u}_{h},q)-\lambda^{-1}c_{\iota}(p_{h},q)&=0\qquad&&\text{for all\quad}q\in P_{h},\end{aligned}\right.

where

aι,h(𝒗,𝒘):\displaystyle a_{\iota,h}(\boldsymbol{v},\boldsymbol{w}){:} =2μ((ϵ(𝒗),ϵ(𝒘))+ι2(hϵ(𝒗),hϵ(𝒘)))\displaystyle=2\mu\Bigl{(}(\boldsymbol{\epsilon}(\boldsymbol{v}),\boldsymbol{\epsilon}(\boldsymbol{w}))+\iota^{2}(\nabla_{h}\boldsymbol{\epsilon}(\boldsymbol{v}),\nabla_{h}\boldsymbol{\epsilon}(\boldsymbol{w}))\Bigr{)}\quad for all 𝒗,𝒘Vh,\displaystyle\text{for all\,}\boldsymbol{v},\boldsymbol{w}\in V_{h},
bι,h(𝒗,q):\displaystyle b_{\iota,h}(\boldsymbol{v},q){:} =(div𝒗,q)+ι2(hdiv𝒗,q)\displaystyle=(\operatorname{div}\boldsymbol{v},q)+\iota^{2}(\nabla_{h}\operatorname{div}\boldsymbol{v},\nabla q)\quad for all 𝒗Vh,qPh.\displaystyle\text{for all\,}\boldsymbol{v}\in V_{h},q\in P_{h}.

Note that VhVV_{h}\not\subset V, and this is a nonconforming method, we introduce the broken norm

𝒗ι,h:=𝒗L2+ιh2𝒗L2for all𝒗Vh\|\,\nabla\boldsymbol{v}\,\|_{\iota,h}{:}=\|\,\nabla\boldsymbol{v}\,\|_{L^{2}}+\iota\|\,\nabla_{h}^{2}\boldsymbol{v}\,\|_{L^{2}}\qquad\text{for all}\quad\boldsymbol{v}\in V_{h}

Due to the continuity of 𝒗\boldsymbol{v}, 𝒗ι,h\|\,\nabla\boldsymbol{v}\,\|_{\iota,h} is a norm over VhV_{h}.

The following broken Korn’s inequality was proved in [LiMingWang:2021]*Theorem 2:

hϵ(𝒗)L2(11/2)h2𝒗L2,\|\,\nabla_{h}\boldsymbol{\epsilon}(\boldsymbol{v})\,\|_{L^{2}}\geq\left(1-1/\sqrt{2}\right)\|\,\nabla_{h}^{2}\boldsymbol{v}\,\|_{L^{2}},

which together with the first Korn’s inequality (2.6) gives

(4.3) aι,h(𝒗,𝒗)μ2𝒗ι,h2for all 𝒗Vh.a_{\iota,h}(\boldsymbol{v},\boldsymbol{v})\geq\dfrac{\mu}{2}\|\,\nabla\boldsymbol{v}\,\|_{\iota,h}^{2}\qquad\text{for all }\boldsymbol{v}\in V_{h}.

It remains to prove the discrete B-B inequality for the pair (Vh,Ph)(V_{h},P_{h}). To this end, we construct a Fortin operator that is uniformly stable in the weighted norm ι,h\|\,\nabla\cdot\,\|_{\iota,h} [Winther:2013]. The key is to construct different Fortin operators for ι/h\iota/h in different regimes.

Firstly we define an interpolation operator Πh:VVh\varPi_{h}:V\to V_{h} by Πh|K:=ΠK=[πK]2\varPi_{h}|_{K}{:}=\varPi_{K}=[\pi_{K}]^{2}, which satisfies

Lemma 4.1.

For all 𝐯V\boldsymbol{v}\in V, there holds

(4.4) bι,h(Πh𝒗,p)=bι(𝒗,p)for all pPh.b_{\iota,h}(\varPi_{h}\boldsymbol{v},p)=b_{\iota}(\boldsymbol{v},p)\qquad\text{for all\;}p\in P_{h}.
Proof.

Using the fact that Πh𝒗Vh[H01(Ω)]2\varPi_{h}\boldsymbol{v}\in V_{h}\subset[H_{0}^{1}(\Omega)]^{2}, an integration by parts gives

(4.5) Ωdiv(𝒗Πh𝒗)pd𝒙=K𝒯hK(𝒗ΠK𝒗)pd𝒙=0,\int_{\Omega}\operatorname{div}(\boldsymbol{v}-\varPi_{h}\boldsymbol{v})p\,\mathrm{d}\boldsymbol{x}=-\sum_{K\in\mathcal{T}_{h}}\int_{K}(\boldsymbol{v}-\varPi_{K}\boldsymbol{v})\cdot\nabla p\,\mathrm{d}\boldsymbol{x}=0,

where we have used the identity (3.5)4 in the last step.

Next, integration by parts yields

Ωdiv(𝒗Πh𝒗)pd𝒙=K𝒯hKdiv(𝒗ΠK𝒗)𝒏pdσ(𝒙)K𝒯hKdiv(𝒗ΠK𝒗)Δpd𝒙=0.\int_{\Omega}\nabla\operatorname{div}(\boldsymbol{v}-\varPi_{h}\boldsymbol{v})\cdot\nabla p\,\mathrm{d}\boldsymbol{x}=\sum_{K\in\mathcal{T}_{h}}\int_{\partial K}\operatorname{div}(\boldsymbol{v}-\varPi_{K}\boldsymbol{v})\partial_{\boldsymbol{n}}p\,\mathrm{d}\sigma(\boldsymbol{x})-\sum_{K\in\mathcal{T}_{h}}\int_{K}\operatorname{div}(\boldsymbol{v}-\varPi_{K}\boldsymbol{v})\Delta p\,\mathrm{d}\boldsymbol{x}=0.

