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

The inf-sup constant for hphp-Crouzeix-Raviart triangular elements

S. Sauter (stas@math.uzh.ch), Institut für Mathematik, Universität Zürich, Winterthurerstr 190, CH-8057 Zürich, Switzerland
Abstract

In this paper, we consider the discretization of the two-dimensional stationary Stokes equation by Crouzeix-Raviart elements for the velocity of polynomial order k1k\geq 1 on conforming triangulations and discontinuous pressure approximations of order k1k-1. We will bound the inf-sup constant from below independent of the mesh size and show that it depends only logarithmically on kk. Our assumptions on the mesh are very mild: for odd kk we require that the triangulations contain at least one inner vertex while for even kk we assume that the triangulations consist of more than a single triangle.

AMS Subject Classification: 65N30, 65N12, 76D07, 33C45,

Key Words:Non-conforming finite elements, Crouzeix-Raviart elements, high order finite elements, Stokes equation.

1 Introduction

In this paper we consider the numerical discretization of the two-dimensional stationary Stokes problem by Crouzeix-Raviart elements. They were introduced in the seminal paper [16] in 1973 by Crouzeix and Raviart with the goal to obtain a stable and economic discretization of the Stokes equation. They can be considered as an non-conforming enrichment of conforming finite elements of polynomial degree kk for the velocity and discontinuous pressures of degree k1k-1. It is well known that the conforming (k,k1)\left(k,k-1\right) pair of finite elements can be unstable; for two-dimensions the proof of the inf-sup stability of Crouzeix-Raviart discretizations of general order kk has been evolved over the last 50 years, the inf-sup stability for k=1k=1 has been proved in [16] and only recently the last open case k=3k=3, has been proved in [11]. We mention the papers [21], [32], [15], [7], [27], [10] which contain essential milestones in this development. There is a vast of literature on various further aspects of Crouzeix-Raviart elements; we omit to present a comprehensive review here but refer to the overview article [9] instead.

Since higher order methods are becoming increasingly popular a natural question arises how the inf-sup constant depends on the polynomial degree kk. It is the goal of this paper to investigate this dependency.

The paper is organized as follows. In Section 1.1 we introduce the Stokes problem and the Crouzeix-Raviart discretization of polynomial order kk. We state our main theorem that the discrete inf-sup constant can be estimated from below by c(log(k+1))αc\left(\log\left(k+1\right)\right)^{-\alpha}, where the positive constant cc depends only on the shape-regularity of the mesh and on the maximal outer angle of the domain Ω\Omega. The explicit value of α\alpha depends on the mesh topology. The simplest case is that each triangle in triangulation contains at least one inner vertex and then α=1/2\alpha=1/2 holds. We will give the value of α\alpha also for more general triangulations.

The proof is given in Section 2. The key ingredient is to show that for any discrete pressure qq, there exists a velocity field 𝐯q\mathbf{v}_{q} from the Crouzeix-Raviart space such that q~:=qdiv𝒯𝐯q\tilde{q}:=q-\operatorname{div}_{\mathcal{T}}\mathbf{v}_{q} belongs to the Scott-Vogelius pressure space [37, R.1, R.2], [32, R.1, R.2] (see (2.25)) and 𝐯q\mathbf{v}_{q} depends continuously on qq. These properties allow us to construct a conforming velocity field 𝐯~q~\mathbf{\tilde{v}}_{\tilde{q}} of order kk with div𝐯~q~=q~\operatorname*{div}\mathbf{\tilde{v}}_{\tilde{q}}=\tilde{q}. For this step, the construction in [32, Thm. 5.1], [27, Thm. 1] is modified such that the norm of the right-inverse does not deteriorate if a triangle vertex is a “nearly-critical” point – a notion which will be introduced in Definition 2.11. This key result is proved for odd polynomial degree in Section 2.2 and for even polynomial degree in Section 2.3.

In the conclusions (Sec. 3) we summarize our main findings and compare our results with existing results in the literature on some other pairs of finite elements for the Stokes equation.

In the appendices, we prove a technical result for a Gram matrix related to the bilinear form (div𝒯,)L2(Ω)\left(\operatorname*{div}_{\mathcal{T}}\cdot,\cdot\right)_{L^{2}\left(\Omega\right)} applied to the Crouzeix-Raviart element and discontinuous pressure space (see §A), an estimate of traces of non-conforming Crouzeix-Raviart basis functions (see §B), give some explicit formulae for integrals related to orthogonal polynomials (see §C), and prove a discrete Friedrichs inequality for Crouzeix-Raviart spaces (see §D).

1.1 The Stokes problem and its numerical discretization

Let Ω2\Omega\subset\mathbb{R}^{2} denote a bounded polygonal Lipschitz domain with boundary Ω\partial\Omega. For a vertex 𝐳\mathbf{z} in Ω\partial\Omega, we denote by α𝐳\alpha_{\mathbf{z}} the exterior angle between the two segments in Ω\partial\Omega with joint 𝐳\mathbf{z}. The minimal outer angle at the boundary vertices is given by

αΩ:=min𝐳 is a vertex in Ωα𝐳\alpha_{\Omega}:=\min_{\mathbf{z}\text{ is a vertex in }\partial\Omega}\alpha_{\mathbf{z}} (1.1)

and satisfies 0<αΩ<2π0<\alpha_{\Omega}<2\pi since Ω\Omega is Lipschitz. We consider the Stokes equation

Δ𝐮p=𝐟in Ω,div𝐮=0in Ω\begin{array}[c]{llll}-\Delta\mathbf{u}&-\nabla p&=\mathbf{f}&\text{in }\Omega,\\ \operatorname*{div}\mathbf{u}&&=0&\text{in }\Omega\end{array}

with Dirichlet boundary conditions for the velocity and a usual normalization condition for the pressure

𝐮=𝟎on Ωand Ωp=0.\mathbf{u}=\mathbf{0}\quad\text{on }\partial\Omega\quad\text{and\quad}\int_{\Omega}p=0.

To formulate this equation in a variational form we first introduce the relevant function spaces. Throughout the paper we restrict to vector spaces over the field of real numbers.

For s0s\geq 0, 1p1\leq p\leq\infty, Ws,p(Ω)W^{s,p}\left(\Omega\right) denote the classical Sobolev spaces of functions with norm Ws,p(Ω)\left\|\cdot\right\|_{W^{s,p}\left(\Omega\right)}. As usual we write Lp(Ω)L^{p}\left(\Omega\right) instead of W0,p(Ω)W^{0,p}\left(\Omega\right) and Hs(Ω)H^{s}\left(\Omega\right) for Ws,2(Ω)W^{s,2}\left(\Omega\right). For s0s\geq 0, we denote by H0s(Ω)H_{0}^{s}\left(\Omega\right) the closure of the space of infinitely smooth functions with compact support in Ω\Omega with respect to the Hs(Ω)H^{s}\left(\Omega\right) norm. Its dual space is denoted by Hs(Ω)H^{-s}\left(\Omega\right). For the pressure pp, the space L02(Ω):={uL2(Ω):Ωu=0}L_{0}^{2}\left(\Omega\right):=\left\{u\in L^{2}\left(\Omega\right):\int_{\Omega}u=0\right\} will be relevant.

The scalar product and norm in L2(Ω)L^{2}\left(\Omega\right) are denoted respectively by

(u,v)L2(Ω):=ΩuvanduL2(Ω):=(u,u)L2(Ω)1/2in L2(Ω).\begin{array}[c]{llll}\left(u,v\right)_{L^{2}\left(\Omega\right)}:=\int_{\Omega}uv&\text{and}&\left\|u\right\|_{L^{2}\left(\Omega\right)}:=\left(u,u\right)_{L^{2}\left(\Omega\right)}^{1/2}&\text{in }L^{2}\left(\Omega\right).\end{array}

Vector-valued and 2×22\times 2 tensor-valued analogues of the function spaces are denoted by bold and blackboard bold letters, e.g., 𝐇s(Ω)=(Hs(Ω))2\mathbf{H}^{s}\left(\Omega\right)=\left(H^{s}\left(\Omega\right)\right)^{2} and s=(Hs(Ω))2×2\mathbb{H}^{s}=\left(H^{s}\left(\Omega\right)\right)^{2\times 2}.

The 𝐋2(Ω)\mathbf{L}^{2}\left(\Omega\right) scalar product and norm for vector valued functions are given by

(𝐮,𝐯)𝐋2(Ω):=Ω𝐮,𝐯and 𝐮𝐋2(Ω):=(𝐮,𝐮)𝐋2(Ω)1/2,\left(\mathbf{u},\mathbf{v}\right)_{\mathbf{L}^{2}\left(\Omega\right)}:=\int_{\Omega}\left\langle\mathbf{u},\mathbf{v}\right\rangle\quad\text{and\quad}\left\|\mathbf{u}\right\|_{\mathbf{L}^{2}\left(\Omega\right)}:=\left(\mathbf{u},\mathbf{u}\right)_{\mathbf{L}^{2}\left(\Omega\right)}^{1/2},

where 𝐮,𝐯\left\langle\mathbf{u},\mathbf{v}\right\rangle denotes the Euclidean scalar product in 3\mathbb{R}^{3}. In a similar fashion, we define for 𝐆,𝐇𝕃2(Ω)\mathbf{G},\mathbf{H}\in\mathbb{L}^{2}\left(\Omega\right) the scalar product and norm by

(𝐆,𝐇)𝕃2(Ω):=Ω𝐆,𝐇and 𝐆𝕃2(Ω):=(𝐆,𝐆)𝕃2(Ω)1/2,\left(\mathbf{G},\mathbf{H}\right)_{\mathbb{L}^{2}\left(\Omega\right)}:=\int_{\Omega}\left\langle\mathbf{G},\mathbf{H}\right\rangle\quad\text{and\quad}\left\|\mathbf{G}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}:=\left(\mathbf{G},\mathbf{G}\right)_{\mathbb{L}^{2}\left(\Omega\right)}^{1/2},

where 𝐆,𝐇=i,j=13Gi,jHi,j\left\langle\mathbf{G},\mathbf{H}\right\rangle=\sum_{i,j=1}^{3}G_{i,j}H_{i,j}. We also need fractional order Sobolev norms on boundary of triangles and introduce the relevant notation; for details see, e.g., [29]. For a bounded Lipschitz domain ω2\omega\subset\mathbb{R}^{2} with boundary ω\partial\omega, let L2(ω)L^{2}\left(\partial\omega\right) and H1(ω)H^{1}\left(\partial\omega\right) denote the usual Lebesgue and Sobolev space on ω\partial\omega with norm L2(ω)\left\|\cdot\right\|_{L^{2}\left(\partial\omega\right)} and H1(ω)\left\|\cdot\right\|_{H^{1}\left(\partial\omega\right)}. For 0<s<10<s<1 the fractional Sobolev space on ω\partial\omega of order ss is denoted by Hs(ω)H^{s}\left(\partial\omega\right) and equipped with the norm

vHs(ω):=(vL2(ω)2+|v|Hs(ω)2)1/2\left\|v\right\|_{H^{s}\left(\partial\omega\right)}:=\left(\left\|v\right\|_{L^{2}\left(\partial\omega\right)}^{2}+\left|v\right|_{H^{s}\left(\partial\omega\right)}^{2}\right)^{1/2}

and seminorm

|v|Hs(ω):=(ωω|v(𝐱)v(𝐲)|2𝐱𝐲1+2s𝑑𝐲𝑑𝐱)1/2.\left|v\right|_{H^{s}\left(\partial\omega\right)}:=\left(\int_{\partial\omega}\int_{\partial\omega}\frac{\left|v\left(\mathbf{x}\right)-v\left(\mathbf{y}\right)\right|^{2}}{\left\|\mathbf{x}-\mathbf{y}\right\|^{1+2s}}d\mathbf{y}d\mathbf{x}\right)^{1/2}.

We introduce the bilinear form a:𝐇1(Ω)×𝐇1(Ω)a:\mathbf{H}^{1}\left(\Omega\right)\times\mathbf{H}^{1}\left(\Omega\right)\rightarrow\mathbb{R} by

a(𝐮,𝐯):=(𝐮,𝐯)𝕃2(Ω),a\left(\mathbf{u},\mathbf{v}\right):=\left(\nabla\mathbf{u},\nabla\mathbf{v}\right)_{\mathbb{L}^{2}\left(\Omega\right)}, (1.2)

where 𝐮\nabla\mathbf{u} and 𝐯\nabla\mathbf{v} denote the derivatives of 𝐮\mathbf{u} and 𝐯\mathbf{v}. The variational form of the Stokes problem is given by: For given 𝐅𝐇1(Ω),\mathbf{F}\in\mathbf{H}^{-1}\left(\Omega\right),

find (𝐮,p)𝐇01(Ω)×L02(Ω)s.t. {a(𝐮,𝐯)+(p,div𝐯)L2(Ω)=𝐅(𝐯)𝐯𝐇01(Ω),(div𝐮,q)L2(Ω)=0qL02(Ω).\text{find }\left(\mathbf{u},p\right)\in\mathbf{H}_{0}^{1}\left(\Omega\right)\times L_{0}^{2}\left(\Omega\right)\;\text{s.t.\ }\left\{\begin{array}[c]{lll}a\left(\mathbf{u},\mathbf{v}\right)+\left(p,\operatorname*{div}\mathbf{v}\right)_{L^{2}\left(\Omega\right)}&=\mathbf{F}\left(\mathbf{v}\right)&\forall\mathbf{v}\in\mathbf{H}_{0}^{1}\left(\Omega\right),\\ \left(\operatorname*{div}\mathbf{u},q\right)_{L^{2}\left(\Omega\right)}&=0&\forall q\in L_{0}^{2}\left(\Omega\right).\end{array}\right. (1.3)

It is well-known (see, e.g., [22]) that (1.3) is well posed. Since we consider non-conforming discretizations we restrict the space 𝐇1(Ω)\mathbf{H}^{-1}\left(\Omega\right) for the right-hand side to a smaller space and assume from now on for simplicity that 𝐅(𝐯)=(𝐟,𝐯)𝐋2(Ω)\mathbf{F}\left(\mathbf{v}\right)=\left(\mathbf{f},\mathbf{v}\right)_{\mathbf{L}^{2}\left(\Omega\right)} for some 𝐟𝐋2(Ω)\mathbf{f}\in\mathbf{L}^{2}\left(\Omega\right); for a more general setting we refer to [35], [36].

In the following a discretization for problem (1.3) is introduced. Let 𝒯={Ki:1in}\mathcal{T}=\left\{K_{i}:1\leq i\leq n\right\} denote a triangulation of Ω\Omega consisting of closed triangles which are conforming: the intersection of two different triangles is either empty, a common edge, or a common point. We also assume Ω=dom𝒯\Omega=\operatorname*{dom}\mathcal{T}, where

dom𝒯:=int(K𝒯K)\operatorname*{dom}\mathcal{T}:=\operatorname*{int}\left({\displaystyle\bigcup\limits_{K\in\mathcal{T}}}K\right) (1.4)

and int(M):=M\operatorname*{int}\left(M\right):=\overset{\circ}{M} denotes the interior of a set M2M\subset\mathbb{R}^{2}.

Piecewise versions of differential operators such as \nabla and div\operatorname*{div} are defined for functions uu and vector fields 𝐰\mathbf{w} which are sufficiently smooth in the interior of the triangles K𝒯K\subset\mathcal{T} by

(𝒯u)|K:=(u|K)and (div𝒯𝐰)|K:=div(𝐰|K).\left.\left(\nabla_{\mathcal{T}}u\right)\right|_{\overset{\circ}{K}}:=\nabla\left(\left.u\right|_{\overset{\circ}{K}}\right)\quad\text{and\quad}\left.\left(\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{w}\right)\right|_{\overset{\circ}{K}}:=\operatorname*{div}\left(\left.\mathbf{w}\right|_{\overset{\circ}{K}}\right).

The values on K\partial K are arbitrary since K\partial K has measure zero.

An important measure for the quality of a finite element triangulation is the shape-regularity constant given by

γ𝒯:=maxK𝒯hKρK\gamma_{\mathcal{T}}:=\max_{K\in\mathcal{T}}\frac{h_{K}}{\rho_{K}} (1.5)

with the local mesh width hK:=diamKh_{K}:=\operatorname*{diam}K and ρK\rho_{K} denoting the diameter of the largest inscribed ball in KK. The global mesh width is h𝒯:=max{hK:K𝒯}h_{\mathcal{T}}:=\max\left\{h_{K}:K\in\mathcal{T}\right\}.

Remark 1.1

It is well known that the shape-regularity implies that there exists some minimal angle ϕ𝒯>0\phi_{\mathcal{T}}>0 depending only on γ𝒯\gamma_{\mathcal{T}} such that every triangle angle in 𝒯\mathcal{T} is bounded from below by ϕ𝒯\phi_{\mathcal{T}}. In turn, every triangle angle in 𝒯\mathcal{T} is bounded from above by π2ϕ𝒯\pi-2\phi_{\mathcal{T}}.

The set of edges in 𝒯\mathcal{T} is denoted by (𝒯)\mathcal{E}\left(\mathcal{T}\right), while the subset of boundary edges is Ω(𝒯):={E(𝒯):E(dom𝒯)}\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right):=\left\{E\in\mathcal{E}\left(\mathcal{T}\right):E\subset\partial\left(\operatorname*{dom}\mathcal{T}\right)\right\}; the subset of inner edges is given by Ω(𝒯):=(𝒯)\Ω(𝒯)\mathcal{E}_{\Omega}\left(\mathcal{T}\right):=\mathcal{E}\left(\mathcal{T}\right)\backslash\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right). For each edge EE\in\mathcal{E} we fix a unit vector 𝐧E\mathbf{n}_{E} orthogonal to EE with the convention that 𝐧E\mathbf{n}_{E} is the outer normal vector for boundary edges EΩE\in\mathcal{E}_{\partial\Omega}.

The set of triangle vertices in 𝒯\mathcal{T} is denoted by 𝒱(𝒯)\mathcal{V}\left(\mathcal{T}\right), while the subset of inner vertices is 𝒱Ω(𝒯):={𝐕𝒱(𝒯):𝐕(dom𝒯)}\mathcal{V}_{\Omega}\left(\mathcal{T}\right):=\left\{\mathbf{V}\in\mathcal{V}\left(\mathcal{T}\right):\mathbf{V}\notin\partial\left(\operatorname*{dom}\mathcal{T}\right)\right\} and 𝒱Ω(𝒯):=𝒱(𝒯)\𝒱Ω(𝒯)\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right):=\mathcal{V}\left(\mathcal{T}\right)\backslash\mathcal{V}_{\Omega}\left(\mathcal{T}\right). For K𝒯K\in\mathcal{T}, the set of its vertices is denoted by 𝒱(K)\mathcal{V}\left(K\right). For E(𝒯)E\in\mathcal{E}\left(\mathcal{T}\right), we define the edge patch by

𝒯E:={K𝒯:EK}and ωE:=K𝒯EK.\mathcal{T}_{E}:=\left\{K\in\mathcal{T}:E\subset K\right\}\quad\text{and\quad}\omega_{E}:={\displaystyle\bigcup\limits_{K\in\mathcal{T}_{E}}}K.

For 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right), the nodal patch is defined by

𝒯𝐳:={K𝒯:𝐳K}and ω𝐳:=K𝒯𝐳K\mathcal{T}_{\mathbf{z}}:=\left\{K\in\mathcal{T}:\mathbf{z}\in K\right\}\quad\text{and\quad}\omega_{\mathbf{z}}:={\displaystyle\bigcup\limits_{K\in\mathcal{T}_{\mathbf{z}}}}K (1.6)

with local mesh width h𝐳:=max{hK:K𝒯𝐳}h_{\mathbf{z}}:=\max\left\{h_{K}:K\in\mathcal{T}_{\mathbf{z}}\right\}. For K𝒯K\in\mathcal{T}, we set

𝒯K:={K𝒯KK}and ωK:=K𝒯KK.\mathcal{T}_{K}:=\left\{K^{\prime}\in\mathcal{T}\mid K\cap K^{\prime}\neq\emptyset\right\}\quad\text{and\quad}\omega_{K}:={\displaystyle\bigcup\limits_{K^{\prime}\in\mathcal{T}_{K}}}K^{\prime}. (1.7)

For a subset M2M\subset\mathbb{R}^{2}, we denote by [M]\left[M\right] its convex hull; in this way an edge EE with endpoints 𝐚,𝐛\mathbf{a},\mathbf{b} can be written as E=[𝐚,𝐛]=[𝐛,𝐚]E=\left[\mathbf{a},\mathbf{b}\right]=\left[\mathbf{b},\mathbf{a}\right].

Let ={1,2,}\mathbb{N}=\left\{1,2,\ldots\right\} and 0:={0}\mathbb{N}_{0}:=\mathbb{N\cup}\left\{0\right\}. For mm\in\mathbb{N}, we employ the usual multiindex notation for 𝝁=(μi)i=1m0m\mbox{\boldmath$\mu$}=\left(\mu_{i}\right)_{i=1}^{m}\in\mathbb{N}_{0}^{m} and points 𝐱=(xi)i=1mm\mathbf{x}=\left(x_{i}\right)_{i=1}^{m}\in\mathbb{R}^{m}

|𝝁|:=μ1++μm,𝐱𝝁:=j=1mxjμj.\left|\mbox{\boldmath$\mu$}\right|:=\mu_{1}+\ldots+\mu_{m},\quad\mathbf{x}^{\mbox{\boldmath$\mu$}}:={\displaystyle\prod\limits_{j=1}^{m}}x_{j}^{\mu_{j}}.

Let m,k\mathbb{P}_{m,k} denote the space of mm-variate polynomials of maximal degree kk, consisting of functions of the form

𝝁0m|𝝁|ka𝝁𝐱𝝁\sum_{\begin{subarray}{c}\mbox{\boldmath$\mu$}\in\mathbb{N}_{0}^{m}\\ \left|\mbox{\boldmath$\mu$}\right|\leq k\end{subarray}}a_{\mbox{\boldmath$\mu$}}\mathbf{x}^{\mbox{\boldmath$\mu$}}

for real coefficients a𝝁a_{\mbox{\boldmath$\mu$}}. Formally, we set m,1:={0}\mathbb{P}_{m,-1}:=\left\{0\right\}. To indicate the domain explicitly in notation we write sometimes k(D)\mathbb{P}_{k}\left(D\right) for DmD\subset\mathbb{R}^{m} and skip the index mm since it is then clear from the argument DD.