The first term vanishes because j=tj𝒕+nj𝒏\partial_{j}=t_{j}\partial_{\boldsymbol{t}}+n_{j}\partial_{\boldsymbol{n}} for components j=1,2j=1,2, and using (3.5)2, we obtain that for each edge eKe\in\partial K,

etj𝒕(vjπKvj)𝒏pdσ(𝒙)=etj(vjπKvj)2ptndσ(𝒙)=0.\int_{e}t_{j}\dfrac{\partial}{\partial\boldsymbol{t}}(v_{j}-\pi_{K}v_{j})\partial_{\boldsymbol{n}}p\,\mathrm{d}\sigma(\boldsymbol{x})=-\int_{e}t_{j}(v_{j}-\pi_{K}v_{j})\dfrac{\partial^{2}p}{\partial t\partial n}\,\mathrm{d}\sigma(\boldsymbol{x})=0.

Using (3.5)3, we obtain

enj𝒏(vjπKvj)𝒏pdσ(𝒙)=0.\int_{e}n_{j}\dfrac{\partial}{\partial\boldsymbol{n}}(v_{j}-\pi_{K}v_{j})\partial_{\boldsymbol{n}}p\,\mathrm{d}\sigma(\boldsymbol{x})=0.

While the second term vanishes because

Kdiv(𝒗ΠK𝒗)Δpd𝒙=K(𝒗ΠK𝒗)𝒏Δpdσ(𝒙)K(𝒗ΠK𝒗)Δpd𝒙=0,\int_{K}\operatorname{div}(\boldsymbol{v}-\varPi_{K}\boldsymbol{v})\Delta p\,\mathrm{d}\boldsymbol{x}=\int_{\partial K}(\boldsymbol{v}-\varPi_{K}\boldsymbol{v})\cdot\boldsymbol{n}\Delta p\,\mathrm{d}\sigma(\boldsymbol{x})-\int_{K}(\boldsymbol{v}-\varPi_{K}\boldsymbol{v})\cdot\nabla\Delta p\,\mathrm{d}\boldsymbol{x}=0,

where we have used (3.5)2 and (3.5)4. ∎

The operator Πh\varPi_{h} is not H1{}^{1}-bounded by (3.6), and we construct an H1{}^{1}-bounded regularized interpolation operator as follows.

Lemma 4.2.

There exists an operator Ih:VVhI_{h}:\,V\mapsto V_{h} satisfying

(4.6) Ωdiv(𝒗Ih𝒗)pd𝒙=0for all pPh,\int_{\Omega}\operatorname{div}(\boldsymbol{v}-I_{h}\boldsymbol{v})p\,\mathrm{d}\boldsymbol{x}=0\qquad\text{for all\quad}p\in P_{h},

and if 𝐯V[Hs(Ω)]2\boldsymbol{v}\in V\cap[H^{s}(\Omega)]^{2} with 1sr+11\leq s\leq r+1, then

(4.7) hj(𝒗Ih𝒗)L2Chsjs𝒗L20js.\|\,\nabla^{j}_{h}(\boldsymbol{v}-I_{h}\boldsymbol{v})\,\|_{L^{2}}\leq Ch^{s-j}\|\,\nabla^{s}\boldsymbol{v}\,\|_{L^{2}}\qquad 0\leq j\leq s.

The construction of IhI_{h} is based on a regularized interpolation operator in [GuzmanLeykekhmanNeilan:2012] and the standard construction of the Fortin operator [Fortin:1977]. The operator IhI_{h} is also well-defined for functions in [H2(Ω)H01(Ω)]2[H^{2}(\Omega)\cap H_{0}^{1}(\Omega)]^{2}.

Proof.

Define Ih:VVhI_{h}\,:V\mapsto V_{h} with Ih:=[Π1]2I_{h}{:}=[\varPi_{1}]^{2} and

Π1:=Π0(IΠ2)+Π2,\varPi_{1}{:}=\varPi_{0}(I-\varPi_{2})+\varPi_{2},

where the regularized interpolation operator Π2:H02(Ω)Xh\varPi_{2}:H_{0}^{2}(\Omega)\mapsto X_{h} was constructed in [GuzmanLeykekhmanNeilan:2012]*Lemma 2, which satisfies

(4.8) hj(vΠ2v)L2ChsjsvL2,1sr+1,0js.\|\,\nabla^{j}_{h}(v-\varPi_{2}v)\,\|_{L^{2}}\leq Ch^{s-j}\|\,\nabla^{s}v\,\|_{L^{2}},\qquad 1\leq s\leq r+1,0\leq j\leq s.

The operator Π0:H01(Ω)Xh\varPi_{0}\,:H_{0}^{1}(\Omega)\mapsto X_{h} is defined for any element K𝒯hK\in\mathcal{T}_{h} as

{Π0v(a)=0for all verticesa,eΠ0vqdσ(𝒙)=0qr2(e)for all edgese,e𝒏Π0vqdσ(𝒙)=0qr2(e)for all edgese,KΠ0vqd𝒙=Kvqd𝒙for all qr2(K).\left\{\begin{aligned} \varPi_{0}v(a)&=0\quad\text{for all vertices}\,a,\\ \int_{e}\varPi_{0}vq\,\mathrm{d}\sigma(\boldsymbol{x})&=0\quad\forall q\in\mathbb{P}_{r-2}(e)\quad\text{for all edges}\,e,\\ \int_{e}\partial_{\boldsymbol{n}}\varPi_{0}vq\,\mathrm{d}\sigma(\boldsymbol{x})&=0\quad\forall q\in\mathbb{P}_{r-2}(e)\quad\text{for all edges}\,e,\\ \int_{K}\varPi_{0}vq\,\mathrm{d}\boldsymbol{x}&=\int_{K}vq\,\mathrm{d}\boldsymbol{x}\quad\text{for all\quad}q\in\mathbb{P}_{r-2}(K).\end{aligned}\right.

On each element KK, we have (Π0v)|KH01(K)(\varPi_{0}v)|_{K}\in H_{0}^{1}(K) for any vH01(Ω)v\in H_{0}^{1}(\Omega), and hence Π0vXh\varPi_{0}v\in X_{h}. A standard scaling argument gives

Π0vL2CvL2for all vH01(Ω).\|\,\varPi_{0}v\,\|_{L^{2}}\leq C\|\,v\,\|_{L^{2}}\qquad\text{for all }v\in H_{0}^{1}(\Omega).

For any 𝒗V\boldsymbol{v}\in V, using the fact that Ih𝒗VhI_{h}\boldsymbol{v}\in V_{h} and pPhp\in P_{h}, an integration by parts gives

Ωdiv(𝒗Ih𝒗)pd𝒙\displaystyle\int_{\Omega}\operatorname{div}(\boldsymbol{v}-I_{h}\boldsymbol{v})p\,\mathrm{d}\boldsymbol{x} =K𝒯hK(𝒗Ih𝒗)pd𝒙+K𝒯hK(𝒗Ih𝒗)𝒏pdσ(𝒙)\displaystyle=-\sum_{K\in\mathcal{T}_{h}}\int_{K}(\boldsymbol{v}-I_{h}\boldsymbol{v})\cdot\nabla p\,\mathrm{d}\boldsymbol{x}+\sum_{K\in\mathcal{T}_{h}}\int_{\partial K}(\boldsymbol{v}-I_{h}\boldsymbol{v})\cdot\boldsymbol{n}p\,\mathrm{d}\sigma(\boldsymbol{x})
=K𝒯hK(𝒗Ih𝒗)pd𝒙=K𝒯hK(IΠ0)(IΠ2)vpd𝒙\displaystyle=-\sum_{K\in\mathcal{T}_{h}}\int_{K}(\boldsymbol{v}-I_{h}\boldsymbol{v})\cdot\nabla p\,\mathrm{d}\boldsymbol{x}=-\sum_{K\in\mathcal{T}_{h}}\int_{K}(I-\varPi_{0})(I-\varPi_{2})v\cdot\nabla p\,\mathrm{d}\boldsymbol{x}
=0,\displaystyle=0,

where we have used the last property of Π0\varPi_{0}. This gives (4.6).

Using (4.8), the L2-stability of Π0\varPi_{0} and the inverse inequality, we obtain

hj(𝒗Π1𝒗)L2\displaystyle\|\,\nabla^{j}_{h}(\boldsymbol{v}-\varPi_{1}\boldsymbol{v})\,\|_{L^{2}}\leq hj(𝒗Π2𝒗)L2+hjΠ0(𝒗Π2𝒗)L2\displaystyle\|\,\nabla^{j}_{h}(\boldsymbol{v}-\varPi_{2}\boldsymbol{v})\,\|_{L^{2}}+\|\,\nabla^{j}_{h}\varPi_{0}(\boldsymbol{v}-\varPi_{2}\boldsymbol{v})\,\|_{L^{2}}
\displaystyle\leq Chsjs𝒗L2+Chj𝒗Π2𝒗L2\displaystyle Ch^{s-j}\|\,\nabla^{s}\boldsymbol{v}\,\|_{L^{2}}+Ch^{-j}\|\,\boldsymbol{v}-\varPi_{2}\boldsymbol{v}\,\|_{L^{2}}
\displaystyle\leq Chsjs𝒗L2.\displaystyle Ch^{s-j}\|\,\nabla^{s}\boldsymbol{v}\,\|_{L^{2}}.

This implies (4.7) and completes the proof. ∎

We are ready to prove the following discrete B-B inequality.

Theorem 4.3.

There exists β\beta independent of ι\iota and hh, such that

(4.9) sup𝒗Vhbι,h(𝒗,p)𝒗ι,hβpιfor all pPh.\sup_{\boldsymbol{v}\in V_{h}}\dfrac{b_{\iota,h}(\boldsymbol{v},p)}{\|\,\nabla\boldsymbol{v}\,\|_{\iota,h}}\geq\beta\|p\|_{\iota}\qquad\text{for all\quad}p\in P_{h}.
Proof.

Using (2.3), for any pPhPp\in P_{h}\subset P, there exists 𝒗0V\boldsymbol{v}_{0}\in V such that

bι(𝒗0,p)=pι2and 𝒗0ιCpι.b_{\iota}(\boldsymbol{v}_{0},p)=\|p\|_{\iota}^{2}\qquad\text{and\quad}\|\nabla\boldsymbol{v}_{0}\|_{\iota}\leq C\|p\|_{\iota}.