We introduce the following finite element spaces

k(𝒯):={qL2(Ω)K𝒯:q|Kk(K)},and (cf. (1.4))k,0(𝒯):={qk(𝒯):dom𝒯q=0}.\begin{array}[c]{ll}&\mathbb{P}_{k}\left(\mathcal{T}\right):=\left\{q\in L^{2}\left(\Omega\right)\mid\forall K\in\mathcal{T}:\left.q\right|_{\overset{\circ}{K}}\in\mathbb{P}_{k}\left(\overset{\circ}{K}\right)\right\},\\ \text{and (cf. (\ref{defdomT}))}&\mathbb{P}_{k,0}\left(\mathcal{T}\right):=\left\{q\in\mathbb{P}_{k}\left(\mathcal{T}\right):\int_{\operatorname*{dom}\mathcal{T}}q=0\right\}.\end{array} (1.8)

Furthermore, let

Sk(𝒯):=k(𝒯)H1(dom𝒯),andSk,0(𝒯):=Sk(𝒯)H01(dom𝒯).\begin{array}[c]{ll}&S_{k}\left(\mathcal{T}\right):=\mathbb{P}_{k}\left(\mathcal{T}\right)\cap H^{1}\left(\operatorname*{dom}\mathcal{T}\right),\\ \text{and}&S_{k,0}\left(\mathcal{T}\right):=S_{k}\left(\mathcal{T}\right)\cap H_{0}^{1}\left(\operatorname*{dom}\mathcal{T}\right).\end{array}

The vector-valued versions are denoted by 𝐒k(𝒯):=Sk(𝒯)2\mathbf{S}_{k}\left(\mathcal{T}\right):=S_{k}\left(\mathcal{T}\right)^{2} and 𝐒k,0(𝒯):=Sk,0(𝒯)2\mathbf{S}_{k,0}\left(\mathcal{T}\right):=S_{k,0}\left(\mathcal{T}\right)^{2}. Finally, we define the Crouzeix-Raviart space by

CRk(𝒯)\displaystyle\operatorname*{CR}\nolimits_{k}\left(\mathcal{T}\right) :={vk(𝒯)qk1(E)EΩ(𝒯)E[v]Eq=0},\displaystyle:=\left\{v\in\mathbb{P}_{k}\left(\mathcal{T}\right)\mid\forall q\in\mathbb{P}_{k-1}\left(E\right)\quad\forall E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)\quad\int_{E}\left[v\right]_{E}q=0\right\}, (1.9a)
CRk,0(𝒯)\displaystyle\operatorname*{CR}\nolimits_{k,0}\left(\mathcal{T}\right) :={vCRk(𝒯)qk1(E)EΩ(𝒯)Evq=0}.\displaystyle:=\left\{v\in\operatorname*{CR}\nolimits_{k}\left(\mathcal{T}\right)\mid\forall q\in\mathbb{P}_{k-1}\left(E\right)\quad\forall E\in\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right)\quad\int_{E}vq=0\right\}. (1.9b)

Here, [v]E\left[v\right]_{E} denotes the jump of vk(𝒯)v\in\mathbb{P}_{k}\left(\mathcal{T}\right) across an edge EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)

[u]E(𝐱):=limε0(u(𝐱+ε𝐧E)u(𝐱ε𝐧E)).\left[u\right]_{E}\left(\mathbf{x}\right):=\lim_{\varepsilon\searrow 0}\left(u\left(\mathbf{x}+\varepsilon\mathbf{n}_{E}\right)-u\left(\mathbf{x}-\varepsilon\mathbf{n}_{E}\right)\right).

and k1(E)\mathbb{P}_{k-1}\left(E\right) is the space of polynomials of maximal degree k1k-1 with respect to the local variable in EE.

We have collected all ingredients for defining the Crouzeix-Raviart discretization for the Stokes equation. For kk\in\mathbb{N}, let the discrete velocity space and pressure space be defined by

𝐂𝐑k,0(𝒯):=(CRk,0(𝒯))2and Mk1(𝒯):=k1,0(𝒯).\mathbf{CR}_{k,0}\left(\mathcal{T}\right):=\left(\operatorname*{CR}\nolimits_{k,0}\left(\mathcal{T}\right)\right)^{2}\quad\text{and\quad}M_{k-1}\left(\mathcal{T}\right):=\mathbb{P}_{k-1,0}\left(\mathcal{T}\right).

Then, the discretization is given by: find (𝐮CR,pdisc)𝐂𝐑k,0(𝒯)×Mk1(𝒯)\left(\mathbf{u}_{\operatorname*{CR}},p_{\operatorname*{disc}}\right)\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right)\times M_{k-1}\left(\mathcal{T}\right)\;such that

{a𝒯(𝐮CR,𝐯)b𝒯(𝐯,pdisc)=(𝐟,𝐯)𝐋2(Ω)𝐯𝐂𝐑k,0(𝒯),b𝒯(𝐮CR,q)=0qMk1(𝒯),\left\{\begin{array}[c]{lll}a_{\mathcal{T}}\left(\mathbf{u}_{\operatorname*{CR}},\mathbf{v}\right)-b_{\mathcal{T}}\left(\mathbf{v},p_{\operatorname*{disc}}\right)&=\left(\mathbf{f},\mathbf{v}\right)_{\mathbf{L}^{2}\left(\Omega\right)}&\forall\mathbf{v}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right),\\ b_{\mathcal{T}}\left(\mathbf{u}_{\operatorname*{CR}},q\right)&=0&\forall q\in M_{k-1}\left(\mathcal{T}\right),\end{array}\right. (1.10)

where the bilinear forms a𝒯:𝐂𝐑k,0(𝒯)×𝐂𝐑k,0(𝒯)a_{\mathcal{T}}:\mathbf{CR}_{k,0}\left(\mathcal{T}\right)\times\mathbf{CR}_{k,0}\left(\mathcal{T}\right)\rightarrow\mathbb{R} and b𝒯:𝐂𝐑k,0(𝒯)×Mk1(𝒯)b_{\mathcal{T}}:\mathbf{CR}_{k,0}\left(\mathcal{T}\right)\times M_{k-1}\left(\mathcal{T}\right)\rightarrow\mathbb{R} are given by

a𝒯(𝐮,𝐯):=(𝒯𝐮,𝒯𝐯)𝕃2(Ω)and b𝒯(𝐯,q):=(div𝒯𝐯,q)L2(Ω).a_{\mathcal{T}}\left(\mathbf{u},\mathbf{v}\right):=\left(\nabla_{\mathcal{T}}\mathbf{u},\nabla_{\mathcal{T}}\mathbf{v}\right)_{\mathbb{L}^{2}\left(\Omega\right)}\quad\text{and\quad}b_{\mathcal{T}}\left(\mathbf{v},q\right):=\left(\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{v},q\right)_{L^{2}\left(\Omega\right)}.

It is well known that problem (1.10) is well-posed if (i): the bilinear form a𝒯(,)a_{\mathcal{T}}\left(\cdot,\cdot\right) is coercive and (ii): b𝒯(,)b_{\mathcal{T}}\left(\cdot,\cdot\right) satisfies the inf-sup condition.

To verify the condition (i) we introduce, for a conforming triangulation 𝒯\mathcal{T} of the domain Ω\Omega, the broken Sobolev space

H1(𝒯):={uL2(Ω)K𝒯:u|KH1(K)}H^{1}\left(\mathcal{T}\right):=\left\{u\in L^{2}\left(\Omega\right)\mid\forall K\in\mathcal{T}:\left.u\right|_{K}\in H^{1}\left(K\right)\right\}

and define, for uH1(𝒯)u\in H^{1}\left(\mathcal{T}\right), the broken H1H^{1}–seminorm by

uH1(𝒯):=𝒯u𝐋2(Ω)=(K𝒯u𝐋2(K)2)1/2.\left\|u\right\|_{H^{1}\left(\mathcal{T}\right)}:=\left\|\nabla_{\mathcal{T}}u\right\|_{\mathbf{L}^{2}\left(\Omega\right)}=\left(\sum_{K\in\mathcal{T}}\left\|\nabla u\right\|_{\mathbf{L}^{2}\left(K\right)}^{2}\right)^{1/2}.

In [16, Lem. 2]) it is proved that 𝐇1(𝒯)\left\|\cdot\right\|_{\mathbf{H}^{1}\left(\mathcal{T}\right)} defines a norm in 𝐂𝐑k,0(𝒯)+𝐇01(Ω)\mathbf{CR}_{k,0}\left(\mathcal{T}\right)+\mathbf{H}_{0}^{1}\left(\Omega\right) which is equivalent to the norm (K𝒯𝐮𝐇1(K)2)1/2\left(\sum_{K\in\mathcal{T}}\left\|\mathbf{u}\right\|_{\mathbf{H}^{1}\left(K\right)}^{2}\right)^{1/2} with equivalence constants independent of the polynomial degree and the mesh width (see Theorem D.1). This directly implies the coercivity of a𝒯(,)a_{\mathcal{T}}\left(\cdot,\cdot\right):

a𝒯(𝐮,𝐮)𝐮𝐇1(𝒯)2𝐮𝐂𝐑k,0(𝒯).a_{\mathcal{T}}\left(\mathbf{u},\mathbf{u}\right)\geq\left\|\mathbf{u}\right\|_{\mathbf{H}^{1}\left(\mathcal{T}\right)}^{2}\quad\forall\mathbf{u}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right).

Hence, well-posedness of (1.10) follows from the inf-sup condition for b𝒯(,)b_{\mathcal{T}}\left(\cdot,\cdot\right).

Definition 1.2

Let 𝒯\mathcal{T} denote a conforming triangulation for Ω\Omega. The pair 𝐂𝐑k,0(𝒯)×Mk1(𝒯)\mathbf{CR}_{k,0}\left(\mathcal{T}\right)\times M_{k-1}\left(\mathcal{T}\right) is inf-sup stable if there exists a constant c𝒯,kc_{\mathcal{T},k} such that

infpMk1(𝒯)\{0}sup𝐯𝐂𝐑k,0(𝒯)\{𝟎}(p,div𝒯𝐯)L2(Ω)𝐯𝐇1(𝒯)pL2(Ω)c𝒯,k>0.\inf_{p\in M_{k-1}\left(\mathcal{T}\right)\backslash\left\{0\right\}}\sup_{\mathbf{v}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right)\backslash\left\{\mathbf{0}\right\}}\frac{\left(p,\operatorname*{div}_{\mathcal{T}}\mathbf{v}\right)_{L^{2}\left(\Omega\right)}}{\left\|\mathbf{v}\right\|_{\mathbf{H}^{1}\left(\mathcal{T}\right)}\left\|p\right\|_{L^{2}\left(\Omega\right)}}\geq c_{\mathcal{T},k}>0. (1.11)

We are now in the position to formulate our main theorem.

Theorem 1.3

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded polygonal Lipschitz domain and let 𝒯\mathcal{T} denote a conforming triangulation of Ω\Omega consisting of more than a single triangle. Let kk\in\mathbb{N}. If k3k\geq 3 is odd we assume that 𝒯\mathcal{T} contains at least one inner vertex. Then, the inf-sup condition (1.11) holds:

c𝒯,kc𝒯(log(k+1))αc_{\mathcal{T},k}\geq c_{\mathcal{T}}\left(\log\left(k+1\right)\right)^{-\alpha} (1.12)

for a constant c𝒯>0c_{\mathcal{T}}>0 depending only on the shape-regularity of the mesh and on the maximal outer angle αΩ\alpha_{\Omega}. In particular c𝒯c_{\mathcal{T}} is independent of the mesh width h𝒯h_{\mathcal{T}} and the polynomial degree kk. The value of α0\alpha\geq 0 is given by

α={1/2{if k is even,or k3 is odd and all triangles in 𝒯 have at least one inner vertex,(1+L)/2otherwise,\alpha=\left\{\begin{array}[c]{ll}1/2&\left\{\begin{array}[c]{l}\text{if }k\text{ is even,}\\ \text{or }k\geq 3\text{ is odd and all triangles in }\mathcal{T}\text{ have at least one inner vertex,}\end{array}\right.\\ \left(1+L\right)/2&\text{otherwise,}\end{array}\right.

where LL depends only on the mesh topology via the number of steps involved in the step-by-step construction introduced in (2.13).

Proof. The estimate c𝒯,k>0c_{\mathcal{T},k}>0 follows, for k=1k=1 from [16], for k=2k=2 from [12, Thm. 3.1], for even k4k\geq 4 from [7], for odd k5k\geq 5 from [10], and for k=3k=3 from [11]. We set

c𝒯,low:=min{c𝒯,k:1k3}.c_{\mathcal{T},\operatorname*{low}}:=\min\left\{c_{\mathcal{T},k}:1\leq k\leq 3\right\}.

Estimate (1.12) for some c𝒯:=c𝒯,highodd>0c_{\mathcal{T}}:=c_{\mathcal{T},\operatorname*{high}}^{\operatorname*{odd}}>0 for odd k5k\geq 5 is proved in Section 2.2, Lem. 2.22, while the estimate for some c𝒯:=c𝒯,higheven>0c_{\mathcal{T}}:=c_{\mathcal{T},\operatorname*{high}}^{\operatorname*{even}}>0 for even k4k\geq 4 is proved in Section 2.3. Both constants c𝒯,highoddc_{\mathcal{T},\operatorname*{high}}^{\operatorname*{odd}}, c𝒯,highevenc_{\mathcal{T},\operatorname*{high}}^{\operatorname*{even}} depend only on the shape-regularity of the mesh and αΩ\alpha_{\Omega}. Hence, c𝒯min{c𝒯,low,c𝒯,highodd,c𝒯,higheven}c_{\mathcal{T}}\geq\min\left\{c_{\mathcal{T},\operatorname*{low}},c_{\mathcal{T},\operatorname*{high}}^{\operatorname*{odd}},c_{\mathcal{T},\operatorname*{high}}^{\operatorname*{even}}\right\} 

We emphasize that the original definition in [16] allows for slightly more general finite element spaces, more precisely, the spaces CRk(𝒯)\operatorname*{CR}_{k}\left(\mathcal{T}\right) can be enriched by locally supported functions. From this point of view, the definition (1.9) describes a minimal Crouzeix-Raviart space.

The possibility for enrichment has been used frequently in the literature to prove inf-sup stability for the arising finite element spaces (see, e.g., [16], [26], [28]). In contrast, we will prove the kk-explicit estimate of the inf-sup constant for the Crouzeix-Raviart space CRk(𝒯)\operatorname*{CR}_{k}\left(\mathcal{T}\right).

2 Proof of Theorem 1.3

In this section, we will analyse the kk-dependence of the inf-sup constant in the form (1.12), first for odd polynomial degree k5k\geq 5 and then for even degree k4k\geq 4.

2.1 Barycentric coordinates and basis functions for the velocity

In this section, we introduce basis functions for the finite element spaces in Section 1.1. We begin with introducing some general notation.

Notation 2.1

For vectors 𝐚in\mathbf{a}_{i}\in\mathbb{R}^{n}, 1im1\leq i\leq m, we write [𝐚1𝐚2𝐚m]\left[\mathbf{a}_{1}\mid\mathbf{a}_{2}\mid\ldots\mid\mathbf{a}_{m}\right] for the n×mn\times m matrix with column vectors 𝐚i\mathbf{a}_{i}. For 𝐯=(v1,v2)T2\mathbf{v}=\left(v_{1},v_{2}\right)^{T}\in\mathbb{R}^{2} we set 𝐯:=(v2,v1)T\mathbf{v}^{\perp}:=\left(v_{2},-v_{1}\right)^{T}. Let 𝐞k,ik\mathbf{e}_{k,i}\in\mathbb{R}^{k} be the ii-th canonical unit vector in k\mathbb{R}^{k}.

For 𝐯n\mathbf{v}\in\mathbb{R}^{n}, 𝐯\left\|\mathbf{v}\right\| is the Euclidean vector norm while the induced matrix norm is given for 𝐁n×n\mathbf{B}\in\mathbb{R}^{n\times n} by 𝐁:=sup{𝐁𝐱/𝐱:𝐱n\{𝟎}}\left\|\mathbf{B}\right\|:=\sup\left\{\left\|\mathbf{Bx}\right\|/\left\|\mathbf{x}\right\|:\mathbf{x}\in\mathbb{R}^{n}\backslash\left\{\mathbf{0}\right\}\right\}.

Vertices in a triangle are numbered counterclockwise. In a triangle KK with vertices 𝐀1\mathbf{A}_{1}, 𝐀2\mathbf{A}_{2}, 𝐀3\mathbf{A}_{3} the angle at 𝐀i\mathbf{A}_{i} is called αi\alpha_{i}. If a triangle is numbered by an index (e.g., KK_{\ell}), the angle at A,iA_{\ell,i} is called α,i\alpha_{\ell,i}. For quantities in a triangle KK as, e.g., angles αj\alpha_{j}, 1j31\leq j\leq 3, we use the cyclic numbering convention α3+1:=α1\alpha_{3+1}:=\alpha_{1} and α11:=α3\alpha_{1-1}:=\alpha_{3}.

For a dd-dimensional measurable set D,D, we write |D|\left|D\right| for its measure; for a discrete set, say 𝒥\mathcal{J}, we denote by |𝒥|\left|\mathcal{J}\right| its cardinality.

In the proofs, we consider frequently nodal patches 𝒯𝐳\mathcal{T}_{\mathbf{z}} for inner vertices 𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\Omega}\left(\mathcal{T}\right). The number mm denotes the number of triangles in 𝒯𝐳\mathcal{T}_{\mathbf{z}}. Various quantities in this patch such as, e.g., the triangles in 𝒯𝐳\mathcal{T}_{\mathbf{z}}, have an index which runs from 11 to mm. Here, we use the cyclic numbering convention Km+1:=K1K_{m+1}:=K_{1} and K11:=KmK_{1-1}:=K_{m} and apply this analogously for other quantities in the nodal patch.

Let the closed reference triangle K^\widehat{K} be the triangle with vertices 𝐀^1:=(0,0)T\mathbf{\hat{A}}_{1}:=\left(0,0\right)^{T}, 𝐀^2:=(1,0)T\mathbf{\hat{A}}_{2}:=\left(1,0\right)^{T}, 𝐀^3:=(0,1)T\mathbf{\hat{A}}_{3}:=\left(0,1\right)^{T}. The nodal points on the reference element of order k0k\in\mathbb{N}_{0} are given by

𝒩^k:={{1k𝝁𝝁02:|𝝁|k}k1,{(13,13)}k=0.\widehat{\mathcal{N}}_{k}:=\left\{\begin{array}[c]{ll}\left\{\dfrac{1}{k}\mbox{\boldmath$\mu$}\mid\mbox{\boldmath$\mu$}\in\mathbb{N}_{0}^{2}:\mathbb{\quad}\left|\mbox{\boldmath$\mu$}\right|\leq k\right\}&k\geq 1,\\ &\\ \left\{\left(\dfrac{1}{3},\dfrac{1}{3}\right)\right\}&k=0.\end{array}\right.

For a triangle K2K\subset\mathbb{R}^{2}, we denote by χK:K^K\chi_{K}:\widehat{K}\rightarrow K an affine bijection. The mapped nodal points of order k0k\in\mathbb{N}_{0} on KK are given by

𝒩k(K):={χK(𝐳):𝐳𝒩^k}.\mathcal{N}_{k}\left(K\right):=\left\{\chi_{K}\left(\mathbf{z}\right):\mathbf{z}\in\widehat{\mathcal{N}}_{k}\right\}.

Nodal points of order kk on 𝒯\mathcal{T} are defined by

𝒩k(𝒯):=K𝒯𝒩k(K)𝒩Ωk(𝒯):=𝒩k(𝒯)Ω,and 𝒩k,Ω(𝒯):=𝒩k(𝒯)Ω.\mathcal{N}_{k}\left(\mathcal{T}\right):={\displaystyle\bigcup\limits_{K\in\mathcal{T}}}\mathcal{N}_{k}\left(K\right)\text{,\quad}\mathcal{N}_{\partial\Omega}^{k}\left(\mathcal{T}\right):=\mathcal{N}_{k}\left(\mathcal{T}\right)\cap\partial\Omega,\quad\text{and\quad}\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right):=\mathcal{N}_{k}\left(\mathcal{T}\right)\cap\Omega.

We introduce the Lagrange basis for the space Sk(𝒯)S_{k}\left(\mathcal{T}\right), which is indexed by the nodal points 𝐳𝒩k(𝒯)\mathbf{z}\in\mathcal{N}_{k}\left(\mathcal{T}\right) and characterized by

Bk,𝐳Sk(𝒯)and 𝐳𝒩k(𝒯)Bk,𝐳(𝐳)=δ𝐳,𝐳,B_{k,\mathbf{z}}\in S_{k}\left(\mathcal{T}\right)\quad\text{and\quad}\forall\mathbf{z}^{\prime}\in\mathcal{N}_{k}\left(\mathcal{T}\right)\qquad B_{k,\mathbf{z}}\left(\mathbf{z}^{\prime}\right)=\delta_{\mathbf{z},\mathbf{z}^{\prime}}, (2.1)

where δ𝐳,𝐳\delta_{\mathbf{z},\mathbf{z}^{\prime}} is the Kronecker delta. A basis for the space Sk,0(𝒯)S_{k,0}\left(\mathcal{T}\right) is given by Bk,𝐳B_{k,\mathbf{z}}, 𝐳𝒩k,Ω(𝒯)\mathbf{z}\in\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right).