First, we consider the case ι/hγ\iota/h\leq\gamma with γ\gamma to be determined later on. By (4.7), we obtain

Ih𝒗0ι,h𝒗0ι+(𝒗0Ih𝒗0)ι,hC𝒗0ι,\|\,\nabla I_{h}\boldsymbol{v}_{0}\,\|_{\iota,h}\leq\|\nabla\boldsymbol{v}_{0}\|_{\iota}+\|\,\nabla(\boldsymbol{v}_{0}-I_{h}\boldsymbol{v}_{0})\,\|_{\iota,h}\leq C\|\nabla\boldsymbol{v}_{0}\|_{\iota},

and

hdiv(𝒗0Ih𝒗0)L2C2𝒗0L2CpL2.\|\,\nabla_{h}\operatorname{div}(\boldsymbol{v}_{0}-I_{h}\boldsymbol{v}_{0})\,\|_{L^{2}}\leq C\|\,\nabla^{2}\boldsymbol{v}_{0}\,\|_{L^{2}}\leq C\|\,\nabla p\,\|_{L^{2}}.

Combining the above inequality and using the inverse inequality for any pPhp\in P_{h}, we obtain

ι2|(hdiv(𝒗0Ih𝒗0),p)|Cι2pL22C(ι/h)2pL22γ2Cpι2.\iota^{2}\lvert(\nabla_{h}\operatorname{div}(\boldsymbol{v}_{0}-I_{h}\boldsymbol{v}_{0}),\nabla p)\rvert\leq C\iota^{2}\|\,\nabla p\,\|_{L^{2}}^{2}\leq C_{\ast}(\iota/h)^{2}\|\,p\,\|_{L^{2}}^{2}\leq\gamma^{2}C_{\ast}\|p\|_{\iota}^{2}.

Fix γ\gamma such that γ2C<1\gamma^{2}C_{\ast}<1, we obtain

bι,h(Ih𝒗0,p)=bι(𝒗0,p)ι2(hdiv(𝒗0Ih𝒗0),p)(1γ2C)pι2.b_{\iota,h}(I_{h}\boldsymbol{v}_{0},p)=b_{\iota}(\boldsymbol{v}_{0},p)-\iota^{2}(\nabla_{h}\operatorname{div}(\boldsymbol{v}_{0}-I_{h}\boldsymbol{v}_{0}),\nabla p)\geq(1-\gamma^{2}C_{\ast})\|p\|_{\iota}^{2}.

This gives

(4.10) supvVhbι,h(𝒗,p)𝒗ι,hbι,h(Ih𝒗0,p)Ihv0ι,h1γ2CCpι.\sup_{v\in V_{h}}\dfrac{b_{\iota,h}(\boldsymbol{v},p)}{\|\,\nabla\boldsymbol{v}\,\|_{\iota,h}}\geq\dfrac{b_{\iota,h}(I_{h}\boldsymbol{v}_{0},p)}{\|\,\nabla I_{h}v_{0}\,\|_{\iota,h}}\geq\dfrac{1-\gamma^{2}C_{*}}{C}\|p\|_{\iota}.

Next, if ι/h>γ\iota/h>\gamma, then we use (3.6) and obtain

(𝒗0Πh𝒗0)L2Ch2𝒗0L2,andh2(𝒗0Πh𝒗0)L2C2𝒗0L2.\|\,\nabla(\boldsymbol{v}_{0}-\varPi_{h}\boldsymbol{v}_{0})\,\|_{L^{2}}\leq Ch\|\,\nabla^{2}\boldsymbol{v}_{0}\,\|_{L^{2}},\quad\text{and}\quad\|\,\nabla^{2}_{h}(\boldsymbol{v}_{0}-\varPi_{h}\boldsymbol{v}_{0})\,\|_{L^{2}}\leq C\|\,\nabla^{2}\boldsymbol{v}_{0}\,\|_{L^{2}}.

Therefore,

(𝒗0Πh𝒗0)ι,hC(h+ι)2𝒗0L2C(1+h/ι)𝒗0ιC(1+1/γ)𝒗0ι.\|\,\nabla(\boldsymbol{v}_{0}-\varPi_{h}\boldsymbol{v}_{0})\,\|_{\iota,h}\leq C(h+\iota)\|\,\nabla^{2}\boldsymbol{v}_{0}\,\|_{L^{2}}\leq C(1+h/\iota)\|\nabla\boldsymbol{v}_{0}\|_{\iota}\leq C(1+1/\gamma)\|\nabla\boldsymbol{v}_{0}\|_{\iota}.

Hence,

Πh𝒗0ι,h𝒗0ι+(𝒗0Πh𝒗0)ι,hC(2+1/γ)𝒗0ι,\|\,\nabla\varPi_{h}\boldsymbol{v}_{0}\,\|_{\iota,h}\leq\|\nabla\boldsymbol{v}_{0}\|_{\iota}+\|\,\nabla(\boldsymbol{v}_{0}-\varPi_{h}\boldsymbol{v}_{0})\,\|_{\iota,h}\leq C(2+1/\gamma)\|\nabla\boldsymbol{v}_{0}\|_{\iota},

which together with (4.4) gives

(4.11) supvVhbι,h(𝒗,p)𝒗ι,hbι,h(Πh𝒗0,p)Πh𝒗0ι,h=bι,h(𝒗0,p)Πh𝒗0ι,h=pι2Πh𝒗0ι,hγC(1+2γ)pι.\sup_{v\in V_{h}}\dfrac{b_{\iota,h}(\boldsymbol{v},p)}{\|\,\nabla\boldsymbol{v}\,\|_{\iota,h}}\geq\dfrac{b_{\iota,h}(\varPi_{h}\boldsymbol{v}_{0},p)}{\|\,\nabla\varPi_{h}\boldsymbol{v}_{0}\,\|_{\iota,h}}=\dfrac{b_{\iota,h}(\boldsymbol{v}_{0},p)}{\|\,\nabla\varPi_{h}\boldsymbol{v}_{0}\,\|_{\iota,h}}=\dfrac{\|p\|_{\iota}^{2}}{\|\,\nabla\varPi_{h}\boldsymbol{v}_{0}\,\|_{\iota,h}}\geq\dfrac{\gamma}{C(1+2\gamma)}\|p\|_{\iota}.

A combination of (4.10) and (4.11) shows that (4.9) holds true with β\beta independent of ι\iota and hh. ∎

The well-posedness of the mixed approximation problem (4.2) follows from the ellipticity of aι,ha_{\iota,h} and the discrete B-B inequality of bι,hb_{\iota,h}. We are ready to derive the error estimate.

4.2. Error estimates

To carry out the error estimate, we define the bilinear form 𝒜\mathcal{A} as

𝒜(𝒗,q;𝒘,z):=aι,h(𝒗,𝒘)+bι,h(𝒘,q)+bι,h(𝒗,z)λ1cι(q,z)\mathcal{A}(\boldsymbol{v},q;\boldsymbol{w},z){:}=a_{\iota,h}(\boldsymbol{v},\boldsymbol{w})+b_{\iota,h}(\boldsymbol{w},q)+b_{\iota,h}(\boldsymbol{v},z)-\lambda^{-1}c_{\iota}(q,z)

for all 𝒗,𝒘Vh\boldsymbol{v},\boldsymbol{w}\in V_{h} and q,zPhq,z\in P_{h}.

We prove the following inf-sup inequality for 𝒜\mathcal{A} with the aid of the discrete B-B inequality (4.9).

Lemma 4.4.

There exists α\alpha depending on μ\mu and β\beta such that

(4.12) inf(𝒗,q)Vh×Phsup(𝒘,z)Vh×Ph𝒜(𝒗,q;𝒘,z)|(𝒘,z)|ι,h|(𝒗,q)|ι,hα,\inf_{(\boldsymbol{v},q)\in V_{h}\times P_{h}}\sup_{(\boldsymbol{w},z)\in V_{h}\times P_{h}}\dfrac{\mathcal{A}(\boldsymbol{v},q;\boldsymbol{w},z)}{|\!|\!|(\boldsymbol{w},z)|\!|\!|_{\iota,h}|\!|\!|(\boldsymbol{v},q)|\!|\!|_{\iota,h}}\geq\alpha,

where |(𝐰,z)|ι,h:=𝐰ι,h+zι+λ1/2zι|\!|\!|(\boldsymbol{w},z)|\!|\!|_{\iota,h}{:}=\|\,\nabla\boldsymbol{w}\,\|_{\iota,h}+\|z\|_{\iota}+\lambda^{-1/2}\|z\|_{\iota} and β\beta has appeared in (4.9).

Proof.

Noting that aι,ha_{\iota,h} is elliptic over VhV_{h} (4.3) and the discrete B-B inequality for bι,hb_{\iota,h} holds (4.9), we obtain (4.12) by [Braess:1996]*Theorem 2. ∎

We are ready to prove error estimates.

Theorem 4.5.

There exists CC independent of ι,λ\iota,\lambda and hh such that

(4.13) |(𝒖𝒖h,pph)|ι,hC(hr+ιhr1)(𝒖Hr+1+pHr),|\!|\!|(\boldsymbol{u}-\boldsymbol{u}_{h},p-p_{h})|\!|\!|_{\iota,h}\leq C(h^{r}+\iota h^{r-1})(\|\,\boldsymbol{u}\,\|_{H^{r+1}}+\|\,p\,\|_{H^{r}}),

and

(4.14) |(𝒖𝒖h,pph)|ι,hCh1/2𝒇L2.|\!|\!|(\boldsymbol{u}-\boldsymbol{u}_{h},p-p_{h})|\!|\!|_{\iota,h}\leq Ch^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}.
Proof.

Let 𝒗=𝒖h𝒖I\boldsymbol{v}=\boldsymbol{u}_{h}-\boldsymbol{u}_{I} and q=phpIq=p_{h}-p_{I} with 𝒖IVh\boldsymbol{u}_{I}\in V_{h} and pIPhp_{I}\in P_{h}, for any 𝒘Vh\boldsymbol{w}\in V_{h} and zPhz\in P_{h},