Let KK denote a triangle with vertices 𝐀i\mathbf{A}_{i}, 1i31\leq i\leq 3, and let λK,𝐀i1(K)\lambda_{K,\mathbf{A}_{i}}\in\mathbb{P}_{1}\left(K\right) be the barycentric coordinate for the node 𝐀i\mathbf{A}_{i} defined by

λK,𝐀i(𝐀j)=δi,j1i,j3.\lambda_{K,\mathbf{A}_{i}}\left(\mathbf{A}_{j}\right)=\delta_{i,j}\quad 1\leq i,j\leq 3. (2.2)

If the numbering of the vertices in KK is fixed, we write λK,i\lambda_{K,i} short for λK,𝐀i\lambda_{K,\mathbf{A}_{i}}. For the barycentric coordinate on the reference element K^\widehat{K} for the vertex 𝐀^j\mathbf{\hat{A}}_{j} we write λ^j\widehat{\lambda}_{j}, j=1,2,3j=1,2,3. Elementary calculation yield (see, e.g., [10, Appendix A])

𝐧kλK,𝐀i=|Ei|2|K|×{1i=k,cosα s.t. {,i,k}={1,2,3},\partial_{\mathbf{n}_{k}}\lambda_{K,\mathbf{A}_{i}}=\frac{\left|E_{i}\right|}{2\left|K\right|}\times\left\{\begin{array}[c]{ll}-1&i=k,\\ \cos\alpha_{\ell}&\ell\text{ s.t. }\left\{\ell,i,k\right\}=\left\{1,2,3\right\},\end{array}\right. (2.3)

where EiE_{i} is the edge of KK opposite to 𝐀i\mathbf{A}_{i}, 𝐧k\mathbf{n}_{k} the outward unit normal at EkE_{k}, and α\alpha_{\ell} the angle in KK at 𝐀\mathbf{A}_{\ell}.

Definition 2.2

Let LkL_{k} denote the usual univariate Legendre polynomial of degree kk (see [17, Table 18.3.1]). Let kk\in\mathbb{N} be even and K𝒯K\in\mathcal{T}. Then, the non-conforming triangle bubble is given by

Bk,KCR:={12(1+i=13Lk(12λK,i))on K,0on Ω\K.B_{k,K}^{\operatorname*{CR}}:=\left\{\begin{array}[c]{ll}\dfrac{1}{2}\left(-1+{\displaystyle\sum\limits_{i=1}^{3}}L_{k}\left(1-2\lambda_{K,i}\right)\right)&\text{on }K,\\ 0&\text{on }\Omega\backslash K.\end{array}\right.

For kk odd and E(𝒯)E\in\mathcal{E}\left(\mathcal{T}\right), the non-conforming edge bubble is given by

Bk,ECR:={Lk(12λK,𝐀K,E)on K for K𝒯E,0on Ω\ωE,B_{k,E}^{\operatorname*{CR}}:=\left\{\begin{array}[c]{ll}L_{k}\left(1-2\lambda_{K,\mathbf{A}_{K,E}}\right)&\text{on }K\text{ for }K\in\mathcal{T}_{E},\\ 0&\text{on }\Omega\backslash\omega_{E},\end{array}\right. (2.4)

where 𝐀K,E\mathbf{A}_{K,E} denotes the vertex in KK opposite to EE.

Different representations of the functions Bk,ECRB_{k,E}^{\operatorname*{CR}}, Bk,KCRB_{k,K}^{\operatorname*{CR}} exist in the literature, see [34], [5], [12, for p=4,6.p=4,6.], [13] while the formula for Bk,KCRB_{k,K}^{\operatorname*{CR}} has been introduced in [7] and the one for Bk,ECRB_{k,E}^{\operatorname*{CR}} in [10].

Proposition 2.3

A basis for the space CRk,0(𝒯)\operatorname*{CR}_{k,0}\left(\mathcal{T}\right) is given

  1. 1.

    for even kk by

    {Bk,𝐳𝐳𝒩k,Ω(𝒯)}{Bk,KCRK𝒯},\left\{B_{k,\mathbf{z}}\mid\mathbf{z}\in\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right)\right\}\cup\left\{B_{k,K}^{\operatorname*{CR}}\mid K\in\mathcal{T}\right\},
  2. 2.

    for odd kk by

    {Bk,𝐳𝐳𝒩k,Ω(𝒯)\𝒱Ω(𝒯)}{Bk,ECREΩ(𝒯)}.\left\{B_{k,\mathbf{z}}\mid\mathbf{z}\in\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right)\backslash\mathcal{V}_{\Omega}\left(\mathcal{T}\right)\right\}\cup\left\{B_{k,E}^{\operatorname*{CR}}\mid E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)\right\}.

The proof of this proposition and the following corollary can be found, e.g., in [34, Rem. 3], [13, Thm. 22], [10, Cor. 3.4].

Corollary 2.4

A basis for the space 𝐂𝐑k,0(𝒯)\mathbf{CR}_{k,0}\left(\mathcal{T}\right) is given

  1. 1.

    for even kk by

    {Bk,𝐳𝐯𝐳𝐳𝒩k,Ω(𝒯)}{Bk,𝐳𝐰𝐳𝐳𝒩k,Ω(𝒯)}{Bk,KCR𝐯KK𝒯}{Bk,KCR𝐰KK𝒯},\begin{array}[c]{l}\left\{B_{k,\mathbf{z}}\mathbf{v}_{\mathbf{z}}\mid\mathbf{z}\in\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right)\right\}\cup\left\{B_{k,\mathbf{z}}\mathbf{w}_{\mathbf{z}}\mid\mathbf{z}\in\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right)\right\}\\ \\ \quad\cup\left\{B_{k,K}^{\operatorname*{CR}}\mathbf{v}_{K}\mid K\in\mathcal{T}\right\}\cup\left\{B_{k,K}^{\operatorname*{CR}}\mathbf{w}_{K}\mid K\in\mathcal{T}\right\},\end{array} (2.5)
  2. 2.

    for odd kk by

    {Bk,𝐳𝐯𝐳𝐳𝒩k,Ω(𝒯)\𝒱Ω(𝒯)}{Bk,𝐳𝐰𝐳𝐳𝒩k,Ω(𝒯)\𝒱Ω(𝒯)}{Bk,ECR𝐯EEΩ(𝒯)}{Bk,ECR𝐰EEΩ(𝒯)}.\begin{array}[c]{l}\left\{B_{k,\mathbf{z}}\mathbf{v}_{\mathbf{z}}\mid\mathbf{z}\in\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right)\backslash\mathcal{V}_{\Omega}\left(\mathcal{T}\right)\right\}\cup\left\{B_{k,\mathbf{z}}\mathbf{w}_{\mathbf{z}}\mid\mathbf{z}\in\mathcal{N}_{k,\Omega}\left(\mathcal{T}\right)\backslash\mathcal{V}_{\Omega}\left(\mathcal{T}\right)\right\}\\ \\ \quad\cup\left\{B_{k,E}^{\operatorname*{CR}}\mathbf{v}_{E}\mid E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)\right\}\cup\left\{B_{k,E}^{\operatorname*{CR}}\mathbf{w}_{E}\mid E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)\right\}.\end{array} (2.6)

Here, for any nodal point 𝐳\mathbf{z}, the linearly independent vectors 𝐯𝐳,𝐰𝐳2\mathbf{v}_{\mathbf{z}},\mathbf{w}_{\mathbf{z}}\in\mathbb{R}^{2} can be chosen arbitrarily. The same holds for any triangle KK for the vectors 𝐯K,𝐰K2\mathbf{v}_{K},\mathbf{w}_{K}\in\mathbb{R}^{2} in (2.5) and for any EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right) for the vectors 𝐯E,𝐰E2\mathbf{v}_{E},\mathbf{w}_{E}\in\mathbb{R}^{2} in (2.6).

Remark 2.5

The original definition of Crouzeix-Raviart spaces by [16] is implicit and given for conforming simplicial finite element meshes in d\mathbb{R}^{d}, d=2,3d=2,3. For their practical implementation, a basis is needed and Corollary 2.4 provides a simple definition. A basis for Crouzeix-Raviart finite elements in 3\mathbb{R}^{3} is introduced in [20] for k=2k=2, a general construction is given in [14], and a basis for a minimal Crouzeix-Raviart spaces in general dimension dd is presented in [30].

2.2 The case of odd k5k\geq 5

In this section, we assume for the following

a)k5 is odd andb)𝒯 is a conforming triangulation and has at least one inner vertex.\begin{array}[c]{ll}\text{a)}&k\geq 5\text{ is odd and}\\ \text{b)}&\mathcal{T}\text{ is a conforming triangulation and has at least one inner vertex.}\end{array} (2.7)

This section is structured as follows. In §2.2.1 we generalize the concept of critical points (see [37], [32]) to η\eta-critical points which turn out to be essential for estimates with constants depending on the mesh only via the shape-regularity constant and αΩ\alpha_{\Omega}. We split these η\eta-critical points into a set of “obtuse” η\eta-critical points and “acute” η\eta-critical points. In §2.2.2, we provide the proof of Theorem 1.3 for a maximal partial triangulation that does not contain acute η\eta-critical points and satisfies (2.7). Finally, in §2.2.3 we present the argument to allow for acute η\eta-critical points.

2.2.1 Geometric preliminaries

For the analysis of the inf-sup constant we start with the definition of critical points (see [37], [32]).

Definition 2.6

Let 𝒯\mathcal{T} denote a triangulation as in §1.1. For 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right), let

𝐳:={E(𝒯):𝐳 is an endpoint of E}.\mathcal{E}_{\mathbf{z}}:=\left\{E\in\mathcal{E}\left(\mathcal{T}\right):\mathbf{z}\text{ is an endpoint of }E\right\}.

The point 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right) is a critical point for 𝒯\mathcal{T} if there exist two straight infinite lines L1L_{1}, L2L_{2} in 2\mathbb{R}^{2} such that all edges E𝐳E\in\mathcal{E}_{\mathbf{z}} satisfy EL1L2E\subset L_{1}\cup L_{2}. The set of all critical points in 𝒯\mathcal{T} is 𝒞𝒯\mathcal{C}_{\mathcal{T}}.

Refer to caption
Figure 1: Illustration of the four critical cases as in Remark 2.7. Left top: inner critical point, right top: acute critical point, left bottom: flat critical point, right bottom: obtuse critical point.
Remark 2.7

Geometric configurations where critical points occur are well studied in the literature (see, e.g., [32]). Any critical point 𝐳𝒞𝒯\mathbf{z}\in\mathcal{C}_{\mathcal{T}} belongs to one of the following cases (see Fig. 1):

  1. 1.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\Omega}\left(\mathcal{T}\right) and 𝒯𝐳\mathcal{T}_{\mathbf{z}} consists of four triangles and 𝐳\mathbf{z} is the intersections of the two diagonals in the quadrilateral ω𝐳\omega_{\mathbf{z}}.

  2. 2.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and card𝐳=2\operatorname*{card}\mathcal{E}_{\mathbf{z}}=2, i.e., both edges E𝐳E\in\mathcal{E}_{\mathbf{z}} are boundary edges with joint 𝐳\mathbf{z}.

  3. 3.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and card𝐳=3\operatorname*{card}\mathcal{E}_{\mathbf{z}}=3 and two edges E𝐳E\in\mathcal{E}_{\mathbf{z}} are boundary edges which lie on a straight boundary piece.

  4. 4.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and card𝐳=4\operatorname*{card}\mathcal{E}_{\mathbf{z}}=4 and each of the two boundary edges is aligned with one edge of 𝐳Ω(𝒯)\mathcal{E}_{\mathbf{z}}\cap\mathcal{E}_{\Omega}\left(\mathcal{T}\right).

Definition 2.8

Let 𝒯\mathcal{T} denote a triangulation as in §1.1. Let 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right) and the nodal patch 𝒯𝐳\mathcal{T}_{\mathbf{z}} as in (1.6). Let the triangles KK_{\ell}, 1m1\leq\ell\leq m, in 𝒯𝐳\mathcal{T}_{\mathbf{z}} be numbered counterclockwise and denote the angle in KK_{\ell} at 𝐳\mathbf{z} by ω\omega_{\ell}. Then,

Θ(𝐳):={max{|sin(ω1+ω2)|,|sin(ω2+ω3)|,,|sin(ωm+ω1)|}if 𝐳𝒱Ω(𝒯),max{|sin(ω1+ω2)|,|sin(ω2+ω3)|,,|sin(ωm1+ωm)|}if 𝐳Γm>1,0if 𝐳Γm=1.\Theta\left(\mathbf{z}\right):=\left\{\begin{array}[c]{ll}\max\left\{\left|\sin\left(\omega_{1}+\omega_{2}\right)\right|,\left|\sin\left(\omega_{2}+\omega_{3}\right)\right|,\ldots,\left|\sin\left(\omega_{m}+\omega_{1}\right)\right|\right\}&\text{if }\mathbf{z}\in\mathcal{V}_{\Omega}\left(\mathcal{T}\right),\\ \max\left\{\left|\sin\left(\omega_{1}+\omega_{2}\right)\right|,\left|\sin\left(\omega_{2}+\omega_{3}\right)\right|,\ldots,\left|\sin\left(\omega_{m-1}+\omega_{m}\right)\right|\right\}&\text{if }\mathbf{z}\in\Gamma\wedge m>1,\\ 0&\text{if }\mathbf{z\in}\Gamma\wedge m=1.\end{array}\right.
Remark 2.9

It is easy to see that 𝐳𝒞𝒯\mathbf{z}\in\mathcal{C}_{\mathcal{T}} if and only if Θ(𝐳)=0\Theta\left(\mathbf{z}\right)=0.

Lemma 2.10

Let ϕ𝒯\phi_{\mathcal{T}} be as in Remark 1.1. Set

η0:=min{12,c1,3ϕ𝒯π,sinϕ𝒯}\eta_{0}:=\min\left\{\frac{1}{2},c_{1},\frac{3\phi_{\mathcal{T}}}{\pi},\sin\phi_{\mathcal{T}}\right\}

with

c1:={min{sin2ϕ𝒯,|sin(2π4ϕ𝒯)|}ϕ𝒯π/8,sin2ϕ𝒯π/8<ϕ𝒯π/4,1ϕ𝒯>π/4.c_{1}:=\left\{\begin{array}[c]{ll}\min\left\{\sin 2\phi_{\mathcal{T}},\left|\sin\left(2\pi-4\phi_{\mathcal{T}}\right)\right|\right\}&\phi_{\mathcal{T}}\leq\pi/8,\\ \sin 2\phi_{\mathcal{T}}&\pi/8<\phi_{\mathcal{T}}\leq\pi/4,\\ 1&\phi_{\mathcal{T}}>\pi/4.\end{array}\right.

Let 0η<η00\leq\eta<\eta_{0} be fixed. If, for 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right), it holds Θ(𝐳)η\Theta\left(\mathbf{z}\right)\leq\eta, then, for any edge E=[𝐳,𝐳]Ω(𝒯)E=\left[\mathbf{z},\mathbf{z}^{\prime}\right]\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right) it holds

Θ(𝐳)η0.\Theta\left(\mathbf{z}^{\prime}\right)\geq\eta_{0}.

Proof. Let 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right) and consider an edge E=[𝐳,𝐳]Ω(𝒯)E=\left[\mathbf{z},\mathbf{z}^{\prime}\right]\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right). Then, there are two triangles K,K𝒯K,K^{\prime}\in\mathcal{T} which are adjacent to EE. The angle in KK resp. KK^{\prime} at 𝐳\mathbf{z} is denoted by ω\omega resp. ω\omega^{\prime}.

1st case. Let ω+ωπ/2\omega+\omega^{\prime}\leq\pi/2 or ω+ω32π\omega+\omega^{\prime}\geq\frac{3}{2}\pi. Then, we conclude from Remark 1.1 that

2ϕ𝒯ω+ωπ2or 32πω+ω2π4ϕ𝒯.2\phi_{\mathcal{T}}\leq\omega+\omega^{\prime}\leq\frac{\pi}{2}\quad\text{or\quad}\frac{3}{2}\pi\leq\omega+\omega^{\prime}\leq 2\pi-4\phi_{\mathcal{T}}.

For the left inequality to hold, the minimal angle must satisfy ϕ𝒯π/4\phi_{\mathcal{T}}\leq\pi/4 while for the right inequality, it must hold ϕ𝒯π/8\phi_{\mathcal{T}}\leq\pi/8. For ϕ𝒯>π/4\phi_{\mathcal{T}}>\pi/4, the 1st case is empty. For ϕ𝒯π/4\phi_{\mathcal{T}}\leq\pi/4 we get

Θ(𝐳){min{sin2ϕ𝒯,|sin(2π4ϕ𝒯)|}ϕ𝒯π/8sin2ϕ𝒯π/8<ϕ𝒯π/4}c1η0.\Theta\left(\mathbf{z}\right)\geq\left\{\begin{array}[c]{ll}\min\left\{\sin 2\phi_{\mathcal{T}},\left|\sin\left(2\pi-4\phi_{\mathcal{T}}\right)\right|\right\}&\phi_{\mathcal{T}}\leq\pi/8\\ \sin 2\phi_{\mathcal{T}}&\pi/8<\phi_{\mathcal{T}}\leq\pi/4\end{array}\right\}\geq c_{1}\geq\eta_{0}.

Since η<η0c1\eta<\eta_{0}\leq c_{1} this case cannot appear.

2nd case. Let π/2<ω+ω<3π/2\pi/2<\omega+\omega^{\prime}<3\pi/2. The condition |sin(ω+ω)|η\left|\sin\left(\omega+\omega^{\prime}\right)\right|\leq\eta implies that ω+ω=π+δ\omega+\omega^{\prime}=\pi+\delta with

|δ|arcsinη=[17, 4.24.1]η=0η2(2)!(!)24(2+1)η01/2η=022(2)!(!)24(2+1)=πη3.\left|\delta\right|\leq\arcsin\eta\overset{\text{\cite[cite]{[\@@bibref{}{NIST:DLMF}{}{}, 4.24.1]}}}{=}\eta\sum_{\ell=0}^{\infty}\eta^{2\ell}\frac{\left(2\ell\right)!}{\left(\ell!\right)^{2}4^{\ell}\left(2\ell+1\right)}\overset{\eta_{0}\leq 1/2}{\leq}\eta\sum_{\ell=0}^{\infty}2^{-2\ell}\frac{\left(2\ell\right)!}{\left(\ell!\right)^{2}4^{\ell}\left(2\ell+1\right)}=\frac{\pi\eta}{3}. (2.8)

Consequently the two angles α\alpha in KK and α\alpha^{\prime} in KK^{\prime} at 𝐳\mathbf{z}^{\prime} satisfy

α+α=2πωωββ=πδββπ+πη32ϕ𝒯πη/3ϕ𝒯πϕ𝒯,\alpha+\alpha^{\prime}=2\pi-\omega-\omega^{\prime}-\beta-\beta^{\prime}=\pi-\delta-\beta-\beta^{\prime}\leq\pi+\frac{\pi\eta}{3}-2\phi_{\mathcal{T}}\overset{\pi\eta/3\leq\phi_{\mathcal{T}}}{\leq}\pi-\phi_{\mathcal{T}},

where β\beta (resp. β\beta^{\prime}) denotes the third angle in KK (resp. KK^{\prime}). Hence, in this case

Θ(𝐳)|sin(πϕ𝒯)|=sinϕ𝒯η0.\Theta\left(\mathbf{z}^{\prime}\right)\geq\left|\sin\left(\pi-\phi_{\mathcal{T}}\right)\right|=\sin\phi_{\mathcal{T}}\geq\eta_{0}.

 

Definition 2.11

Let η0\eta_{0} be as in Lemma 2.10. For 0η<η00\leq\eta<\eta_{0}, the set of η\eta-critical points 𝒞𝒯(η)\mathcal{C}_{\mathcal{T}}\left(\eta\right) is given by

𝒞𝒯(η):={𝐳𝒱(𝒯)Θ(𝐳)η}.\mathcal{C}_{\mathcal{T}}\left(\eta\right):=\left\{\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right)\mid\Theta\left(\mathbf{z}\right)\leq\eta\right\}.

A point 𝐳𝒞𝒯(η)\𝒞𝒯(0)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right)\backslash\mathcal{C}_{\mathcal{T}}\left(0\right) is called a nearly critical point. An η\eta-critical point 𝐳𝒞𝒯(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right) is isolated if all edge [𝐳,𝐳](𝒯)\left[\mathbf{z,z}^{\prime}\right]\in\mathcal{E}\left(\mathcal{T}\right) satisfy: 𝐳\mathbf{z}^{\prime} is not an η\eta-critical point.

By perturbing the geometric configurations in Remark 2.7 we obtain the following subcases (see Fig. 2).

Definition 2.12

Let η0\eta_{0} be as in Lemma 2.10 and 0η<η00\leq\eta<\eta_{0}. If 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right) satisfies

  1. 1.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\Omega}\left(\mathcal{T}\right) and card𝒯𝐳=4\operatorname*{card}\mathcal{T}_{\mathbf{z}}=4 and Θ(𝐳)η\Theta\left(\mathbf{z}\right)\leq\eta. Then 𝐳\mathbf{z} is an inner η\eta-critical point. Let

    𝒞𝒯inner(η):={𝐳𝒞𝒯(η):𝐳 is an inner η-critical point}.\mathcal{C}_{\mathcal{T}}^{\operatorname*{inner}}\left(\eta\right):=\left\{\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right):\mathbf{z}\text{ is an inner }\eta\text{-critical point}\right\}.
  2. 2.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and card𝐳=2\operatorname*{card}\mathcal{E}_{\mathbf{z}}=2. Then 𝐳\mathbf{z} is an acute critical point. Let111Note that the set of acute critical points is independent of η\eta.

    𝒞𝒯acute:={𝐳𝒞𝒯:𝐳 is an acute critical point}.\mathcal{C}_{\mathcal{T}}^{\operatorname{acute}}:=\left\{\mathbf{z}\in\mathcal{C}_{\mathcal{T}}:\mathbf{z}\text{ is an acute critical point}\right\}.
  3. 3.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and card𝐳=3\operatorname*{card}\mathcal{E}_{\mathbf{z}}=3 and Θ(𝐳)η\Theta\left(\mathbf{z}\right)\leq\eta. Then 𝐳\mathbf{z} is flat η\eta-critical point. Let

    𝒞𝒯flat(η):={𝐳𝒞𝒯(η):𝐳 is a flat η-critical point}.\mathcal{C}_{\mathcal{T}}^{\operatorname*{flat}}\left(\eta\right):=\left\{\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right):\mathbf{z}\text{ is a flat }\eta\text{-critical point}\right\}.
  4. 4.

    𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and card𝐳=4\operatorname*{card}\mathcal{E}_{\mathbf{z}}=4 and Θ(𝐳)η\Theta\left(\mathbf{z}\right)\leq\eta. Then 𝐳\mathbf{z} is a (locally) concave η\eta-critical point. Let

    𝒞𝒯concave(η):={𝐳𝒞𝒯(η):𝐳 is a concave η-critical point}.\mathcal{C}_{\mathcal{T}}^{\operatorname*{concave}}\left(\eta\right):=\left\{\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right):\mathbf{z}\text{ is a concave }\eta\text{-critical point}\right\}.
Refer to caption
Figure 2: Illustration of the four η\eta-critical cases as in Remark 2.12. Left top: inner η\eta-critical point, right top: acute critical point, left bottom: flat η\eta-critical point, right bottom: concave η\eta-critical point.

The acute critical points require some special treatment and we denote the union of the others by

𝒞𝒯obtuse(η):=𝒞𝒯inner(η)𝒞𝒯flat(η)𝒞𝒯concave(η).\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right):=\mathcal{C}_{\mathcal{T}}^{\operatorname*{inner}}\left(\eta\right)\cup\mathcal{C}_{\mathcal{T}}^{\operatorname*{flat}}\left(\eta\right)\cup\mathcal{C}_{\mathcal{T}}^{\operatorname*{concave}}\left(\eta\right).

The following lemma states that for a possibly adjusted η0\eta_{0}, still depending only on the shape-regularity of the mesh and the maximal outer angle αΩ\alpha_{\Omega}, the η\eta-critical points belong to one of the four categories described in Definition 2.12.

Lemma 2.13

Let 𝒯\mathcal{T} be a conforming triangulation such that D:=dom𝒯D:=\operatorname*{dom}\mathcal{T} is a Lipschitz domain.

Then, there exists some η0]0,η0]\eta_{0}^{\prime}\in\left]0,\eta_{0}\right] depending only on the shape-regularity of the mesh and the minimal outer angle αD\alpha_{D} such that for 0η<η00\leq\eta<\eta_{0}^{\prime} any η\eta-critical point belongs to one of the four categories described in Definition 2.12.

Proof. Let 𝐳\mathbf{z} be an η\eta-critical point. We set m:=card𝒯𝐳m:=\operatorname*{card}\mathcal{T}_{\mathbf{z}} and choose a counterclockwise numbering for the triangles in 𝒯𝐳\mathcal{T}_{\mathbf{z}}, i.e., KiK_{i}, 1im1\leq i\leq m. The shape-regularity of the mesh implies that there is some mmaxm_{\max} depending only on ϕ𝒯\phi_{\mathcal{T}} such that mmmaxm\leq m_{\max}. Denote by ωi\omega_{i} the angle in KiK_{i} at 𝐳\mathbf{z}. Let η0]0,η0]\eta_{0}^{\prime}\in\left]0,\eta_{0}\right] which will be fixed later and assume 0η<η00\leq\eta<\eta_{0}^{\prime}. Since 𝐳\mathbf{z} is an η\eta-critical it holds

|sin(ωi+ωi+1)|η<η01imfor m:={mif 𝐳𝒱Ω(𝒯),m1if 𝐳𝒱Ω(𝒯).\left|\sin\left(\omega_{i}+\omega_{i+1}\right)\right|\leq\eta<\eta_{0}^{\prime}\quad\forall 1\leq i\leq m^{\prime}\quad\text{for }m^{\prime}:=\left\{\begin{array}[c]{ll}m&\text{if }\mathbf{z}\in\mathcal{V}_{\Omega}\left(\mathcal{T}\right),\\ m-1&\text{if }\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right).\end{array}\right.

The shape-regularity implies ϕ𝒯ωiπ2ϕ𝒯\phi_{\mathcal{T}}\leq\omega_{i}\leq\pi-2\phi_{\mathcal{T}} and, for δ=arcsinη0\delta=\arcsin\eta_{0}^{\prime}, we get

ωi+ωi+1[2ϕ𝒯,δ][πδ,π+δ][2πδ,2π4ϕ𝒯]for all 1im.\omega_{i}+\omega_{i+1}\in\left[2\phi_{\mathcal{T}},\delta\right]\cup\left[\pi-\delta,\pi+\delta\right]\cup\left[2\pi-\delta,2\pi-4\phi_{\mathcal{T}}\right]\quad\text{for all }1\leq i\leq m^{\prime}. (2.9)

Since arcsin:[0,1[0\arcsin:\left[0,1\right[\rightarrow\mathbb{R}_{\geq 0} is monotonously increasing with arcsin0=0\arcsin 0=0 and limx1arcsinx=+\lim_{x\rightarrow 1}\arcsin x=+\infty, we can select η0\eta_{0}^{\prime} such 0<δ<2ϕ𝒯0<\delta<2\phi_{\mathcal{T}}. In turn, the first and last interval in (2.9) are empty and

ωi+ωi+1=:π+δifor some δi with |δi|δ for all 1im.\omega_{i}+\omega_{i+1}=:\pi+\delta_{i}\quad\text{for some }\delta_{i}\text{ with }\left|\delta_{i}\right|\leq\delta\text{ for all }1\leq i\leq m^{\prime}. (2.10)

Case 1: 𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\Omega}\left(\mathcal{T}\right).

In this case we obtain

4π=i=1m(ωi+ωi+1)=(2.10)mπ+i=1mδi.4\pi=\sum_{i=1}^{m}\left(\omega_{i}+\omega_{i+1}\right)\overset{\text{(\ref{defdeltai})}}{=}m\pi+\sum_{i=1}^{m}\delta_{i}. (2.11)

By adjusting η0\eta_{0}^{\prime} such that mmaxδ<πm_{\max}\delta<\pi we conclude that m=4m=4 and i=1mδi=0\sum_{i=1}^{m}\delta_{i}=0. Hence, 𝐳\mathbf{z} is an inner η\eta-critical point according to Definition 2.12(1).

Case 2a: 𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and m3m\leq 3.

These cases correspond to acute critical/flat η\eta-critical/concave η\eta-critical points according to Definition 2.12(2-4).

Case 2b: 𝐳𝒱Ω(𝒯)\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right) and m4m\geq 4.

We argue as in Case 1 but take into account that the patch 𝒯𝐳\mathcal{T}_{\mathbf{z}} is not “closed” since 𝐳\mathbf{z} is a boundary point. Let α:=2πi=1mωi\alpha:=2\pi-\sum_{i=1}^{m}\omega_{i} be the “outer angle” of the domain at 𝐳\mathbf{z}. Then

ω1+ωm+2i=2m1ωi=i=1m1(ωi+ωi+1)=(m1)π+i=1m1δi.\omega_{1}+\omega_{m}+2\sum_{i=2}^{m-1}\omega_{i}=\sum_{i=1}^{m-1}\left(\omega_{i}+\omega_{i+1}\right)=\left(m-1\right)\pi+\sum_{i=1}^{m-1}\delta_{i}.

By the definition of α\alpha, we obtain

(m1)π+i=1m1δi\displaystyle\left(m-1\right)\pi+\sum_{i=1}^{m-1}\delta_{i} =2πα+i=2m1ωi=2πα+=1m22(ω2+ω2+1)+Q(m)ωm1\displaystyle=2\pi-\alpha+\sum_{i=2}^{m-1}\omega_{i}=2\pi-\alpha+\sum_{\ell=1}^{\left\lfloor\frac{m-2}{2}\right\rfloor}\left(\omega_{2\ell}+\omega_{2\ell+1}\right)+Q\left(m\right)\omega_{m-1}
=2πα+m22π+=1m22δ2+Q(m)ωm1,\displaystyle=2\pi-\alpha+\left\lfloor\frac{m-2}{2}\right\rfloor\pi+\sum_{\ell=1}^{\left\lfloor\frac{m-2}{2}\right\rfloor}\delta_{2\ell}+Q\left(m\right)\omega_{m-1},

where Q(m)=0Q\left(m\right)=0 if mm is even and Q(m)=1Q\left(m\right)=1 if mm is odd. By rearranging the terms we get

Q(m)ωm1+Δm=(m3m22)π+αfor Δm:==1m22δ2i=1m1δi.Q\left(m\right)\omega_{m-1}+\Delta_{m}=\left(m-3-\left\lfloor\frac{m-2}{2}\right\rfloor\right)\pi+\alpha\quad\text{for\quad}\Delta_{m}:=\sum_{\ell=1}^{\left\lfloor\frac{m-2}{2}\right\rfloor}\delta_{2\ell}-\sum_{i=1}^{m-1}\delta_{i}. (2.12)

We adjust η0\eta_{0}^{\prime} such that δ=arcsinη0\delta=\arcsin\eta_{0}^{\prime} satisfies mmaxδ<αm_{\max}\delta<\alpha and, in turn, |Δm|mmaxδ<α\left|\Delta_{m}\right|\leq m_{\max}\delta<\alpha. Then, it is easy to verify that

|Q(m)ωm1+Δm|<Q(m)π+α(m3m22)π+α\left|Q\left(m\right)\omega_{m-1}+\Delta_{m}\right|<Q\left(m\right)\pi+\alpha\leq\left(m-3-\left\lfloor\frac{m-2}{2}\right\rfloor\right)\pi+\alpha

holds for all m4m\geq 4. Hence, (2.12) cannot hold and there exists no η\eta-critical boundary point 𝐳\mathbf{z} for m4m\geq 4.

 

Next, we collect the η\eta-critical points in pairwise disjoint, edge-connected sets which we will define in the following.

Refer to caption
Figure 3: Three types of obtuse η\eta-critical points 𝐳𝒞𝒯obtuse(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right) with associated inner edge 𝔈(𝐳)\mathfrak{E}\left(\mathbf{z}\right), normal vector 𝔑(𝐳)\mathfrak{N}\left(\mathbf{z}\right) and opposite endpoint 𝔙(𝐳)\mathfrak{V}\left(\mathbf{z}\right) of 𝔈(𝐳)\mathfrak{E}\left(\mathbf{z}\right); left: inner η\eta-critical point, middle: flat η\eta-critical point, right: concave η\eta-critical point.

We say two points 𝐲,𝐲𝒱(𝒯)\mathbf{y},\mathbf{y}^{\prime}\in\mathcal{V}\left(\mathcal{T}\right) are edge-connected if there is an edge E(𝒯)E\in\mathcal{E}\left(\mathcal{T}\right) with endpoints 𝐲,𝐲\mathbf{y,y}^{\prime}. A subset 𝒱𝒱(𝒯)\mathcal{V}^{\prime}\subset\mathcal{V}\left(\mathcal{T}\right) is edge-connected if there is a numbering of the points in 𝒱={𝐲j:1jn}\mathcal{V}^{\prime}=\left\{\mathbf{y}_{j}:1\leq j\leq n\right\} such that 𝐲j1\mathbf{y}_{j-1}, 𝐲j\mathbf{y}_{j} are edge-connected for all 2jn2\leq j\leq n. A point 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right) is edge-connected to 𝒱\mathcal{V}^{\prime} if 𝐳𝒱\mathbf{z}\in\mathcal{V}^{\prime} or there is 𝐲𝒱\mathbf{y\in}\mathcal{V}^{\prime} such that 𝐳\mathbf{z}, 𝐲\mathbf{y} are edge-connected.

From Lemma 2.10 we know that two edge-connected points 𝐳,𝐳𝒱(𝒯)\mathbf{z},\mathbf{z}^{\prime}\in\mathcal{V}\left(\mathcal{T}\right) can be both critical only if the connecting edge EE belongs to Ω(𝒯)\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right); in this case it holds 𝐳,𝐳𝒱Ω(𝒯)\mathbf{z},\mathbf{z}^{\prime}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right). Next, we will group the points in 𝒞𝒯(η)\mathcal{C}_{\mathcal{T}}\left(\eta\right) into subsets called fans.

From Lemma 2.10 it follows that the points in 𝒞𝒯inner(η)\mathcal{C}_{\mathcal{T}}^{\operatorname*{inner}}\left(\eta\right) are isolated (see Def. 2.11). All other η\eta-critical points lie on the boundary. Next, we define mappings 𝔈:𝒞𝒯obtuse(η)Ω(𝒯)\mathfrak{E}:\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right)\rightarrow\mathcal{E}_{\Omega}\left(\mathcal{T}\right), 𝔑:𝒞𝒯obtuse(η)𝕊2\mathfrak{N}:\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right)\rightarrow\mathbb{S}_{2}, and 𝔙:𝒞𝒯obtuse(η)𝒱(𝒯)\𝒞𝒯(η)\mathfrak{V}:\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right)\rightarrow\mathcal{V}\left(\mathcal{T}\right)\backslash\mathcal{C}_{\mathcal{T}}\left(\eta\right). The construction is illustrated in Figure 3.

For 𝐳𝒞𝒯obtuse(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right), Definition 2.12 implies that |𝐳|3\left|\mathcal{E}_{\mathbf{z}}\right|\geq 3 and hence 𝐳Ω(𝒯)\mathcal{E}_{\mathbf{z}}\cap\mathcal{E}_{\Omega}\left(\mathcal{T}\right)\neq\emptyset. We fix one edge E𝐳Ω(𝒯)E\in\mathcal{E}_{\mathbf{z}}\cap\mathcal{E}_{\Omega}\left(\mathcal{T}\right) and set 𝔈(𝐳):=E\mathfrak{E}\left(\mathbf{z}\right):=E. Note that the choice of EE is unique for 𝐳𝒞𝒯flat(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{flat}}\left(\eta\right). For 𝐳𝒞𝒯inner(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{inner}}\left(\eta\right) the choice is arbitrary. For 𝐳𝒞𝒯concave(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{concave}}\left(\eta\right), the set 𝐳Ω(𝒯)\mathcal{E}_{\mathbf{z}}\cap\mathcal{E}_{\Omega}\left(\mathcal{T}\right) consists of two edges, say E1E_{1}, E2E_{2}. We fix one of them and set 𝔈(𝐳):=E2\mathfrak{E}\left(\mathbf{z}\right):=E_{2}. Let 𝐳𝒱(𝒯)\mathbf{z}^{\prime}\in\mathcal{V}\left(\mathcal{T}\right) be such that 𝔈(𝐳)=[𝐳,𝐳]\mathfrak{E}\left(\mathbf{z}\right)=\left[\mathbf{z},\mathbf{z}^{\prime}\right]. Then 𝔙(𝐳):=𝐳\mathfrak{V}\left(\mathbf{z}\right):=\mathbf{z}^{\prime}. Lemma 2.10 implies that 𝐳\mathbf{z}^{\prime} is not an η\eta-critical point. A unit vector 𝔑(𝐳)\mathfrak{N}\left(\mathbf{z}\right) orthogonal to 𝔈(𝐳)\mathfrak{E}\left(\mathbf{z}\right) is defined by the condition that 𝐳𝐳\mathbf{z}^{\prime}-\mathbf{z} and 𝔑(𝐳)\mathfrak{N}\left(\mathbf{z}\right) form a right-handed system.

Definition 2.14

We decompose C𝒯obtuse(η)C_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right) into disjoint fans 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right), 𝒥\ell\in\mathcal{J}, such that the following conditions are satisfies

  1. 1.

    𝒞𝒯obtuse(η)=𝒥𝒞𝒯,(η),\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right)={\displaystyle\bigcup\limits_{\ell\in\mathcal{J}}}\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right),

  2. 2.

    for any 𝒥\ell\in\mathcal{J}, the set 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right) is edge-connected,

  3. 3.

    for any 𝒥\ell\in\mathcal{J}, there is 𝐳𝒱(𝒯)\𝒞𝒯(η)\mathbf{z}_{\ell}\in\mathcal{V}\left(\mathcal{T}\right)\backslash\mathcal{C}_{\mathcal{T}}\left(\eta\right) such that for all 𝐳𝒞𝒯,(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right) it holds 𝔙(𝐳)=𝐳\mathfrak{V}\left(\mathbf{z}\right)=\mathbf{z}_{\ell} and, vice versa:

  4. 4.

    any 𝐳𝒞𝒯obtuse(η)\mathbf{z}^{\prime}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right) which is edge-connected to some 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right) and satisfies 𝔙(𝐳)=𝐳\mathfrak{V}\left(\mathbf{z}^{\prime}\right)=\mathbf{z}_{\ell} belongs to 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right).

The following lemma will allow us to construct a right-inverse for the divergence operator separately for each fan.

Lemma 2.15

Let η0\eta_{0} be as in in Lemma 2.10 and let 0η<η00\leq\eta<\eta_{0} be fixed.

  1. a.

    Then, the mapping 𝔈:𝒞𝒯obtuse(η)Ω(𝒯)\mathfrak{E}:\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right)\rightarrow\mathcal{E}_{\Omega}\left(\mathcal{T}\right) is injective.

  2. b.

    For 𝒥\ell\in\mathcal{J}, let ω:=𝐳𝒞𝒯,(η)ω𝔈(𝐳)\omega_{\ell}:={\displaystyle\bigcup\limits_{\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)}}\omega_{\mathfrak{E}\left(\mathbf{z}\right)}. The domains ω\omega_{\ell} have pairwise disjoint interior.

Proof. Part a. The injectivity of the mapping 𝔈:𝒞𝒯obtuse(η)Ω(𝒯)\mathfrak{E}:\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right)\rightarrow\mathcal{E}_{\Omega}\left(\mathcal{T}\right) follows from Lemma 2.10: if 𝐳𝒞𝒯(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right) and 𝐳𝒱(𝒯)\mathbf{z}^{\prime}\in\mathcal{V}\left(\mathcal{T}\right) is such that E:=[𝐳,𝐳]Ω(𝒯)E:=\left[\mathbf{z},\mathbf{z}^{\prime}\right]\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right) then 𝐳𝒞𝒯(η)\mathbf{z}^{\prime}\notin\mathcal{C}_{\mathcal{T}}\left(\eta\right).

Part b. The following construction is illustrated in Figure 4.

Refer to caption
Figure 4: Nodal patch, illustrating edge-connected obtuse η\eta-critical points of a fan 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right). In this example, the left-most η\eta-critical point is 𝐳,n\mathbf{z}_{\ell,n_{\ell}} and of type “flat”, the right-most is 𝐳,1\mathbf{z}_{\ell,1} of type “concave” with 𝔈(𝐳,1)=[𝐳,1,𝐳]\mathfrak{E}\left(\mathbf{z}_{\ell,1}\right)=\left[\mathbf{z}_{\ell,1},\mathbf{z}\right]. The extremal points 𝐳,0\mathbf{z}_{\ell,0} and 𝐳,n+1\mathbf{z}_{\ell,n_{\ell}+1} do not belong to 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right). The edge connecting 𝐳\mathbf{z}_{\ell} with 𝐳,j\mathbf{z}_{\ell,j} is denoted by E,jE_{\ell,j}.

For fixed 𝒥\ell\in\mathcal{J}, we number the points in 𝒞𝒯,\mathcal{C}_{\mathcal{T},\ell} by 𝐳,j\mathbf{z}_{\ell,j}, 1jn1\leq j\leq n_{\ell}, such that 𝐳,j1\mathbf{z}_{\ell,j-1} and 𝐳,j\mathbf{z}_{\ell,j} are edge-connected for all 2jn2\leq j\leq n_{\ell} and 𝐳,1\mathbf{z}_{\ell,1} and 𝐳,n\mathbf{z}_{\ell,n_{\ell}} are the endpoints in the polygonal line through these points. Let K,jK_{\ell,j} be the triangle with vertices 𝐳,j1\mathbf{z}_{\ell,j-1}, 𝐳\mathbf{z}_{\ell}, 𝐳,j\mathbf{z}_{\ell,j}, 2jn2\leq j\leq n_{\ell}\ and let 𝒯:={K,j:2jn}\mathcal{T}_{\ell}:=\left\{K_{\ell,j}:2\leq j\leq n_{\ell}\right\}. Note that this set is empty if |𝒞𝒯,|=1\left|\mathcal{C}_{\mathcal{T},\ell}\right|=1. Let K,1,K,n+1𝒯\𝒯K_{\ell,1},K_{\ell,n_{\ell}+1}\in\mathcal{T}\backslash\mathcal{T}_{\ell} be two different triangles such that E,1:=[𝐳,1,𝐳]K,1E_{\ell,1}:=\left[\mathbf{z}_{\ell,1},\mathbf{z}_{\ell}\right]\subset\partial K_{\ell,1} and E,n:=[𝐳,n,𝐳]K,n+1E_{\ell,n_{\ell}}:=\left[\mathbf{z}_{\ell,n_{\ell}},\mathbf{z}_{\ell}\right]\subset\partial K_{\ell,n_{\ell}+1}. Let 𝐳,0\mathbf{z}_{\ell,0} be the third vertex in K,1K_{\ell,1} and observe that it does not belong to 𝒞𝒯,\mathcal{C}_{\mathcal{T},\ell}. Since E,1E_{\ell,1} is an inner edge and 𝐳,1\mathbf{z}_{\ell,1} is an η\eta-critical point, Lemma 2.10 implies that 𝐳\mathbf{z}_{\ell} is not an η\eta-critical point. Next, we show that E,0:=[𝐳,0,𝐳]E_{\ell,0}:=\left[\mathbf{z}_{\ell,0},\mathbf{z}_{\ell}\right] does not belong to 𝒯obtuse:={𝔈(𝐳):𝐳𝒞𝒯obtuse(η)}\mathcal{E}_{\mathcal{T}}^{\operatorname*{obtuse}}:=\left\{\mathfrak{E}\left(\mathbf{z}\right):\mathbf{z}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right)\right\}; from this the assertion follows. We assume E,0𝒯obtuseE_{\ell,0}\in\mathcal{E}_{\mathcal{T}}^{\operatorname*{obtuse}} and derive a contradiction. Since 𝐳𝒞𝒯obtuse(η)\mathbf{z}_{\ell}\notin\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right), this assumption implies that 𝐳,0𝒞𝒯obtuse(η)\mathbf{z}_{\ell,0}\in\mathcal{C}_{\mathcal{T}}^{\operatorname*{obtuse}}\left(\eta\right). If 𝔙(𝐳,0)=𝐳\mathfrak{V}\left(\mathbf{z}_{\ell,0}\right)=\mathbf{z}_{\ell} then Definition 2.14(4) implies that 𝐳,0𝒞𝒯,\mathbf{z}_{\ell,0}\in\mathcal{C}_{\mathcal{T},\ell} and this is a contradiction. If 𝔙(𝐳,0)𝐳\mathfrak{V}\left(\mathbf{z}_{\ell,0}\right)\neq\mathbf{z}_{\ell}, then E,0𝒯obtuseE_{\ell,0}\notin\mathcal{E}_{\mathcal{T}}^{\operatorname*{obtuse}}.   