𝒜(𝒗,q;𝒘,z)\displaystyle\mathcal{A}(\boldsymbol{v},q;\boldsymbol{w},z) =𝒜(𝒖h,ph;𝒘,z)𝒜(𝒖,p;𝒘,z)+𝒜(𝒖𝒖I,ppI;𝒘,z)\displaystyle=\mathcal{A}(\boldsymbol{u}_{h},p_{h};\boldsymbol{w},z)-\mathcal{A}(\boldsymbol{u},p;\boldsymbol{w},z)+\mathcal{A}(\boldsymbol{u}-\boldsymbol{u}_{I},p-p_{I};\boldsymbol{w},z)
=(𝒇,𝒘)𝒜(𝒖,p;𝒘,z)+𝒜(𝒖𝒖I,ppI;𝒘,z)\displaystyle=(\boldsymbol{f},\boldsymbol{w})-\mathcal{A}(\boldsymbol{u},p;\boldsymbol{w},z)+\mathcal{A}(\boldsymbol{u}-\boldsymbol{u}_{I},p-p_{I};\boldsymbol{w},z)
=𝒜(𝒖𝒖I,ppI;𝒘,z)ι2ehe(𝒏𝝈𝒏)[[𝒏𝒘]]dσ(𝒙).\displaystyle=\mathcal{A}(\boldsymbol{u}-\boldsymbol{u}_{I},p-p_{I};\boldsymbol{w},z)-\iota^{2}\sum_{e\in\mathcal{E}_{h}}\int_{e}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot[\![\partial_{\boldsymbol{n}}\boldsymbol{w}]\!]\,\mathrm{d}\sigma(\boldsymbol{x}).

The boundedness of 𝒜\mathcal{A} yields

|𝒜(𝒖𝒖I,ppI;𝒘,z)|max(1,2μ)|(𝒖𝒖I,ppI)|ι,h|(𝒘,z)|ι,h.\lvert\mathcal{A}(\boldsymbol{u}-\boldsymbol{u}_{I},p-p_{I};\boldsymbol{w},z)\rvert\leq\max(1,2\mu)|\!|\!|(\boldsymbol{u}-\boldsymbol{u}_{I},p-p_{I})|\!|\!|_{\iota,h}|\!|\!|(\boldsymbol{w},z)|\!|\!|_{\iota,h}.

Let 𝒖I=Πh𝒖\boldsymbol{u}_{I}=\varPi_{h}\boldsymbol{u} be the interpolation of 𝒖\boldsymbol{u} and pIp_{I} be the r1r-1 order Lagrangian interpolation of pp, respectively. The standard interpolation error estimates in (3.6) gives

|(𝒖𝒖I,ppI)|ι,hC(hr+ιhr1)(𝒖Hr+1+pHr).|\!|\!|(\boldsymbol{u}-\boldsymbol{u}_{I},p-p_{I})|\!|\!|_{\iota,h}\leq C(h^{r}+\iota h^{r-1})(\|\,\boldsymbol{u}\,\|_{H^{r+1}}+\|\,p\,\|_{H^{r}}).

Note that

e[[𝒏𝒘]]qdσ(𝒙)=0for all qr2(e).\int_{e}[\![\partial_{\boldsymbol{n}}\boldsymbol{w}]\!]q\,\mathrm{d}\sigma(\boldsymbol{x})=0\qquad\text{for all }q\in\mathbb{P}_{r-2}(e).

A standard estimate for the consistency error functional with trace inequality (4.1) gives

ι2|ehe(𝒏𝝈𝒏)[[𝒏𝒘]]dσ(𝒙)|\displaystyle\iota^{2}\left\lvert\,\sum_{e\in\mathcal{E}_{h}}\int_{e}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot[\![\partial_{\boldsymbol{n}}\boldsymbol{w}]\!]\,\mathrm{d}\sigma(\boldsymbol{x})\,\right\rvert Cι2hr1(𝒖Hr+1+pHr)h2𝒘L2\displaystyle\leq C\iota^{2}h^{r-1}(\|\,\boldsymbol{u}\,\|_{H^{r+1}}+\|\,p\,\|_{H^{r}})\|\,\nabla_{h}^{2}\boldsymbol{w}\,\|_{L^{2}}
Cιhr1(𝒖Hr+1+pHr)𝒘ι,h.\displaystyle\leq C\iota h^{r-1}(\|\,\boldsymbol{u}\,\|_{H^{r+1}}+\|\,p\,\|_{H^{r}})\|\,\nabla\boldsymbol{w}\,\|_{\iota,h}.

A combination of the above three inequalities, the discrete inf-sup condition (4.12) and the triangle inequalities gives (4.13).

Next, let 𝒖I=Ih𝒖\boldsymbol{u}_{I}=I_{h}\boldsymbol{u} and let pIp_{I} be the Clément interpolation [Clement:1975] of pp, respectively. The interpolation error (4.7) and the error estimates for the Clément interpolation give

|(𝒖𝒖I,ppI)|ι,hCh1/2(𝒖H3/2+pH1/2+ι(𝒖H5/2+pH3/2))Ch1/2𝒇L2,|\!|\!|(\boldsymbol{u}-\boldsymbol{u}_{I},p-p_{I})|\!|\!|_{\iota,h}\leq Ch^{1/2}\Bigl{(}\|\,\boldsymbol{u}\,\|_{H^{3/2}}+\|\,p\,\|_{H^{1/2}}+\iota(\|\,\boldsymbol{u}\,\|_{H^{5/2}}+\|\,p\,\|_{H^{3/2}})\Bigr{)}\leq Ch^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}},

where we have used (2.29) and (2.30) in the last step.