Since the acute critical points need some special treatment we define a sequence of triangulations 𝒯i\mathcal{T}_{i}, 1iL1\leq i\leq L, with the properties

  1. 1.
    𝒯1𝒯2𝒯L=𝒯,\mathcal{T}_{1}\subset\mathcal{T}_{2}\subset\ldots\subset\mathcal{T}_{L}=\mathcal{T}, (2.13)
  2. 2.

    𝒯1\mathcal{T}_{1} is a maximal subset of 𝒯\mathcal{T} such that 𝒞𝒯1acute=\mathcal{C}_{\mathcal{T}_{1}}^{\operatorname{acute}}=\emptyset,

  3. 3.

    for j=1,2,,Lj=1,2,\ldots,L,

    𝒯j={K𝒯at least one edge of K belongs to (𝒯j1)}.\mathcal{T}_{j}=\left\{K\in\mathcal{T}\mid\text{at least one edge of }K\text{ belongs to }\mathcal{E}\left(\mathcal{T}_{j-1}\right)\right\}.

By this step-by-step procedure, triangles are attached to a previous triangulation 𝒯j1\mathcal{T}_{j-1} which have an edge in common with the set of edges in 𝒯j1\mathcal{T}_{j-1}. The proof of Theorem 1.3 under assumption (2.7) then consists of first proving the inf-sup stability for 𝒯1\mathcal{T}_{1} and then to investigate the effect of attaching a triangle to an inf-sup stable triangulation. A sufficient condition for L=0L=0 is that every triangle in 𝒯\mathcal{T} has an interior point.

2.2.2 The case 𝒞𝒯acute=\mathcal{C}_{\mathcal{T}}^{\operatorname{acute}}=\emptyset

In this section, we prove the inf-sup stability for the triangulation 𝒯1\mathcal{T}_{1} in (2.13) where 𝒞𝒯1acute=\mathcal{C}_{\mathcal{T}_{1}}^{\operatorname{acute}}=\emptyset. For simplicity we skip the index 11 and write 𝒯\mathcal{T}, 𝒞𝒯(η)\mathcal{C}_{\mathcal{T}}\left(\eta\right), etc.

Next, we define some fundamental non-conforming Crouzeix-Raviart vector fields which will be used to eliminate the critical pressures in the Stokes element (𝐒k,0(𝒯),k1,0(𝒯))\left(\mathbf{S}_{k,0}\left(\mathcal{T}\right),\mathbb{P}_{k-1,0}\left(\mathcal{T}\right)\right).

Essential properties of the Crouzeix-Raviart function Bk,ECRB_{k,E}^{\operatorname*{CR}} are: it is a polynomial of degree kk on each K𝒯EK\subset\mathcal{T}_{E} and a Legendre polynomial on each edge EωEE^{\prime}\subset\partial\omega_{E} so that the jump relations in (1.9b) are satisfied. Furthermore, [Bk,ECR]E=0\left[B_{k,E}^{\operatorname*{CR}}\right]_{E}=0 and suppBk,ECRωE\operatorname*{supp}B_{k,E}^{\operatorname*{CR}}\subset\omega_{E}.

In the first step, we modify the function Bk,ECRB_{k,E}^{\operatorname*{CR}} by adding a conforming edge bubble in Sk,0(𝒯)S_{k,0}\left(\mathcal{T}\right) such that the H1(Ω)H^{1}\left(\Omega\right) norm of the modified function has an improved behaviour with respect to kk.

Let EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right) with endpoints 𝐕1\mathbf{V}_{1}, 𝐕2\mathbf{V}_{2}. Set 𝐭E=(𝐕2𝐕1)/𝐕2𝐕1\mathbf{t}_{E}=\left(\mathbf{V}_{2}-\mathbf{V}_{1}\right)/\left\|\mathbf{V}_{2}-\mathbf{V}_{1}\right\| and consider a function wEk(𝒯)w_{E}\in\mathbb{P}_{k}\left(\mathcal{T}\right) with suppwE=ωE\operatorname*{supp}w_{E}=\omega_{E} and

wE|K|E\displaystyle\left.\left.w_{E}\right|_{K}\right|_{E^{\prime}} =Bk,ECR|K|EK𝒯E and EKωE,\displaystyle=\left.\left.B_{k,E}^{\operatorname*{CR}}\right|_{K}\right|_{E^{\prime}}\quad\forall K\in\mathcal{T}_{E}\text{ and }E^{\prime}\subset\partial K\cap\partial\omega_{E}, (2.14a)
[wE]E\displaystyle\left[w_{E}\right]_{E} =0and 𝐭EwE(𝐕1)=𝐭EwE(𝐕2)=0.\displaystyle=0\quad\text{and\quad}\partial_{\mathbf{t}_{E}}w_{E}\left(\mathbf{V}_{1}\right)=\partial_{\mathbf{t}_{E}}w_{E}\left(\mathbf{V}_{2}\right)=0. (2.14b)

Then, wEw_{E} also belongs to the space CRk,0(𝒯)\operatorname*{CR}_{k,0}\left(\mathcal{T}\right) and

(wE|K)(𝐳)=(Bk,ECR|K)(𝐳)K𝒯𝐳𝒱(K).\nabla\left(\left.w_{E}\right|_{K}\right)\left(\mathbf{z}\right)=\nabla\left(\left.B_{k,E}^{\operatorname*{CR}}\right|_{K}\right)\left(\mathbf{z}\right)\quad\forall K\in\mathcal{T\quad\forall}\mathbf{z}\in\mathcal{V}\left(K\right). (2.15)

The last relation can be derived from the following reasoning. For K𝒯K\in\mathcal{T} and 𝐳𝒱(K)\mathbf{z}\in\mathcal{V}\left(K\right), set 𝐭𝐲:=(𝐲𝐳)/𝐲𝐳\mathbf{t}_{\mathbf{y}}:=\left(\mathbf{y}-\mathbf{z}\right)/\left\|\mathbf{y-z}\right\| for all 𝐲𝒱(K)\{𝐳}\mathbf{y}\in\mathcal{V}\left(K\right)\backslash\left\{\mathbf{z}\right\}. Let 𝐜2\mathbf{c}\in\mathbb{R}^{2} be arbitrary. Clearly, 𝐜=𝐲𝒱(K)\{𝐳}α𝐲𝐭𝐲\mathbf{c}=\sum_{\mathbf{y}\in\mathcal{V}\left(K\right)\backslash\left\{\mathbf{z}\right\}}\alpha_{\mathbf{y}}\mathbf{t}_{\mathbf{y}} for some α𝐲\alpha_{\mathbf{y}}\in\mathbb{R}. The conditions in (2.14) imply that for 𝐲𝒱(K)\{𝐳}\mathbf{y}\in\mathcal{V}\left(K\right)\backslash\left\{\mathbf{z}\right\} it holds

wE|K𝐭𝐲=Bk,ECR|K𝐭𝐲.\frac{\partial\left.w_{E}\right|_{K}}{\partial\mathbf{t}_{\mathbf{y}}}=\frac{\partial\left.B_{k,E}^{\operatorname*{CR}}\right|_{K}}{\partial\mathbf{t}_{\mathbf{y}}}.

Hence,

(wE|K),𝐜(𝐳)\displaystyle\left\langle\nabla\left(\left.w_{E}\right|_{K}\right),\mathbf{c}\right\rangle\left(\mathbf{z}\right) =𝐲𝒱(K)\{𝐳}α𝐲wE|K𝐭𝐲(𝐳)\displaystyle=\sum_{\mathbf{y}\in\mathcal{V}\left(K\right)\backslash\left\{\mathbf{z}\right\}}\alpha_{\mathbf{y}}\frac{\partial\left.w_{E}\right|_{K}}{\partial\mathbf{t}_{\mathbf{y}}}\left(\mathbf{z}\right) (2.16)
=𝐲𝒱(K)\{𝐳}α𝐲Bk,ECR|K𝐭𝐲(𝐳)=((Bk,ECR)|K),𝐜(𝐳).\displaystyle=\sum_{\mathbf{y}\in\mathcal{V}\left(K\right)\backslash\left\{\mathbf{z}\right\}}\alpha_{\mathbf{y}}\frac{\partial\left.B_{k,E}^{\operatorname*{CR}}\right|_{K}}{\partial\mathbf{t}_{\mathbf{y}}}\left(\mathbf{z}\right)=\left\langle\nabla\left(\left.\left(B_{k,E}^{\operatorname*{CR}}\right)\right|_{K}\right),\mathbf{c}\right\rangle\left(\mathbf{z}\right).

Since 𝐜\mathbf{c} was arbitrary, (2.15) follows.

Lemma 2.16

Let k5k\geq 5 be odd and for EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right), let Bk,ECRB_{k,E}^{\operatorname*{CR}} be as in (2.4). Then, there exists a function B~k,ECRCRk,0(𝒯)\tilde{B}_{k,E}^{\operatorname*{CR}}\in\operatorname*{CR}_{k,0}\left(\mathcal{T}\right) with

  1. i.

    suppB~k,ECR=ωE,\operatorname*{supp}\tilde{B}_{k,E}^{\operatorname*{CR}}=\omega_{E},

  2. ii.

    for all K𝒯K\in\mathcal{T}, for all 𝐳𝒱(K)\mathbf{z}\in\mathcal{V}\left(K\right):

    (B~k,ECR|K)(𝐳)=(Bk,ECR|K)(𝐳),\nabla\left(\left.\tilde{B}_{k,E}^{\operatorname*{CR}}\right|_{K}\right)\left(\mathbf{z}\right)=\nabla\left(\left.B_{k,E}^{\operatorname*{CR}}\right|_{K}\right)\left(\mathbf{z}\right), (2.17)
  3. iii.

    for all K𝒯EK\in\mathcal{T}_{E}, for all EKωEE^{\prime}\subset\partial K\cap\mathcal{\partial}\omega_{E}:

    B~k,ECR|K|E=Bk,ECR|K|Eand [B~k,ECR]E=0,\left.\left.\tilde{B}_{k,E}^{\operatorname*{CR}}\right|_{K}\right|_{E^{\prime}}=\left.\left.B_{k,E}^{\operatorname*{CR}}\right|_{K}\right|_{E^{\prime}}\quad\text{and\quad}\left[\tilde{B}_{k,E}^{\operatorname*{CR}}\right]_{E}=0,
  4. iv.

    for all K𝒯K\in\mathcal{T}

    Kdiv𝒯(B~k,ECR𝐧E)=0.\int_{K}\operatorname{div}_{\mathcal{T}}\left(\tilde{B}_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}\right)=0. (2.18)
  5. v.

    The piecewise gradient is bounded:

    𝒯B~k,ECR𝐋2(Ω)Clog(k+1).\left\|\nabla_{\mathcal{T}}\tilde{B}_{k,E}^{\operatorname*{CR}}\right\|_{\mathbf{L}^{2}\left(\Omega\right)}\leq C\sqrt{\log\left(k+1\right)}. (2.19)

Proof. We employ the reference triangle as in [6] in order to apply the polynomial extension theorem therein. Let K~\tilde{K} be the equilateral triangle with vertices 𝐀~1:=(1,0)T\mathbf{\tilde{A}}_{1}:=\left(-1,0\right)^{T}, 𝐀~2:=(1,0)T\mathbf{\tilde{A}}_{2}:=\left(1,0\right)^{T}, 𝐀~3:=(0,3)T\mathbf{\tilde{A}}_{3}:=\left(0,\sqrt{3}\right)^{T} and let E~j\tilde{E}_{j} denote the edge in K~\tilde{K} opposite to 𝐀~j\mathbf{\tilde{A}}_{j}, 1j31\leq j\leq 3.

Let EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right) with endpoints 𝐕1\mathbf{V}_{1}, 𝐕2\mathbf{V}_{2}, and let K𝒯EK\in\mathcal{T}_{E}. Choose an affine pullback ϕK:K~K\phi_{K}:\tilde{K}\rightarrow K such that ϕK(E~3)=E\phi_{K}\left(\tilde{E}_{3}\right)=E. We employ the function ψ~k±k([1,1])\tilde{\psi}_{k}^{\pm}\in\mathbb{P}_{k}\left(\left[-1,1\right]\right) given by

ψ~k(x):=ck1(1+x)(1x)24Pk3(3,3)(x)and ψ~k+(x):=ψ~k(x)\tilde{\psi}_{k}^{-}\left(x\right):=c_{k}^{-1}\frac{\left(1+x\right)\left(1-x\right)^{2}}{4}P_{k-3}^{\left(3,3\right)}\left(x\right)\quad\text{and\quad}\tilde{\psi}_{k}^{+}\left(x\right):=-\tilde{\psi}_{k}^{-}\left(-x\right)

with the Jacobi polynomials Pn(α,β)P_{n}^{\left(\alpha,\beta\right)} (see, e.g., [17, §18.3]) and the normalisation factor ck:=(1)k1(k3)c_{k}:=\left(-1\right)^{k-1}\binom{k}{3}. These functions have been analysed in the proof of Lemma A.1 in [3, denoted by FkF_{k}] and we recall relevant properties. It holds ψ~k±(±1)=(ψ~k)(+1)=(ψ~k+)(1)=0\tilde{\psi}_{k}^{\pm}\left(\pm 1\right)=\left(\tilde{\psi}_{k}^{-}\right)^{\prime}\left(+1\right)=\left(\tilde{\psi}_{k}^{+}\right)^{\prime}\left(-1\right)=0 and (ψ~k)(1)=(ψ~k+)(+1)=1\left(\tilde{\psi}_{k}^{-}\right)^{\prime}\left(-1\right)=\left(\tilde{\psi}_{k}^{+}\right)^{\prime}\left(+1\right)=1 (cf. [17, §18.3]). Their norms can be estimated by

ψ~k±L2([1,1])Ck3and (ψ~k±)L2([1,1])Ck1.\left\|\tilde{\psi}_{k}^{\pm}\right\|_{L^{2}\left(\left[-1,1\right]\right)}\leq Ck^{-3}\quad\text{and\quad}\left\|\left(\tilde{\psi}_{k}^{\pm}\right)^{\prime}\right\|_{L^{2}\left(\left[-1,1\right]\right)}\leq Ck^{-1}.

We set

φ~k(x)=Lk1(x)Lk1(1)ψ~k(x)Lk1(1)ψ~k+(x).\tilde{\varphi}_{k}\left(x\right)=L_{k-1}\left(x\right)-L_{k-1}^{\prime}\left(-1\right)\tilde{\psi}_{k}^{-}\left(x\right)-L_{k-1}^{\prime}\left(1\right)\tilde{\psi}_{k}^{+}\left(x\right). (2.20)

Clearly, it holds

φ~k(±1)=1,φ~k(±1)=0.\tilde{\varphi}_{k}\left(\pm 1\right)=1,\quad\tilde{\varphi}_{k}^{\prime}\left(\pm 1\right)=0.

By using Lemma C.1, we get

φ~kL2([1,1])Ck1/2,φ~kH1([1,1])Ck.\left\|\tilde{\varphi}_{k}\right\|_{L^{2}\left(\left[-1,1\right]\right)}\leq Ck^{-1/2},\quad\left\|\tilde{\varphi}_{k}\right\|_{H^{1}\left(\left[-1,1\right]\right)}\leq Ck.

From [6, Thm. 7.4] we conclude that there is w~Ek(K~)\tilde{w}_{E}\in\mathbb{P}_{k}\left(\tilde{K}\right) with

w~E|E~j=Bk,ECR|KϕK|E~j,j=1,2and w~E|E~3=φ~k\left.\tilde{w}_{E}\right|_{\tilde{E}_{j}}=\left.\left.B_{k,E}^{\operatorname*{CR}}\right|_{K}\circ\phi_{K}\right|_{\tilde{E}_{j}},\quad j=1,2\quad\text{and\quad}\left.\tilde{w}_{E}\right|_{\tilde{E}_{3}}=\tilde{\varphi}_{k}

which satisfies

w~EH1(K~)Cw~EH1/2(K~)\left\|\tilde{w}_{E}\right\|_{H^{1}\left(\tilde{K}\right)}\leq C\left\|\tilde{w}_{E}\right\|_{H^{1/2}\left(\partial\tilde{K}\right)}

for a constant CC independent of kk. Lemma B.2 implies the following estimate of the H1/2H^{1/2} norm of w~E\tilde{w}_{E}:

w~EH1/2(K~)Clog(k+1).\left\|\tilde{w}_{E}\right\|_{H^{1/2}\left(\partial\tilde{K}\right)}\leq C\sqrt{\log\left(k+1\right)}. (2.21)

In turn, we get

w~EH1(K~)Clog(k+1).\left\|\tilde{w}_{E}\right\|_{H^{1}\left(\tilde{K}\right)}\leq C\sqrt{\log\left(k+1\right)}. (2.22)

By using the affine lifting ϕK\phi_{K} to the triangle KK we define the function wEw_{E} by

wE|K:={w~EϕK1if K𝒯E,0otherwise.\left.w_{E}\right|_{K}:=\left\{\begin{array}[c]{ll}\tilde{w}_{E}\circ\phi_{K}^{-1}&\text{if }K\in\mathcal{T}_{E}\text{,}\\ 0&\text{otherwise.}\end{array}\right.

This function is continuous across EE (with value φ~kϕK1|E\tilde{\varphi}_{k}\circ\left.\phi_{K}^{-1}\right|_{E}) and, on EωEE^{\prime}\subset\partial\omega_{E}, it is a lifted Legendre polynomial. This implies property (iii) for wEw_{E}. The function wEw_{E} vanishes outside ωE\omega_{E} so that (i) holds. Since the construction implies that the derivative of wEw_{E} in the direction of EE, evaluated at the endpoints 𝐕1\mathbf{V}_{1}, 𝐕2\mathbf{V}_{2} of EE, is zero, we may apply the reasoning in (2.16) to obtain property (ii) for wEw_{E}.

From (2.22) we obtain by the transformation rule for integrals and the chain rule for differentiation

wE𝐋2(K)Cw~EH1(K~)Clog(k+1)K𝒯E.\left\|\nabla w_{E}\right\|_{\mathbf{L}^{2}\left(K\right)}\leq C\left\|\tilde{w}_{E}\right\|_{H^{1}\left(\tilde{K}\right)}\leq C\sqrt{\log\left(k+1\right)}\quad\forall K\in\mathcal{T}_{E}\text{.}

Next, we modify wEw_{E} such that property (iv) holds without affecting the other properties. Let ψES4,0(𝒯)\psi_{E}\in S_{4,0}\left(\mathcal{T}\right) with suppψE=ωE\operatorname*{supp}\psi_{E}=\omega_{E} and

ψE|K:=αKλK,𝐕12λK,𝐕22with αK:=(K𝐧EwE)/(K𝐧E(λK,𝐕12λK,𝐕22))K𝒯E.\left.\psi_{E}\right|_{K}:=\alpha_{K}\lambda_{K,\mathbf{V}_{1}}^{2}\lambda_{K,\mathbf{V}_{2}}^{2}\quad\text{with\quad}\alpha_{K}:=\left(\int_{K}\partial_{\mathbf{n}_{E}}w_{E}\right)/\left(\int_{K}\partial_{\mathbf{n}_{E}}\left(\lambda_{K,\mathbf{V}_{1}}^{2}\lambda_{K,\mathbf{V}_{2}}^{2}\right)\right)\quad\forall K\in\mathcal{T}_{E}. (2.23)

The modified function B~k,ECR\tilde{B}_{k,E}^{\operatorname*{CR}} finally is defined by

B~k,ECR=wEψE.\tilde{B}_{k,E}^{\operatorname*{CR}}=w_{E}-\psi_{E}. (2.24)

Since div𝒯(B~k,ECR𝐧E)|K=𝐧E(B~|Kk,ECR)\left.\operatorname*{div}_{\mathcal{T}}\left(\tilde{B}_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}\right)\right|_{K}=\partial_{\mathbf{n}_{E}}\left(\tilde{B}\left.{}_{k,E}^{\operatorname*{CR}}\right|_{K}\right) property (iv) follows by construction. The gradient 𝒯ψE\nabla_{\mathcal{T}}\psi_{E} vanishes in the vertices of KK so that (𝐧EψE)(𝐳)=0\left(\partial_{\mathbf{n}_{E}}\psi_{E}\right)\left(\mathbf{z}\right)=0 for all 𝐳𝒱(K)\mathbf{z}\in\mathcal{V}\left(K\right) and (ii) is inherited from wEw_{E}. Properties (i), (iii) are obvious. Next, we verify (v). Let 𝐕3\mathbf{V}_{3} denote the vertex in KK opposite to EE. We first compute

K𝐧E(λK,𝐕12λK,𝐕22)\displaystyle\int_{K}\partial_{\mathbf{n}_{E}}\left(\lambda_{K,\mathbf{V}_{1}}^{2}\lambda_{K,\mathbf{V}_{2}}^{2}\right) =j=122𝐧EλK,𝐕jKλK,𝐕1λK,𝐕2λK,𝐕3j=KλK,𝐕12λK,𝐕2j=122𝐧EλK,𝐕j\displaystyle=\sum_{j=1}^{2}2\partial_{\mathbf{n}_{E}}\lambda_{K,\mathbf{V}_{j}}\int_{K}\lambda_{K,\mathbf{V}_{1}}\lambda_{K,\mathbf{V}_{2}}\lambda_{K,\mathbf{V}_{3-j}}=\int_{K}\lambda_{K,\mathbf{V}_{1}}^{2}\lambda_{K,\mathbf{V}_{2}}\sum_{j=1}^{2}2\partial_{\mathbf{n}_{E}}\lambda_{K,\mathbf{V}_{j}}
=115𝐧EλK,𝐕3|K|=(2.3)|E|30,\displaystyle=-\frac{1}{15}\partial_{\mathbf{n}_{E}}\lambda_{K,\mathbf{V}_{3}}\left|K\right|\overset{\text{(\ref{normcomp})}}{=}\frac{\left|E\right|}{30},
|K𝐧EwE|\displaystyle\left|\int_{K}\partial_{\mathbf{n}_{E}}w_{E}\right| |K|1/2wE𝐋2(K)C|K|1/2log(k+1).\displaystyle\leq\left|K\right|^{1/2}\left\|\nabla w_{E}\right\|_{\mathbf{L}^{2}\left(K\right)}\leq C\left|K\right|^{1/2}\sqrt{\log\left(k+1\right)}.