Using the trace inequality (4.1), we bound the consistency error functional as

ι2|ehe(𝒏𝝈𝒏)[[𝒏𝒘]]dσ(𝒙)|\displaystyle\iota^{2}\left\lvert\,\sum_{e\in\mathcal{E}_{h}}\int_{e}(\partial_{\boldsymbol{n}}\boldsymbol{\sigma}\boldsymbol{n})\cdot[\![\partial_{\boldsymbol{n}}\boldsymbol{w}]\!]\,\mathrm{d}\sigma(\boldsymbol{x})\,\right\rvert Cι2h1/2(𝒖H2+pH1)1/2(𝒖H3+pH2)1/2h2𝒘L2\displaystyle\leq C\iota^{2}h^{1/2}(\|\,\boldsymbol{u}\,\|_{H^{2}}+\|\,p\,\|_{H^{1}})^{1/2}(\|\,\boldsymbol{u}\,\|_{H^{3}}+\|\,p\,\|_{H^{2}})^{1/2}\|\,\nabla_{h}^{2}\boldsymbol{w}\,\|_{L^{2}}
Ch1/2𝒇L2𝒘ι,h,\displaystyle\leq Ch^{1/2}\|\,\boldsymbol{f}\,\|_{L^{2}}\|\,\nabla\boldsymbol{w}\,\|_{\iota,h},

where we have used (2.28) in the last step.

Combining these inequalities, the discrete inf-sup condition (4.12) and the triangle inequalities gives (4.14). ∎

Corollary 4.6.

There exists CC independent of ι,λ\iota,\lambda and hh such that

(4.15) |(𝒖0𝒖h,p0ph)|ι,hC(ι1/2+h1/2)𝒇L2,|\!|\!|(\boldsymbol{u}_{0}-\boldsymbol{u}_{h},p_{0}-p_{h})|\!|\!|_{\iota,h}\leq C(\iota^{1/2}+h^{1/2})\|\,\boldsymbol{f}\,\|_{L^{2}},

where 𝐮0\boldsymbol{u}_{0} is the solution of (2.8), and p0=λdiv𝐮0p_{0}=\lambda\operatorname{div}\boldsymbol{u}_{0}.

Proof.

A combination of Theorem 2.9, Theorem 4.5, and the triangle inequality gives (4.15). ∎

5. Numerical Examples

In this part, we report the numerical performance for the proposed element of the lowest-order, i.e., r=2r=2. We test the accuracy and robustness of the element pair for the nearly incompressible materials. All examples are carried out on the nonuniform mesh. We are interested in the case when the Poisson’s ratio ν\nu is close to 0.50.5 and we report the relative errors (𝒖𝒖h)ι,h/𝒖ι\|\,\nabla(\boldsymbol{u}-\boldsymbol{u}_{h})\,\|_{\iota,h}/\|\nabla\boldsymbol{u}\|_{\iota} and the rates of convergence.

We let Ω=(0,1)2\Omega=(0,1)^{2}, and set Young’s modulus E=1E=1. The Lamé constants are determined by

λ=Eν(1+ν)(12ν),μ=E2(1+ν).\lambda=\dfrac{E\nu}{(1+\nu)(1-2\nu)},\quad\mu=\dfrac{E}{2(1+\nu)}.

We set ν=0.3\nu=0.3 for the ordinary cases, hence λ=0.5769\lambda=0.5769, and μ=0.3846\mu=0.3846, and we set ν=0.4999\nu=0.4999 for the nearly incompressible materials, hence λ=1.6664e3\lambda=\text{1.6664e3}, and μ=0.3334\mu=0.3334.

5.1. The first example

We test the performance of the element pair by solving a completely incompressible problem, which means div𝒖=0\operatorname{div}\boldsymbol{u}=0. Let 𝒖=(u1,u2)\boldsymbol{u}=(u_{1},u_{2}) with

u1=sin3(πx)sin(2πy)sin(πy),u2=sin(2πx)sin(πx)sin3(πy).u_{1}=-\sin^{3}(\pi x)\sin(2\pi y)\sin(\pi y),\quad u_{2}=\sin(2\pi x)\sin(\pi x)\sin^{3}(\pi y).

Therefore div𝒖=0\operatorname{div}\boldsymbol{u}=0, and 𝒇\boldsymbol{f} is independent of λ\lambda.

Table 2. Relative errors and convergence rates for the 1st example
ι\h\iota\backslash h 1/8 1/16 1/32 1/64
ν=0.3,λ=0.5769,μ=0.3846\nu=0.3,\lambda=0.5769,\mu=0.3846
1e+00 2.592e-01 1.333e-01 6.519e-02 3.246e-02
rate 0.96 1.03 1.01
1e-06 4.252e-02 1.159e-02 2.784e-03 6.918e-04
rate 1.88 2.06 2.01
ν=0.4999,λ=1.6664e3,μ=0.3334\nu=0.4999,\lambda=\text{1.6664e3},\mu=0.3334
1e+00 2.592e-01 1.333e-01 6.519e-02 3.246e-02
rate 0.96 1.03 1.01
1e-06 4.252e-02 1.159e-02 2.784e-03 6.918e-04
rate 1.88 2.06 2.01

In view of Table 2, the optimal rates of convergence are observed with the completely incompressible media, which is consistent with the error bound (4.13).

5.2. The second example

This example is motivated by [Wihler:2006], which admits a singular solution. The exact solution 𝒖=(u1,u2)\boldsymbol{u}=(u_{1},u_{2}) expressed in the polar coordinates as

u1=uρ(ρ,θ)cosθuθ(ρ,θ)sinθ,u2=uρ(ρ,θ)sinθ+uθ(ρ,θ)cosθ,u_{1}=u_{\rho}(\rho,\theta)\cos\theta-u_{\theta}(\rho,\theta)\sin\theta,\quad u_{2}=u_{\rho}(\rho,\theta)\sin\theta+u_{\theta}(\rho,\theta)\cos\theta,

where

uρ\displaystyle u_{\rho} =12μρα((α+1)cos((α+1)θ)+(C2(α+1))C1cos((α1)θ)),\displaystyle=\dfrac{1}{2\mu}\rho^{\alpha}\Bigl{(}-(\alpha+1)\cos((\alpha+1)\theta)+(C_{2}-(\alpha+1))C_{1}\cos((\alpha-1)\theta)\Bigr{)},
uθ\displaystyle u_{\theta} =12μρα((α+1)sin((α+1)θ)+(C2+α1)C1sin((α1)θ)),\displaystyle=\dfrac{1}{2\mu}\rho^{\alpha}\Bigl{(}(\alpha+1)\sin((\alpha+1)\theta)+(C_{2}+\alpha-1)C_{1}\sin((\alpha-1)\theta)\Bigr{)},

and α=1.5,ω=3π/4\alpha=1.5,\omega=3\pi/4,

C1=cos((α+1)ω)cos((α1)ω)andC2=2(λ+2μ)λ+μ.C_{1}=-\dfrac{\cos((\alpha+1)\omega)}{\cos((\alpha-1)\omega)}\quad\text{and}\quad C_{2}=\dfrac{2(\lambda+2\mu)}{\lambda+\mu}.

It may be verified that 𝒖[H5/2ε(Ω)]2\boldsymbol{u}\in[H^{5/2-\varepsilon}(\Omega)]^{2} for a small number ε>0\varepsilon>0. A direct calculation gives that 𝒇𝟎\boldsymbol{f}\equiv\boldsymbol{0}, while it is nearly incompressible because

div𝒖=3(1+2)λ+μρ1/2cos(θ/2).\operatorname{div}\boldsymbol{u}=-\dfrac{3(1+\sqrt{2})}{\lambda+\mu}\rho^{1/2}\cos(\theta/2).
Table 3. Relative errors and convergence rates for the 2nd example
ι\h\iota\backslash h 1/8 1/16 1/32 1/64
ν=0.3,λ=0.5769,μ=0.3846\nu=0.3,\lambda=0.5769,\mu=0.3846
1e+00 1.062e-01 7.554e-02 5.355e-02 3.792e-02
rate 0.49 0.50 0.50
1e-06 2.809e-03 1.001e-03 3.549e-04 1.257e-04
rate 1.49 1.50 1.50
ν=0.4999,λ=1.6664e3,μ=0.3334\nu=0.4999,\lambda=\text{1.6664e3},\mu=0.3334
1e+00 1.149e-01 8.200e-02 5.824e-02 4.135e-02
rate 0.49 0.49 0.49
1e-06 4.399e-03 1.567e-03 5.558e-04 1.968e-04
rate 1.49 1.50 1.50

It follows from Table 3 that the rates of convergence are sub-optimal. It is reasonable because the solution 𝒖\boldsymbol{u} is singular, which is similar to the results in [Wihler:2006]. The element pair is robust for the nearly incompressible materials.

5.3. The third example

In the last example, we test a problem with strong boundary layer effects. Such effects have been frequently observed in the stain elasticity model [Engel:2002, LiMingShi:2017, LiaoM:2019, LiMingWang:2021]. It is shown that the numerical solution converges to the solution of (2.8) when ιh\iota\ll h.

When ι0\iota\to 0, the boundary value problem (1.1) reduces to (2.8). Let 𝒖0=(u10,u20)\boldsymbol{u}_{0}=(u_{1}^{0},u_{2}^{0}) with

u10=sin2(πx)sin(2πy),u20=sin(2πx)sin2(πy)u_{1}^{0}=-\sin^{2}(\pi x)\sin(2\pi y),\quad u_{2}^{0}=\sin(2\pi x)\sin^{2}(\pi y)

be the solution of problem (2.8). The source term 𝒇\boldsymbol{f} is computed from (2.8). A direct calculation gives that div𝒖0=0\operatorname{div}\boldsymbol{u}_{0}=0, and 𝒇\boldsymbol{f} is independent of λ\lambda. The exact solution 𝒖\boldsymbol{u} for (1.1) is unknown, while it has strong boundary layer effects. In this case, we take ιh\iota\ll h, and report the relative error (𝒖0𝒖h)ι,h/𝒖0ι\|\,\nabla(\boldsymbol{u}_{0}-\boldsymbol{u}_{h})\,\|_{\iota,h}/\|\nabla\boldsymbol{u}_{0}\|_{\iota}.

Table 4. Relative errors and convergence rates for the 3rd example
ι\h\iota\backslash h 1/8 1/16 1/32 1/64
ν=0.3,λ=0.5769,μ=0.3846\nu=0.3,\lambda=0.5769,\mu=0.3846
1e-04 1.311e-01 8.966e-02 6.299e-02 4.476e-02
rate 0.55 0.51 0.49
1e-06 1.311e-01 8.960e-02 6.283e-02 4.432e-02
rate 0.55 0.51 0.50
ν=0.4999,λ=1.6664e3,μ=0.3334\nu=0.4999,\lambda=\text{1.6664e3},\mu=0.3334
1e-04 1.312e-01 8.968e-02 6.300e-02 4.476e-02
rate 0.55 0.51 0.49
1e-06 1.312e-01 8.963e-02 6.284e-02 4.432e-02
rate 0.55 0.51 0.50

It follows from Table 4 that the rate of convergence for the element pair changes to 1/21/2 because of the boundary layer effects, which is consistent with the theoretical result. The element is still robust when the solution has strong boundary layer effects in the nearly incompressible limit.

References