In this way, |αK|Clog(k+1)\left|\alpha_{K}\right|\leq C\sqrt{\log\left(k+1\right)} and an inverse inequality for quartic polynomials gives us

ψE𝐋2(K)ChK1ψEL2(K)ChK1log(k+1)1L2(K)Clog(k+1).\left\|\nabla\psi_{E}\right\|_{\mathbf{L}^{2}\left(K\right)}\leq Ch_{K}^{-1}\left\|\psi_{E}\right\|_{L^{2}\left(K\right)}\leq Ch_{K}^{-1}\sqrt{\log\left(k+1\right)}\left\|1\right\|_{L^{2}\left(K\right)}\leq C\sqrt{\log\left(k+1\right)}.

Hence, property (iv) follows.  

Next we recall a result which goes back to Vogelius [37] and Scott-Vogelius [32], see also [27, Proof of Thm. 1].

Definition 2.17

Let η0\eta_{0} be as in Lemma 2.10. For 0η<η00\leq\eta<\eta_{0}, the subspace Mη,k1SV(𝒯)M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) of the pressure space Mk1(𝒯)M_{k-1}\left(\mathcal{T}\right) is given by

Mη,k1SV(𝒯):={qMk1(𝒯)𝐳𝒞𝒯(η):A𝒯,𝐳(q)=0},M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right):=\left\{q\in M_{k-1}\left(\mathcal{T}\right)\mid\forall\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right):A_{\mathcal{T},\mathbf{z}}\left(q\right)=0\right\}, (2.25)

where, for 𝐳𝒞𝒯(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right), the functional A𝒯,𝐳(q)A_{\mathcal{T},\mathbf{z}}\left(q\right) is as follows: fix the counterclockwise numbering KK_{\ell}, 1m1\leq\ell\leq m, of the triangles in the patch 𝒯𝐳\mathcal{T}_{\mathbf{z}} by the condition K1K2=𝔈(𝐳)K_{1}\cap K_{2}=\mathfrak{E}\left(\mathbf{z}\right) and set

A𝒯,𝐳(q)==1m(1)(q|K)(𝐳).A_{\mathcal{T},\mathbf{z}}\left(q\right)=\sum_{\ell=1}^{m}\left(-1\right)^{\ell}\left(\left.q\right|_{K_{\ell}}\right)\left(\mathbf{z}\right). (2.26)

Note that M0,k1SV(𝒯)M_{0,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) is the pressure space introduced by Vogelius [37] and Scott-Vogelius [32] and the following inclusions hold: for 0ηηη00\leq\eta\leq\eta^{\prime}\leq\eta_{0}

Mη,k1SV(𝒯)Mη,k1SV(𝒯)M0,k1SV(𝒯)=Qhk1M_{\eta^{\prime},k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right)\subset M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right)\subset M_{0,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right)=Q_{h}^{k-1}

with the pressure space Qhk1Q_{h}^{k-1} in [27, p. 517].

For the Scott-Vogelius pressure space M0,k1SV(𝒯)M_{0,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right), the existence of a continuous right-inverse of the divergence operator into 𝐒k,0(𝒯)\mathbf{S}_{k,0}\left(\mathcal{T}\right) was proved in [37] and [32].

Proposition 2.18 (Scott-Vogelius)

For any pM0,k1SV(𝒯)p\in M_{0,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) there exists some 𝐯𝐒k,0(𝒯)\mathbf{v}\in\mathbf{S}_{k,0}\left(\mathcal{T}\right) such that

div𝐯=qand 𝐯𝐇1(Ω)CqL2(Ω),\operatorname*{div}\mathbf{v}=q\mathbf{\quad}\text{and\quad}\left\|\mathbf{v}\right\|_{\mathbf{H}^{1}\left(\Omega\right)}\leq C\left\|q\right\|_{L^{2}\left(\Omega\right)},

for a constant which only depends on the shape-regularity of the mesh, the polynomial degree kk, and on Θmin1\Theta_{\min}^{-1}, where

Θmin:=min𝐳𝒱(𝒯)\𝒞𝒯Θ(𝐳).\Theta_{\min}:=\min_{\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right)\backslash\mathcal{C}_{\mathcal{T}}}\Theta\left(\mathbf{z}\right). (2.27)

In particular, the constant CC is independent of hh.

In Lemma 2.20, we will show that, by subtracting the divergence of a suitable Crouzeix-Raviart velocity from a given pressure in Mk1(𝒯)M_{k-1}\left(\mathcal{T}\right), the resulting modified pressure belongs to the reduced pressure space Mη,k1SV(𝒯)M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right). As a preliminary, we need a bound of the functional A𝒯,𝐳A_{\mathcal{T},\mathbf{z}} in (2.26) which is explicit with respect to the local mesh size and polynomial degree.

Lemma 2.19

There exists a constant CC which only depends on the shape-regularity of the mesh such that

|A𝒯,𝐳(q)|Ck2h𝐳qL2(ω𝐳)qk1(𝒯)\left|A_{\mathcal{T},\mathbf{z}}\left(q\right)\right|\leq C\frac{k^{2}}{h_{\mathbf{z}}}\left\|q\right\|_{L^{2}\left(\omega_{\mathbf{z}}\right)}\quad\forall q\in\mathbb{P}_{k-1}\left(\mathcal{T}\right)

for any kk\in\mathbb{N}.

Proof. Let 𝐳𝒱(𝒯)\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right) and K𝒯𝐳K\in\mathcal{T}_{\mathbf{z}}. The affine pullback to the reference triangle is denoted by χK:K^K\chi_{K}:\widehat{K}\rightarrow K. For qk(K)q\in\mathbb{P}_{k}\left(K\right), let q^:=qχK\widehat{q}:=q\circ\chi_{K} and 𝐳^:=χK1(𝐳)\mathbf{\hat{z}}:=\chi_{K}^{-1}\left(\mathbf{z}\right). Then

|q(𝐳)|=|q^(𝐳^)|[38][2, Lem. 6.1](k+1)(k+2)2q^L2(K^)=(k+22)|K|1/2qL2(K).\left|q\left(\mathbf{z}\right)\right|=\left|\widehat{q}\left(\widehat{\mathbf{z}}\right)\right|\overset{\text{\cite[cite]{[\@@bibref{}{Hesthaven_trace_2003}{}{}]}, \cite[cite]{[\@@bibref{}{Ainsworth_Jiang_hp}{}{}, Lem. 6.1]}}}{\leq}\frac{\left(k+1\right)\left(k+2\right)}{\sqrt{2}}\left\|\widehat{q}\right\|_{L^{2}\left(\widehat{K}\right)}=\binom{k+2}{2}\left|K\right|^{-1/2}\left\|q\right\|_{L^{2}\left(K\right)}. (2.28)

A summation over all K𝒯𝐳K\in\mathcal{T}_{\mathbf{z}} leads to

|A𝒯,𝐳(q)|(k+22)=1m|K|1/2qL2(K)C(k+22)h𝐳1qL2(ω𝐳),\left|A_{\mathcal{T},\mathbf{z}}\left(q\right)\right|\leq\binom{k+2}{2}\sum_{\ell=1}^{m}\left|K_{\ell}\right|^{-1/2}\left\|q\right\|_{L^{2}\left(K_{\ell}\right)}\leq C\binom{k+2}{2}h_{\mathbf{z}}^{-1}\left\|q\right\|_{L^{2}\left(\omega_{\mathbf{z}}\right)},

where CC only depends on the shape-regularity of the mesh.  

The following lemma shows that the non-conforming Crouzeix-Raviart elements allow us to modify a general pressure qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right) in such a way that the result belongs to Mη,k1SV(𝒯)M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) provided 𝒞𝒯acute=\mathcal{C}_{\mathcal{T}}^{\operatorname{acute}}=\emptyset.

Lemma 2.20

Let assumption (2.7) be satisfied and 𝒞𝒯acute=\mathcal{C}_{\mathcal{T}}^{\operatorname{acute}}=\emptyset. There exists a constant η2>0\eta_{2}>0 which only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} (see (1.1)) such that for any fixed 0η<η20\leq\eta<\eta_{2} and any qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right), there exists some 𝐯q𝐂𝐑k,0(𝒯)\mathbf{v}_{q}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right) such that

Kdiv𝐯q\displaystyle\int_{K}\operatorname{div}\mathbf{v}_{q} =0K𝒯,\displaystyle=0\quad\forall K\in\mathcal{T}, (2.29)
qdiv𝒯𝐯q\displaystyle q-\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{v}_{q} Mη,k1SV(𝒯)\displaystyle\in M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) (2.30)

and

𝒯𝐯q𝕃2(Ω)CCRlog(k+1)qL2(Ω).\left\|\nabla_{\mathcal{T}}\mathbf{v}_{q}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}\leq C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\left\|q\right\|_{L^{2}\left(\Omega\right)}. (2.31)

The constant CCRC_{\operatorname*{CR}} depends only on the shape-regularity of the mesh and αΩ\alpha_{\Omega}.

Proof. Let qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right). Let the fans 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right), 𝒥\ell\in\mathcal{J}, be as in Definition 2.14. For each fan 𝒞𝒯,(η)\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right) we employ an ansatz

𝐯:=𝐳𝒞𝒯,(η)α,𝐳B~k,𝔈(𝐳)CR𝔑(𝐳)\mathbf{v}_{\ell}:=\sum_{\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)}\alpha_{\ell,\mathbf{z}}\tilde{B}_{k,\mathfrak{E}\left(\mathbf{z}\right)}^{\operatorname*{CR}}\mathfrak{N}\left(\mathbf{z}\right) (2.32)

for B~k,𝔈(𝐳)CR\tilde{B}_{k,\mathfrak{E}\left(\mathbf{z}\right)}^{\operatorname*{CR}} as in (2.24), where the coefficients α,𝐳\alpha_{\ell,\mathbf{z}}\in\mathbb{R} are defined next. The global function 𝐯q\mathbf{v}_{q} is then given by

𝐯q==1N𝐯.\mathbf{v}_{q}=\sum_{\ell=1}^{N}\mathbf{v}_{\ell}.

Property (2.29) follows from this ansatz by using Lemma 2.16. Next, we define the coefficients α,𝐳\alpha_{\ell,\mathbf{z}} in (2.32) such that (2.30) holds and prove the norm estimates for 𝐯q\mathbf{v}_{q}. Our construction of the fans implies that open interiors of the supports of 𝐯\mathbf{v}_{\ell} are pairwise disjoint (see Lem. 2.15); as a consequence the definition of (α,𝐳)𝐳𝒞𝒯(η)\left(\alpha_{\ell,\mathbf{z}}\right)_{\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right)} and the estimate of 𝒯𝐯q\nabla_{\mathcal{T}}\mathbf{v}_{q} can be performed for each fan separately.

For 𝐳𝒞𝒯,(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right), let E:=𝔈(𝐳)E:=\mathfrak{E}\left(\mathbf{z}\right), 𝐧E:=𝔑(𝐳)\mathbf{n}_{E}:=\mathfrak{N}\left(\mathbf{z}\right). Let K𝐳K_{\mathbf{z}}^{-}, K𝐳+K_{\mathbf{z}}^{+}, denote the triangles in 𝒯E\mathcal{T}_{E} with the convention that 𝐧E\mathbf{n}_{E} points into K𝐳+K_{\mathbf{z}}^{+}. The vertex in K𝐳±K_{\mathbf{z}}^{\pm} opposite to EE is denoted by 𝐀±\mathbf{A}^{\pm}. We use

div(B~k,ECR𝐧E|K)(𝐲)=(2.17)div(Bk,ECR𝐧E|K)(𝐲)𝐲𝒱(K)\operatorname*{div}\left(\left.\tilde{B}_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}\right|_{K}\right)\left(\mathbf{y}\right)\overset{\text{(\ref{neinher})}}{=}\operatorname*{div}\left(\left.B_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}\right|_{K}\right)\left(\mathbf{y}\right)\quad\forall\mathbf{y}\in\mathcal{V}\left(K\right)

and compute the divergence of Bk,ECR𝐧EB_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}

div(Bk,ECR𝐧E|K)={|E||K|Lk(12λK,𝐀±)on K=K𝐳±,i=1,2,0otherwise.\operatorname*{div}\left(\left.B_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}\right|_{K}\right)=\left\{\begin{array}[c]{ll}\mp\frac{\left|E\right|}{\left|K\right|}L_{k}^{\prime}\left(1-2\lambda_{K,\mathbf{A}^{\pm}}\right)&\text{on }K=K_{\mathbf{z}}^{\pm},\quad i=1,2,\\ 0&\text{otherwise.}\end{array}\right. (2.33)

Well-known properties of Legendre polynomials applied to (2.33) imply that for any vertex 𝐲\mathbf{y} of KK and odd polynomial degree kk

div(B~k,ECR𝐧E|K)(𝐲)=(k+12)×{|E||K|𝐲𝒱(K),if K=K𝐳±0otherwise.\operatorname*{div}\left(\left.\tilde{B}_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}\right|_{K}\right)\left(\mathbf{y}\right)=\mp\binom{k+1}{2}\times\left\{\begin{array}[c]{ll}\frac{\left|E\right|}{\left|K\right|}&\forall\mathbf{y}\in\mathcal{V}\left(K\right),\quad\text{if }K=K_{\mathbf{z}}^{\pm}\\ 0&\text{otherwise.}\end{array}\right. (2.34)

Hence the condition A𝒯,𝐲(qdiv𝒯𝐯)=0A_{\mathcal{T},\mathbf{y}}\left(q-\operatorname*{div}_{\mathcal{T}}\mathbf{v}_{\ell}\right)=0 for all 𝐲𝒞𝒯,(η)\mathbf{y}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right) is equivalent to the system of linear equation

𝐌𝜶=𝐫\mathbf{M}_{\ell}\mbox{\boldmath$\alpha$}_{\ell}=\mathbf{r}_{\ell} (2.35)

with

𝐌:=(A𝒯,𝐲(div𝒯(Bk,𝔈(𝐳)CR𝔑(𝐳))))𝐲𝒞𝒯,(η)𝐳𝒞𝒯,(η),𝜶:=(α,𝐳)𝐳𝒞𝒯,(η),𝐫:=(A𝒯,𝐲(q))𝐲𝒞𝒯,(η).\mathbf{M}_{\ell}:=\left(A_{\mathcal{T},\mathbf{y}}\left(\operatorname*{div}\nolimits_{\mathcal{T}}\left(B_{k,\mathfrak{E}\left(\mathbf{z}\right)}^{\operatorname*{CR}}\mathfrak{N}\left(\mathbf{z}\right)\right)\right)\right)_{\begin{subarray}{c}\mathbf{y}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)\\ \mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)\end{subarray}},\quad\mbox{\boldmath$\alpha$}_{\ell}:=\left(\alpha_{\ell,\mathbf{z}}\right)_{\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)},\quad\mathbf{r}_{\ell}:=\left(A_{\mathcal{T},\mathbf{y}}\left(q\right)\right)_{\mathbf{y}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)}. (2.36)

The matrix 𝐌\mathbf{M}_{\ell} is explicitly given by

𝐌:=(k+12)[|E,1||K,1|+|E,1||K,2||E,2||K,2|00|E,1||K,2||E,2||K,2|+|E,2||K,3|00|E,n||K,n|00|E,n1||K,n||E,n||K,n|+|E,n||K,n+1|].\mathbf{M}_{\ell}:=-\binom{k+1}{2}\left[\begin{array}[c]{ccccc}\frac{\left|E_{\ell,1}\right|}{\left|K_{\ell,1}\right|}+\frac{\left|E_{\ell,1}\right|}{\left|K_{\ell,2}\right|}&\frac{\left|E_{\ell,2}\right|}{\left|K_{\ell,2}\right|}&0&\ldots&0\\ \frac{\left|E_{\ell,1}\right|}{\left|K_{\ell,2}\right|}&\frac{\left|E_{\ell,2}\right|}{\left|K_{\ell,2}\right|}+\frac{\left|E_{\ell,2}\right|}{\left|K_{\ell,3}\right|}&\ddots&\ddots&\vdots\\ 0&\ddots&\ddots&&0\\ \vdots&\ddots&&&\frac{\left|E_{\ell,n_{\ell}}\right|}{\left|K_{\ell,n_{\ell}}\right|}\\ 0&\ldots&0&\frac{\left|E_{\ell,n_{\ell}-1}\right|}{\left|K_{\ell,n_{\ell}}\right|}&\frac{\left|E_{\ell,n_{\ell}}\right|}{\left|K_{\ell,n_{\ell}}\right|}+\frac{\left|E_{\ell,n_{\ell}}\right|}{\left|K_{\ell,n_{\ell+1}}\right|}\end{array}\right].

We use (cf. Fig. 5)

Refer to caption
Figure 5: Local numbering convention of the angles in K,jK_{\ell,j} and K,j+1K_{\ell,j+1}. The angle in K,jK_{\ell,j} at 𝐳\mathbf{z}_{\ell} is denoted by α,j,1\alpha_{\ell,j,1}, at 𝐳,j\mathbf{z}_{\ell,j} by α,j,2\alpha_{\ell,j,2}, at 𝐳,j1\mathbf{z}_{\ell,j-1} by α,j,3\alpha_{\ell,j,3} and in K,j+1K_{\ell,j+1} accordingly.
|E,j||K,j|=2sin(α,j,1+α,j,2)|E,j|sinα,j,1sinα,j,2and |E,j||K,j+1|=2sin(α,j+1,1+α,j+1,3)|E,j|sinα,j+1,1sinα,j+1,3\frac{\left|E_{\ell,j}\right|}{\left|K_{\ell,j}\right|}=\frac{2\sin\left(\alpha_{\ell,j,1}+\alpha_{\ell,j,2}\right)}{\left|E_{\ell,j}\right|\sin\alpha_{\ell,j,1}\sin\alpha_{\ell,j,2}}\quad\text{and\quad}\frac{\left|E_{\ell,j}\right|}{\left|K_{\ell,j+1}\right|}=\frac{2\sin\left(\alpha_{\ell,j+1,1}+\alpha_{\ell,j+1,3}\right)}{\left|E_{\ell,j}\right|\sin\alpha_{\ell,j+1,1}\sin\alpha_{\ell,j+1,3}}

(cf. [10, formula before (3.29)]) and obtain

𝐌=𝐃(𝐓+𝚫)\mathbf{M}_{\ell}=\mathbf{D}_{\ell}\left(\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell}\right) (2.37)

with 𝐃=k(k+1)diag[|E,j|1:1jn]\mathbf{D}_{\ell}=-k\left(k+1\right)\operatorname*{diag}\left[\left|E_{\ell,j}\right|^{-1}:1\leq j\leq n_{\ell}\right] and

𝐓\displaystyle\mathbf{T}_{\ell} :=[sin(α,1,1+α,2,1)sinα,1,1sinα,2,11sinα,2,1001sinα,2,1sin(α,2,1+α,3,1)sinα,2,1sinα,3,1001sinα,n,1001sinα,n,1sin(α,n,1+α,n+1,1)sinα,n,1sinα,n+1,1],\displaystyle:=\left[\begin{array}[c]{ccccc}\frac{\sin\left(\alpha_{\ell,1,1}+\alpha_{\ell,2,1}\right)}{\sin\alpha_{\ell,1,1}\sin\alpha_{\ell,2,1}}&\frac{1}{\sin\alpha_{\ell,2,1}}&0&\ldots&0\\ \frac{1}{\sin\alpha_{\ell,2,1}}&\frac{\sin\left(\alpha_{\ell,2,1}+\alpha_{\ell,3,1}\right)}{\sin\alpha_{\ell,2,1}\sin\alpha_{\ell,3,1}}&\ddots&\ddots&\vdots\\ 0&\ddots&\ddots&&0\\ \vdots&\ddots&&&\frac{1}{\sin\alpha_{\ell,n_{\ell},1}}\\ 0&\ldots&0&\frac{1}{\sin\alpha_{\ell,n_{\ell},1}}&\frac{\sin\left(\alpha_{\ell,n_{\ell},1}+\alpha_{\ell,n_{\ell}+1,1}\right)}{\sin\alpha_{\ell,n_{\ell},1}\sin\alpha_{\ell,n_{\ell}+1,1}}\end{array}\right], (2.43)
𝚫\displaystyle\mbox{\boldmath$\Delta$}_{\ell} :=diag[sin(α,j,2+α,j+1,3)sinα,j,2sinα,j+1,3:1jn].\displaystyle:=\operatorname*{diag}\left[\frac{\sin\left(\alpha_{\ell,j,2}+\alpha_{\ell,j+1,3}\right)}{\sin\alpha_{\ell,j,2}\sin\alpha_{\ell,j+1,3}}:1\leq j\leq n_{\ell}\right].

In Lemma A.1, we will prove that the matrix 𝐓+𝚫\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell} is invertible and the inverse is bounded by a constant independent of h𝒯h_{\mathcal{T}} and kk. Hence,

𝜶C~h𝐳k(k+1)𝐫.\left\|\mbox{\boldmath$\alpha$}_{\ell}\right\|\leq\tilde{C}\frac{h_{\mathbf{z}_{\ell}}}{k\left(k+1\right)}\left\|\mathbf{r}_{\ell}\right\|. (2.44)

Let 𝒯:={K,j:1jn+1}\mathcal{T}_{\ell}:=\left\{K_{\ell,j}:1\leq j\leq n_{\ell}+1\right\} and D:=dom(𝒯)D_{\ell}:=\operatorname*{dom}\left(\mathcal{T}_{\ell}\right). We estimate the function 𝐯\mathbf{v}_{\ell} in (2.32) by

𝒯𝐯𝕃2(D)\displaystyle\left\|\nabla_{\mathcal{T}}\mathbf{v}_{\ell}\right\|_{\mathbb{L}^{2}\left(D_{\ell}\right)} (𝐳𝒞𝒯,(η)|α,𝐳|2𝒯B~k,𝔈(𝐳)CR𝔑(𝐳)𝕃2(D)2)1/2\displaystyle\leq\left(\sum_{\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)}\left|\alpha_{\ell,\mathbf{z}}\right|^{2}\left\|\nabla_{\mathcal{T}}\tilde{B}_{k,\mathfrak{E}\left(\mathbf{z}\right)}^{\operatorname*{CR}}\mathfrak{N}\left(\mathbf{z}\right)\right\|_{\mathbb{L}^{2}\left(D_{\ell}\right)}^{2}\right)^{1/2}
max𝐳𝒞𝒯,(η)𝒯B~k,𝔈(𝐳)CR𝔑(𝐳)𝕃2(D)𝜶\displaystyle\leq\max_{\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)}\left\|\nabla_{\mathcal{T}}\tilde{B}_{k,\mathfrak{E}\left(\mathbf{z}\right)}^{\operatorname*{CR}}\mathfrak{N}\left(\mathbf{z}\right)\right\|_{\mathbb{L}^{2}\left(D_{\ell}\right)}\left\|\mbox{\boldmath$\alpha$}_{\ell}\right\|
(2.19)Ch𝐳log(k+1)(k+1)2𝐫.\displaystyle\overset{\text{(\ref{pwgrad})}}{\leq}Ch_{\mathbf{z}_{\ell}}\frac{\sqrt{\log\left(k+1\right)}}{\left(k+1\right)^{2}}\left\|\mathbf{r}_{\ell}\right\|.

The constant CC only depends on the shape-regularity of the mesh. We use Lemma 2.19 and conclude that

𝒯𝐯𝕃2(D)Clog(k+1)qL2(D).\left\|\nabla_{\mathcal{T}}\mathbf{v}_{\ell}\right\|_{\mathbb{L}^{2}\left(D_{\ell}\right)}\leq C\sqrt{\log\left(k+1\right)}\left\|q\right\|_{L^{2}\left(D_{\ell}\right)}\text{.}

Since the interiors of the supports DD_{\ell} have pairwise empty intersection the estimate

𝒯𝐯q𝕃2(Ω)Clog(k+1)qL2(Ω)\left\|\nabla_{\mathcal{T}}\mathbf{v}_{q}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}\leq C\sqrt{\log\left(k+1\right)}\left\|q\right\|_{L^{2}\left(\Omega\right)}

follows.  

Definition 2.21

Let assumption (2.7) be satisfied and η2>0\eta_{2}>0 as in Lemma 2.20. Fix η[0,η2[\eta\in\left[0,\eta_{2}\right[. For qk,0(𝒯)q\in\mathbb{P}_{k,0}\left(\mathcal{T}\right), the linear map ΠkCR:k1,0(𝒯)𝐂𝐑k,0(𝒯)\Pi_{k}^{\operatorname*{CR}}:\mathbb{P}_{k-1,0}\left(\mathcal{T}\right)\rightarrow\mathbf{CR}_{k,0}\left(\mathcal{T}\right) is given by

ΠkCRq:=𝒥𝐳𝒞𝒯,(η)α,𝐳B~k,𝔈(𝐳)CR𝔑(𝐳)\Pi_{k}^{\operatorname*{CR}}q:=\sum_{\ell\in\mathcal{J}}\sum_{\mathbf{z}\in\mathcal{C}_{\mathcal{T},\ell}\left(\eta\right)}\alpha_{\ell,\mathbf{z}}\tilde{B}_{k,\mathfrak{E}\left(\mathbf{z}\right)}^{\operatorname*{CR}}\mathfrak{N}\left(\mathbf{z}\right)

with 𝛂\mbox{\boldmath$\alpha$}_{\ell} as in (2.35).

For the proof of the following lemma we recall the definitions of some linear maps from the literature which are related to the right inverse of the divergence operator acting on some polynomial spaces.

Bernardi and Raugel introduced in [8, Lem. II.4] a linear mapping ΠBR:Mk1(𝒯)𝐒2,0(𝒯)\Pi^{\operatorname*{BR}}:M_{k-1}\left(\mathcal{T}\right)\rightarrow\mathbf{S}_{2,0}\left(\mathcal{T}\right) with the property: for any qMk1(𝒯)q\in M_{k-1}\left(\mathcal{T}\right), the function ΠBRq𝐒2,0(𝒯)\Pi^{\operatorname*{BR}}q\in\mathbf{S}_{2,0}\left(\mathcal{T}\right) satisfies

Kq=Kdiv(ΠBRq)K𝒯\int_{K}q=\int_{K}\operatorname*{div}\left(\Pi^{\operatorname*{BR}}q\right)\mathbf{\quad}\forall K\in\mathcal{T} (2.45)

and

ΠBRq𝐇1(Ω)CBRqL2(Ω)\left\|\Pi^{\operatorname*{BR}}q\right\|_{\mathbf{H}^{1}\left(\Omega\right)}\leq C_{\operatorname*{BR}}\left\|q\right\|_{L^{2}\left(\Omega\right)} (2.46)

for a constant CBRC_{\operatorname*{BR}} which is independent of the mesh width and the polynomial degree.

Next we consider some right inverse of the divergence operator on the space

Mk1V(𝒯):={qMk1(𝒯)(Kq=0K𝒯q|K(𝐲)=0𝐲𝒱(K))}.M_{k-1}^{\operatorname*{V}}\left(\mathcal{T}\right):=\left\{q\in M_{k-1}\left(\mathcal{T}\right)\mid\left(\begin{array}[c]{ll}\int_{K}q=0&\forall K\in\mathcal{T}\\ \left.q\right|_{K}\left(\mathbf{y}\right)=0&\forall\mathbf{y}\in\mathcal{V}\left(K\right)\end{array}\right)\right\}.

Let

𝐒kV(𝒯):={𝐮𝐒k(𝒯)K𝒯:𝐮|K=𝟎}.\mathbf{S}_{k}^{\operatorname*{V}}\left(\mathcal{T}\right):=\left\{\mathbf{u}\in\mathbf{S}_{k}\left(\mathcal{T}\right)\mid\forall K\in\mathcal{T}:\quad\left.\mathbf{u}\right|_{\partial K}=\mathbf{0}\right\}.

There exists a linear operator ΠV:Mk1V(𝒯)𝐒kV(𝒯)\Pi^{\operatorname*{V}}:M_{k-1}^{\operatorname*{V}}\left(\mathcal{T}\right)\rightarrow\mathbf{S}_{k}^{\operatorname*{V}}\left(\mathcal{T}\right) such that for all qMk1V(𝒯)q\in M_{k-1}^{\operatorname*{V}}\left(\mathcal{T}\right)

q\displaystyle q =divΠVq,\displaystyle=\operatorname*{div}\Pi^{\operatorname*{V}}q,
ΠVq𝐇1(Ω)\displaystyle\left\|\Pi^{\operatorname*{V}}q\right\|_{\mathbf{H}^{1}\left(\Omega\right)} CVqL2(Ω),\displaystyle\leq C_{\operatorname*{V}}\left\|q\right\|_{L^{2}\left(\Omega\right)}, (2.47)

where the constant CVC_{\operatorname*{V}} is independent of the mesh width and the polynomial degree. Note that in the original paper [37, Lem. 2.5] by M. Vogelius, the right-hand side in the estimate (2.47) contains an additional factor kβVk^{\beta_{\operatorname*{V}}} for some positive βV\beta_{\operatorname*{V}} (independent of the mesh width). In [3, Thm. 3.4], the operator in [37, Lem. 2.5] is modified and the estimate in the form (2.47) is proved for the modified operator.

Finally, we reconsider the linear operator ΠGS:Mη,k1SV(𝒯)𝐒4,0(𝒯)\Pi^{\operatorname*{GS}}:M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right)\rightarrow\mathbf{S}_{4,0}\left(\mathcal{T}\right) introduced by Guzmán and Scott in [27, Proof of Lem. 6 and Lem. 7] with the property that, for any qMη,k1SV(𝒯)q\in M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right), it holds

(IdivΠGS)q\displaystyle\left(I-\operatorname*{div}\Pi^{\operatorname*{GS}}\right)q Mk1V(𝒯),\displaystyle\in M_{k-1}^{\operatorname*{V}}\left(\mathcal{T}\right), (2.48a)
ΠGSq𝕃2(Ω)\displaystyle\left\|\nabla\Pi^{\operatorname*{GS}}q\right\|_{\mathbb{L}^{2}\left(\Omega\right)} CGSkκ(θmin+η)1qL2(Ω)for κ=2.\displaystyle\leq C_{\operatorname*{GS}}k^{\kappa}\left(\theta_{\min}+\eta\right)^{-1}\left\|q\right\|_{L^{2}\left(\Omega\right)}\quad\text{for }\kappa=2. (2.48b)

We emphasize that in [27, Lemma 7] the constant Θmin1\Theta_{\min}^{-1} (cf. (2.27)) instead of (Θmin+η)1\left(\Theta_{\min}+\eta\right)^{-1} appears in (2.48b) so that the estimate of ΠGSq𝕃2(Ω)\left\|\nabla\Pi^{\operatorname*{GS}}q\right\|_{\mathbb{L}^{2}\left(\Omega\right)} for qM0,k1SV(𝒯)q\in M_{0,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) deteriorates in cases where the 𝐳\mathbf{z} is a nearly critical point, i.e., very close to the geometric situations described in Remark 2.7. The proof of [27, Lemma 7] is split into an estimate related to points 𝐳\mathbf{z} with A𝒯,𝐳(q)=0A_{\mathcal{T},\mathbf{z}}\left(q\right)=0 (cf. (2.26)) and an estimate for the remaining points 𝐳\mathbf{z} with A𝒯,𝐳(q)0A_{\mathcal{T},\mathbf{z}}\left(q\right)\neq 0. Only in this second part, the constant Θmin1\Theta_{\min}^{-1} is involved. The result has been improved in [24, Lem. 4.5] and it was shown that there is an operator Πη,k1:Mη,k1SV(𝒯)𝐒k,0(𝒯)\Pi_{\eta,k-1}:M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right)\rightarrow\mathbf{S}_{k,0}\left(\mathcal{T}\right) such that the properties in (2.48) hold for κ=0\kappa=0: for any qMη,k1SV(𝒯)q\in M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) it holds

(IdivΠη,k1)q\displaystyle\left(I-\operatorname*{div}\Pi_{\eta,k-1}\right)q Mk1V(𝒯),\displaystyle\in M_{k-1}^{\operatorname*{V}}\left(\mathcal{T}\right),
Πη,k1q𝕃2(Ω)\displaystyle\left\|\nabla\Pi_{\eta,k-1}q\right\|_{\mathbb{L}^{2}\left(\Omega\right)} Cπ(Θmin+η)1qL2(Ω).\displaystyle\leq C_{\pi}\left(\Theta_{\min}+\eta\right)^{-1}\left\|q\right\|_{L^{2}\left(\Omega\right)}.

From Lemma 2.20, we conclude that qdiv𝒯(ΠkCRq)Mη,k1SV(𝒯)q-\operatorname*{div}_{\mathcal{T}}\left(\Pi_{k}^{\operatorname*{CR}}q\right)\in M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) and the second part of the proof in [27, Lemma 7] is applied only to points with

min𝐳𝒱(𝒯)\𝒞𝒯(η)Θ(𝐳)max{η,Θmin}.\min_{\mathbf{z}\in\mathcal{V}\left(\mathcal{T}\right)\backslash\mathcal{C}_{\mathcal{T}}\left(\eta\right)}\Theta\left(\mathbf{z}\right)\geq\max\left\{\eta,\Theta_{\min}\right\}.

Hence, (2.48b) follows for η\eta depending only on the shape-regularity of the mesh.

Lemma 2.22

Let assumption (2.7) be satisfied and let 𝒞𝒯acute=\mathcal{C}_{\mathcal{T}}^{\operatorname{acute}}=\emptyset. There exists a constant η2>0\eta_{2}>0 which only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} as in (1.1) such that for any fixed 0η<η20\leq\eta<\eta_{2} and any qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right), there exists some 𝐰q𝐂𝐑k,0(𝒯)\mathbf{w}_{q}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right) such that

q=div𝒯𝐰qq=\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{w}_{q}

and

𝐰q𝐇1(𝒯)Clog(k+1)(Θmin+η)1qL2(Ω).\left\|\mathbf{w}_{q}\right\|_{\mathbf{H}^{1}\left(\mathcal{T}\right)}\leq C\sqrt{\log\left(k+1\right)}\left(\Theta_{\min}+\eta\right)^{-1}\left\|q\right\|_{L^{2}\left(\Omega\right)}.

The constant CC only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} but is independent of the mesh width and the polynomial degree kk.

Proof. For the construction of 𝐰q\mathbf{w}_{q} we follow and modify the lines of proof in [27, Thm. 1] by a) involving the operator ΠkCR\Pi_{k}^{\operatorname*{CR}} and b) employing the concept of η\eta-critical points.

For given qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right), we employ the operators ΠkCR\Pi_{k}^{\operatorname*{CR}}, ΠBR\Pi^{\operatorname*{BR}}, Πη,k1\Pi_{\eta,k-1}, ΠV\Pi^{\operatorname*{V}} in the definition of the function 𝐰q\mathbf{w}_{q}

𝐰q\displaystyle\mathbf{w}_{q} =𝐓1+𝐓2+𝐓3+𝐓4,\displaystyle=\mathbf{T}_{1}+\mathbf{T}_{2}+\mathbf{T}_{3}+\mathbf{T}_{4}, (2.49)
𝐓1\displaystyle\mathbf{T}_{1} :=ΠBRq,\displaystyle:=\Pi^{\operatorname*{BR}}q,
𝐓2\displaystyle\mathbf{T}_{2} :=ΠkCR(IdivΠBR)q,\displaystyle:=\Pi_{k}^{\operatorname*{CR}}\left(I-\operatorname*{div}\Pi^{\operatorname*{BR}}\right)q,
𝐓3\displaystyle\mathbf{T}_{3} :=Πη,k1(Idiv𝒯ΠkCR)(IdivΠBR)q,\displaystyle:=\Pi_{\eta,k-1}\left(I-\operatorname*{div}\nolimits_{\mathcal{T}}\Pi_{k}^{\operatorname*{CR}}\right)\left(I-\operatorname*{div}\Pi^{\operatorname*{BR}}\right)q,
𝐓4\displaystyle\mathbf{T}_{4} :=ΠV(IdivΠη,k1)(Idiv𝒯ΠkCR)(IdivΠBR)q.\displaystyle:=\Pi^{\operatorname*{V}}\left(I-\operatorname*{div}\Pi_{\eta,k-1}\right)\left(I-\operatorname*{div}\nolimits_{\mathcal{T}}\Pi_{k}^{\operatorname*{CR}}\right)\left(I-\operatorname*{div}\Pi^{\operatorname*{BR}}\right)q.

By construction we have

div𝒯𝐰q=q.\operatorname{div}_{\mathcal{T}}\mathbf{w}_{q}=q.

The first two summands in (2.49) satisfy

𝐓1𝐇1(Ω)\displaystyle\left\|\mathbf{T}_{1}\right\|_{\mathbf{H}^{1}\left(\Omega\right)} (2.46)CBRqL2(Ω),\displaystyle\overset{\text{(\ref{CBR})}}{\leq}C_{\operatorname*{BR}}\left\|q\right\|_{L^{2}\left(\Omega\right)}, (2.50)
𝐓2𝐇1(𝒯)\displaystyle\left\|\mathbf{T}_{2}\right\|_{\mathbf{H}^{1}\left(\mathcal{T}\right)} (2.31)CCRlog(k+1)(qL2(Ω)+ΠBRq𝐇1(Ω))\displaystyle\overset{\text{(\ref{CRlifting1})}}{\leq}C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\left(\left\|q\right\|_{L^{2}\left(\Omega\right)}+\left\|\Pi^{\operatorname*{BR}}q\right\|_{\mathbf{H}^{1}\left(\Omega\right)}\right) (2.51)
(2.50)CCRlog(k+1)(1+CBR)qL2(Ω).\displaystyle\overset{\text{(\ref{estPi1})}}{\leq}C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\left(1+C_{\operatorname*{BR}}\right)\left\|q\right\|_{L^{2}\left(\Omega\right)}.

For the third term in (2.49) we get

𝐓3𝕃2(Ω)\displaystyle\left\|\nabla\mathbf{T}_{3}\right\|_{\mathbb{L}^{2}\left(\Omega\right)} Cπ(Θmin+η)1(Idiv𝒯ΠkCR)(IdivΠBR)qL2(Ω)\displaystyle\leq C_{\pi}\left(\Theta_{\min}+\eta\right)^{-1}\left\|\left(I-\operatorname*{div}\nolimits_{\mathcal{T}}\Pi_{k}^{\operatorname*{CR}}\right)\left(I-\operatorname*{div}\Pi^{\operatorname*{BR}}\right)q\right\|_{L^{2}\left(\Omega\right)}
Cπ(Θmin+η)1((IdivΠBR)qL2(Ω)+div𝒯ΠkCR(IdivΠBR)qL2(Ω)).\displaystyle\leq C_{\pi}\left(\Theta_{\min}+\eta\right)^{-1}\left(\left\|\left(I-\operatorname*{div}\Pi^{\operatorname*{BR}}\right)q\right\|_{L^{2}\left(\Omega\right)}+\left\|\operatorname*{div}\nolimits_{\mathcal{T}}\Pi_{k}^{\operatorname*{CR}}\left(I-\operatorname*{div}\Pi^{\operatorname*{BR}}\right)q\right\|_{L^{2}\left(\Omega\right)}\right). (2.52)

The combination with (2.46), (2.31) leads to

𝐓3𝕃2(Ω)CCπ(1+CBR)(1+CCR)log(k+1)Θmin+ηqL2(Ω).\left\|\nabla\mathbf{T}_{3}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}\leq CC_{\pi}\left(1+C_{\operatorname*{BR}}\right)\left(1+C_{\operatorname*{CR}}\right)\frac{\sqrt{\log\left(k+1\right)}}{\Theta_{\min}+\eta}\left\|q\right\|_{L^{2}\left(\Omega\right)}. (2.53)

For the fourth term we get in a similar way

𝐓4𝕃2(Ω)\displaystyle\left\|\nabla\mathbf{T}_{4}\right\|_{\mathbb{L}^{2}\left(\Omega\right)} (2.47)CV((Idiv𝒯ΠkCR)(IdivΠBR)qL2(Ω)+div𝐓3L2(Ω))\displaystyle\overset{\text{(\ref{optests})}}{\leq}C_{\operatorname*{V}}\left(\left\|\left(I-\operatorname*{div}\nolimits_{\mathcal{T}}\Pi_{k}^{\operatorname*{CR}}\right)\left(I-\operatorname*{div}\Pi^{\operatorname*{BR}}\right)q\right\|_{L^{2}\left(\Omega\right)}+\left\|\operatorname*{div}\mathbf{T}_{3}\right\|_{L^{2}\left(\Omega\right)}\right)
CV(qL2(Ω)+div𝐓1L2(Ω)+div𝒯𝐓2L2(Ω)+div𝐓3L2(Ω))\displaystyle\leq C_{\operatorname*{V}}\left(\left\|q\right\|_{L^{2}\left(\Omega\right)}+\left\|\operatorname{div}\mathbf{T}_{1}\right\|_{L^{2}\left(\Omega\right)}+\left\|\operatorname{div}_{\mathcal{T}}\mathbf{T}_{2}\right\|_{L^{2}\left(\Omega\right)}+\left\|\operatorname*{div}\mathbf{T}_{3}\right\|_{L^{2}\left(\Omega\right)}\right)
(2.50), (2.51), (2.53)CV(1+CBR)(2+CCR)log(k+1)(1+CCπ(Θmin+η)1)qL2(Ω).\displaystyle\overset{\text{(\ref{estPi1}), (\ref{estPi2}), (\ref{estPi3})}}{\leq}C_{\operatorname*{V}}\left(1+C_{\operatorname*{BR}}\right)\left(2+C_{\operatorname*{CR}}\right)\sqrt{\log\left(k+1\right)}\left(1+CC_{\pi}\left(\Theta_{\min}+\eta\right)^{-1}\right)\left\|q\right\|_{L^{2}\left(\Omega\right)}. (2.54)

The combination of (2.50), (2.51), (2.53), (2.54) with (2.49) leads to the assertion.   

Lemma 2.22 implies that for conforming triangulations 𝒯\mathcal{T} which satisfy (2.7) and 𝒞𝒯acute=\mathcal{C}_{\mathcal{T}}^{\operatorname{acute}}=\emptyset, there exists a bounded linear operator

Π𝒯,kinv:k1,0(𝒯)𝐂𝐑k,0(𝒯)\Pi_{\mathcal{T},k}^{\operatorname*{inv}}:\mathbb{P}_{k-1,0}\left(\mathcal{T}\right)\rightarrow\mathbf{CR}_{k,0}\left(\mathcal{T}\right)

such that div𝒯Π𝒯,kinv\operatorname*{div}_{\mathcal{T}}\circ\Pi_{\mathcal{T},k}^{\operatorname*{inv}} is the identity on k1,0(𝒯)\mathbb{P}_{k-1,0}\left(\mathcal{T}\right) and

Π𝒯,kinvq𝐇1(𝒯)Cinvlog(k+1)qL2(Ω)\left\|\Pi_{\mathcal{T},k}^{\operatorname*{inv}}q\right\|_{\mathbf{H}^{1}\left(\mathcal{T}\right)}\leq C_{\operatorname*{inv}}\sqrt{\log\left(k+1\right)}\left\|q\right\|_{L^{2}\left(\Omega\right)}

for a constant CinvC_{\operatorname*{inv}} which only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} (cf. (1.1)).

2.2.3 The case 𝒞𝒯acute(η)\mathcal{C}_{\mathcal{T}}^{\operatorname*{acute}}\left(\eta\right)\neq\emptyset

In this section, we remove the condition 𝒞𝒯acute=\mathcal{C}_{\mathcal{T}}^{\operatorname{acute}}=\emptyset and construct a bounded right-inverse of the piecewise divergence operator for odd k5k\geq 5 and conforming triangulations which contain at least one inner point. The construction is based on the step-by-step procedure (cf. (2.13)) from the triangulation 𝒯1\mathcal{T}_{1} to 𝒯\mathcal{T}. Inductively, we assume that there is a triangulation 𝒯j\mathcal{T}_{j} along a bounded right-inverse Πj,kinv:k1,0(𝒯j)𝐂𝐑k,0(𝒯j)\Pi_{j,k}^{\operatorname*{inv}}:\mathbb{P}_{k-1,0}\left(\mathcal{T}_{j}\right)\rightarrow\mathbf{CR}_{k,0}\left(\mathcal{T}_{j}\right) of the piecewise divergence operator. A single extension step is analysed by the following lemma.

Lemma 2.23

Let 𝒯\mathcal{T} denote a conforming triangulation for the domain Ω:=dom𝒯\Omega:=\operatorname*{dom}\mathcal{T} and let 𝒯𝒯\mathcal{T}^{\prime}\subset\mathcal{T} be a subset such that every triangle K𝒯\𝒯K\in\mathcal{T}\backslash\mathcal{T}^{\prime} has one edge, say EE, which belongs to (𝒯)\mathcal{E}\left(\mathcal{T}^{\prime}\right). We assume that 𝒯\mathcal{T}^{\prime} has at least one inner vertex and set Ω:=dom𝒯\Omega^{\prime}:=\operatorname*{dom}\mathcal{T}^{\prime}. Assume that there exists a bounded linear operator Π𝒯,kinv:k1,0(𝒯)𝐂𝐑k,0(𝒯)\Pi_{\mathcal{T}^{\prime},k}^{\operatorname*{inv}}:\mathbb{P}_{k-1,0}\left(\mathcal{T}^{\prime}\right)\rightarrow\mathbf{CR}_{k,0}\left(\mathcal{T}^{\prime}\right) with div𝒯Π𝒯,kinv=Id\operatorname*{div}_{\mathcal{T}^{\prime}}\circ\Pi_{\mathcal{T}^{\prime},k}^{\operatorname*{inv}}=\operatorname*{Id} on k1,0(𝒯)\mathbb{P}_{k-1,0}\left(\mathcal{T}^{\prime}\right) and

Π𝒯,kinvq𝐇1(Ω)C𝒯qL2(Ω).\left\|\Pi_{\mathcal{T}^{\prime},k}^{\operatorname*{inv}}q\right\|_{\mathbf{H}^{1}\left(\Omega^{\prime}\right)}\leq C_{\mathcal{T}^{\prime}}\left\|q\right\|_{L^{2}\left(\Omega^{\prime}\right)}.

Then, there exists a linear operator Π𝒯,kinv:k1,0(𝒯)𝐂𝐑k,0(𝒯)\Pi_{\mathcal{T},k}^{\operatorname*{inv}}:\mathbb{P}_{k-1,0}\left(\mathcal{T}\right)\rightarrow\mathbf{CR}_{k,0}\left(\mathcal{T}\right) with div𝒯Π𝒯,kinv=Id\operatorname*{div}_{\mathcal{T}}\circ\Pi_{\mathcal{T},k}^{\operatorname*{inv}}=\operatorname*{Id} on k1,0(𝒯)\mathbb{P}_{k-1,0}\left(\mathcal{T}\right) and

Π𝒯,kinvq𝐇1(Ω)C𝒯qL2(Ω)with C𝒯:=C3log(k+1)C𝒯\left\|\Pi_{\mathcal{T},k}^{\operatorname*{inv}}q\right\|_{\mathbf{H}^{1}\left(\Omega\right)}\leq C_{\mathcal{T}}\left\|q\right\|_{L^{2}\left(\Omega\right)}\quad\text{with\quad}C_{\mathcal{T}}:=C_{3}\sqrt{\log\left(k+1\right)}C_{\mathcal{T}^{\prime}}

for a constant C3C_{3} which depends only on the shape-regularity of the mesh and on αΩ\alpha_{\Omega}.

Proof. Let qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right). We set 𝐯0:=ΠBRq\mathbf{v}_{0}:=\Pi^{\operatorname*{BR}}q where the operator ΠBR\Pi^{\operatorname*{BR}} is as in (2.45) and satisfies

𝐯0𝐇1(Ω)CBRqL2(Ω).\left\|\mathbf{v}_{0}\right\|_{\mathbf{H}^{1}\left(\Omega\right)}\leq C_{\operatorname*{BR}}\left\|q\right\|_{L^{2}\left(\Omega\right)}.

Hence, q1:=qdiv𝐯0q_{1}:=q-\operatorname*{div}\mathbf{v}_{0} belongs to k1,0(𝒯)\mathbb{P}_{k-1,0}\left(\mathcal{T}\right) and has trianglewise integral mean zero.

The following construction is illustrated in Figure 6.

Refer to caption
Figure 6: The black triangles form the triangulation 𝒯\mathcal{T}^{\prime}. Left: One triangle KK is attached to 𝒯\mathcal{T}^{\prime} having a common side EE with K𝒯K^{\prime}\in\mathcal{T}^{\prime} and 𝒯out(K)={K}\mathcal{T}_{\operatorname*{out}}\left(K^{\prime}\right)=\left\{K\right\}. Right: Two triangles K1K_{1}, K2𝒯K_{2}\notin\mathcal{T}^{\prime} are attached to a triangle K𝒯K^{\prime}\in\mathcal{T}^{\prime} and 𝒯out(K)={K1,K2}\mathcal{T}_{\operatorname*{out}}\left(K^{\prime}\right)=\left\{K_{1},K_{2}\right\}.

Let K𝒯K^{\prime}\in\mathcal{T}^{\prime} be such that there exists a non-empty subset 𝒯out(K)𝒯\𝒯\mathcal{T}_{\operatorname*{out}}\left(K^{\prime}\right)\subset\mathcal{T}\backslash\mathcal{T}^{\prime} having the property that any K𝒯out(K)K\in\mathcal{T}_{\operatorname*{out}}\left(K^{\prime}\right) shares an edge with KK^{\prime}. We have |𝒯out(K)|2\left|\mathcal{T}_{\operatorname*{out}}\left(K^{\prime}\right)\right|\leq 2; indeed, if |𝒯out(K)|=3\left|\mathcal{T}_{\operatorname*{out}}\left(K^{\prime}\right)\right|=3, then all three edges of KK^{\prime} are boundary edges which implies 𝒯={K}\mathcal{T}^{\prime}=\left\{K^{\prime}\right\} and violates the condition that 𝒯\mathcal{T}^{\prime} must contain an inner vertex.

For K𝒯out(K)K\in\mathcal{T}_{\operatorname*{out}}\left(K^{\prime}\right), let 𝐳\mathbf{z} denote the vertex in KK opposite to EE and set ωE=KK\omega_{E}=K\cup K^{\prime}. The endpoints of EE are denoted by 𝐲1\mathbf{y}_{1}, 𝐲2\mathbf{y}_{2}. We employ the ansatz (cf. (2.24))

𝐯1:=αB~k,ECR𝐧E\mathbf{v}_{1}:=\alpha\tilde{B}_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E} (2.55)

with the convention that 𝐧E\mathbf{n}_{E} is the unit vector orthogonal to EE and directed into KK^{\prime}. By construction it holds 𝐯1𝐂𝐑k,0(𝒯)\mathbf{v}_{1}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right) and supp𝐯1ωE\operatorname*{supp}\mathbf{v}_{1}\subset\omega_{E}. We determine α\alpha in (2.55) such that div(𝐯1|K)(𝐳)=q1(𝐳)\operatorname*{div}\left(\left.\mathbf{v}_{1}\right|_{K}\right)\left(\mathbf{z}\right)=q_{1}\left(\mathbf{z}\right) and employ (2.34) to get

|div(B~k,ECR𝐧E)(𝐳)|=(k+12)|E||K|.\left|\operatorname{div}\left(\tilde{B}_{k,E}^{\operatorname*{CR}}\mathbf{n}_{E}\right)\left(\mathbf{z}\right)\right|=\binom{k+1}{2}\frac{\left|E\right|}{\left|K\right|}.

Hence |α|=|q1(𝐳)||K||E|/(k+12)\left|\alpha\right|=\left|q_{1}\left(\mathbf{z}\right)\right|\frac{\left|K\right|}{\left|E\right|}/\binom{k+1}{2} and we conclude as in the proof of Lemma 2.20 that

𝒯𝐯1𝕃2(ωE)CCRlog(k+1)q1L2(K).\left\|\nabla_{\mathcal{T}}\mathbf{v}_{1}\right\|_{\mathbb{L}^{2}\left(\omega_{E}\right)}\leq C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\left\|q_{1}\right\|_{L^{2}\left(K\right)}.

We set

q2:=q1div𝒯𝐯1so that q1=div𝒯𝐯1+q2q_{2}:=q_{1}-\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{v}_{1}\quad\text{so that }q_{1}=\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{v}_{1}+q_{2}

and note that

q2L2(K)(1+CCRlog(k+1))q1L2(K).\left\|q_{2}\right\|_{L^{2}\left(K\right)}\leq\left(1+C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\right)\left\|q_{1}\right\|_{L^{2}\left(K\right)}. (2.56)

The construction implies q2k1,0(𝒯)q_{2}\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right), q2q_{2} has trianglewise integral mean zero, and q2(𝐳)=0q_{2}\left(\mathbf{z}\right)=0. Next, we employ the vector field defined in [24, Lem. 4.9] which is a modification of the cubic vector field defined in [27, (3.5)] but allows for better kk-explicit estimates. We recall the relevant lemma from [24] for the existence of such vector fields and collect important properties.

Lemma 2.24 ([24, Lem. 4.9])

Let 𝒯\mathcal{T} be a conforming triangulation of Ω\Omega and let k3k\geq 3. Let E(𝒯)E\in\mathcal{E}\left(\mathcal{T}\right) with endpoints 𝐲1\mathbf{y}_{1}, 𝐲2\mathbf{y}_{2}. Then there exist vector fields 𝐯E,j\mathbf{v}_{E,j}, j{1,2}j\in\left\{1,2\right\}, with the following properties

𝐯E,j𝐒k(𝒯),supp𝐯E,jωE,Kdiv𝐯E,j=0K𝒯,j{1,2}(div𝐯E,j|K)(𝐯)={1if K𝒯(E)𝐯=𝐯j0otherwise,K𝒯,𝐯𝒱(K),j{1,2}𝐯E,j𝕃2(ωE)ChEk2.\begin{array}[c]{ll}\mathbf{v}_{E,j}\in\mathbf{S}_{k}\left(\mathcal{T}\right)\text{,}&\operatorname*{supp}\mathbf{v}_{E,j}\subset\omega_{E},\\ \int_{K}\operatorname*{div}\mathbf{v}_{E,j}=0&\forall K\in\mathcal{T},\quad j\in\left\{1,2\right\}\\ \left(\operatorname{div}\left.\mathbf{v}_{E,j}\right|_{K}\right)\left(\mathbf{v}\right)=\left\{\begin{array}[c]{cl}1&\text{if }K\in\mathcal{T}\left(E\right)\wedge\mathbf{v}=\mathbf{v}_{j}\\ 0&\text{otherwise,}\end{array}\right.&\forall K\in\mathcal{T},\quad\forall\mathbf{v}\in\mathcal{V}\left(K\right),\quad\forall j\in\left\{1,2\right\}\\ \left\|\nabla\mathbf{v}_{E,j}\right\|_{\mathbb{L}^{2}\left(\omega_{E}\right)}\leq Ch_{E}k^{-2}.&\end{array} (2.57)

We employ this vector field to the edge E=KKE=K\cap K^{\prime}, set

𝐯2:=j=12q2|K(𝐲j)𝐯E,j\mathbf{v}_{2}:=\sum_{j=1}^{2}\left.q_{2}\right|_{K}\left(\mathbf{y}_{j}\right)\mathbf{v}_{E,j}

and define

q3:=q2div𝐯2so that q2=div𝐯2+q3.q_{3}:=q_{2}-\operatorname{div}\mathbf{v}_{2}\quad\text{so that }q_{2}=\operatorname{div}\mathbf{v}_{2}+q_{3}. (2.58)

The function q3k1,0(𝒯)q_{3}\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right) has trianglewise integral mean zero and q3|K\left.q_{3}\right|_{K} vanishes in all vertices of KK. The norm 𝐯2𝕃2(KK)\left\|\nabla\mathbf{v}_{2}\right\|_{\mathbb{L}^{2}\left(K^{\prime}\cup K\right)} can be estimated in the same way as the function 𝐓3\mathbf{T}_{3} in (2.53); however the factor (Θmin+η)1\left(\Theta_{\min}+\eta\right)^{-1} does not appear as in (2.53) since the last estimate in (2.57) does not depend on these quantities. In this way, we get

𝒯𝐯2𝕃2(ωE)\displaystyle\left\|\nabla_{\mathcal{T}}\mathbf{v}_{2}\right\|_{\mathbb{L}^{2}\left(\omega_{E}\right)} C1q2L2(K)\displaystyle\leq C_{1}\left\|q_{2}\right\|_{L^{2}\left(K\right)} (2.59)
(2.56)C1(1+CCRlog(k+1))q1L2(K).\displaystyle\overset{\text{(\ref{q2est})}}{\leq}C_{1}\left(1+C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\right)\left\|q_{1}\right\|_{L^{2}\left(K\right)}. (2.60)

Hence, from [3, Thm. 3.4] we deduce that there exists 𝐯3𝐒k,0(𝒯)\mathbf{v}_{3}\in\mathbf{S}_{k,0}\left(\mathcal{T}\right) with supp𝐯3=K\operatorname*{supp}\mathbf{v}_{3}=K such that div𝐯3=q3\operatorname*{div}\mathbf{v}_{3}=q_{3} on KK and

hK1𝐯3𝐋2(K)+𝐯3𝕃2(K)\displaystyle h_{K}^{-1}\left\|\mathbf{v}_{3}\right\|_{\mathbf{L}^{2}\left(K\right)}+\left\|\nabla\mathbf{v}_{3}\right\|_{\mathbb{L}^{2}\left(K\right)}\leq CVq3L2(K)(2.58), (2.59)CV(1+C1)q2L2(K)\displaystyle C_{\operatorname*{V}}\left\|q_{3}\right\|_{L^{2}\left(K\right)}\overset{\text{(\ref{defq3}), (\ref{v2est1})}}{\leq}C_{\operatorname*{V}}\left(1+C_{1}\right)\left\|q_{2}\right\|_{L^{2}\left(K\right)}
(2.56)CV(1+C1)(1+CCRlog(k+1))q1L2(K).\displaystyle\overset{\text{(\ref{q2est})}}{\leq}C_{\operatorname*{V}}\left(1+C_{1}\right)\left(1+C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\right)\left\|q_{1}\right\|_{L^{2}\left(K\right)}.

In this way we have constructed the function 𝐯K𝐂𝐑k,0(𝒯)\mathbf{v}_{K}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right) by

𝐯K=𝐯1+𝐯2+𝐯3\mathbf{v}_{K}=\mathbf{v}_{1}+\mathbf{v}_{2}+\mathbf{v}_{3}

such that div𝐯K=q1\operatorname*{div}\mathbf{v}_{K}=q_{1} on KK, supp𝐯KωE\operatorname*{supp}\mathbf{v}_{K}\subset\omega_{E} and

𝒯𝐯K𝕃2(Ω)\displaystyle\left\|\nabla_{\mathcal{T}}\mathbf{v}_{K}\right\|_{\mathbb{L}^{2}\left(\Omega\right)} =𝒯𝐯K𝕃2(ωE)=13𝒯𝐯𝕃2(Ω)\displaystyle=\left\|\nabla_{\mathcal{T}}\mathbf{v}_{K}\right\|_{\mathbb{L}^{2}\left(\omega_{E}\right)}\leq\sum_{\ell=1}^{3}\left\|\nabla_{\mathcal{T}}\mathbf{v}_{\ell}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}
C2log(k+1)q1L2(K)C2log(k+1)(qL2(K)+𝐯0𝕃2(K)),\displaystyle\leq C_{2}\sqrt{\log\left(k+1\right)}\left\|q_{1}\right\|_{L^{2}\left(K\right)}\leq C_{2}\sqrt{\log\left(k+1\right)}\left(\left\|q\right\|_{L^{2}\left(K\right)}+\left\|\nabla\mathbf{v}_{0}\right\|_{\mathbb{L}^{2}\left(K\right)}\right),

where C2C_{2} only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} through the constants CVC_{\operatorname*{V}}, CCRC_{\operatorname*{CR}}, C1C_{1}. Let 𝐯q:=𝐯0+K𝒯\𝒯𝐯K\mathbf{v}_{q}:=\mathbf{v}_{0}+\sum_{K\in\mathcal{T}\backslash\mathcal{T}^{\prime}}\mathbf{v}_{K} and note that by construction

div𝒯𝐯q=qon Ω\Ω¯\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{v}_{q}=q\quad\text{on }\Omega\backslash\overline{\Omega^{\prime}}

and

𝒯𝐯q𝕃2(Ω)\displaystyle\left\|\nabla_{\mathcal{T}}\mathbf{v}_{q}\right\|_{\mathbb{L}^{2}\left(\Omega\right)} 𝐯0𝕃2(Ω)+K𝒯\𝒯𝒯𝐯K𝕃2(Ω)\displaystyle\leq\left\|\nabla\mathbf{v}_{0}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}+\sum_{K\in\mathcal{T}\backslash\mathcal{T}^{\prime}}\left\|\nabla_{\mathcal{T}}\mathbf{v}_{K}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}
𝐯0𝕃2(Ω)+2C2K𝒯\𝒯log(k+1)(qL2(K)+𝐯0𝕃2(K))\displaystyle\leq\left\|\nabla\mathbf{v}_{0}\right\|_{\mathbb{L}^{2}\left(\Omega\right)}+2C_{2}\sum_{K\in\mathcal{T}\backslash\mathcal{T}^{\prime}}\sqrt{\log\left(k+1\right)}\left(\left\|q\right\|_{L^{2}\left(K\right)}+\left\|\nabla\mathbf{v}_{0}\right\|_{\mathbb{L}^{2}\left(K\right)}\right)
Clog(k+1)qL2(Ω).\displaystyle\leq C\sqrt{\log\left(k+1\right)}\left\|q\right\|_{L^{2}\left(\Omega\right)}.

Finally, the linear map Π𝒯,kinv:k1,0(𝒯)𝐂𝐑k,0(𝒯)\Pi_{\mathcal{T},k}^{\operatorname*{inv}}:\mathbb{P}_{k-1,0}\left(\mathcal{T}\right)\rightarrow\mathbf{CR}_{k,0}\left(\mathcal{T}\right) is defined by

Π𝒯,kinvq=𝐯q+Π𝒯,kinv((qdiv𝐯q)|Ω)\Pi_{\mathcal{T},k}^{\operatorname*{inv}}q=\mathbf{v}_{q}+\Pi_{\mathcal{T}^{\prime},k}^{\operatorname*{inv}}\left(\left.\left(q-\operatorname*{div}\mathbf{v}_{q}\right)\right|_{\Omega^{\prime}}\right)

and satisfies divΠ𝒯,kinv=Id\operatorname*{div}\circ\Pi_{\mathcal{T},k}^{\operatorname*{inv}}=\operatorname*{Id} on k1,0(𝒯)\mathbb{P}_{k-1,0}\left(\mathcal{T}\right) and

𝒯Π𝒯,kinvq𝕃2(Ω)C3log(k+1)C𝒯qL2(Ω)\left\|\nabla_{\mathcal{T}}\Pi_{\mathcal{T},k}^{\operatorname*{inv}}q\right\|_{\mathbb{L}^{2}\left(\Omega\right)}\leq C_{3}\sqrt{\log\left(k+1\right)}C_{\mathcal{T}^{\prime}}\left\|q\right\|_{L^{2}\left(\Omega\right)}

for some C3C_{3} which only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega}.   

By iterating this argument we can prove Theorem 1.3 for the case (2.7).

Theorem 2.25

Let 𝒯\mathcal{T} be a conforming triangulation which contains at least one interior vertex. Let k5k\geq 5 be odd and let L0L\in\mathbb{N}_{0} the number of steps in the construction (2.13). Then, the inf-sup constant for the corresponding Crouzeix-Raviart discretization satisfies

c𝒯,kc𝒯(log(k+1))(L+1)/2c_{\mathcal{T},k}\geq c_{\mathcal{T}}\left(\log\left(k+1\right)\right)^{-\left(L+1\right)/2}

for some constant c𝒯c_{\mathcal{T}} which depends only on the shape-regularity of the mesh and αΩ\alpha_{\Omega}. If every triangle in 𝒯\mathcal{T} has an interior vertex, then L=0L=0.

2.3 The case of even kk

This case is slightly simpler that the case of odd kk since the non-conforming Crouzeix-Raviart functions for even kk have smaller support (i.e., one triangle) compared to two triangles (which share an edge) for odd kk.

In this section, we assume

a)k4 is even andb)𝒯 is a conforming triangulation and contains more than a single triangle.\begin{array}[c]{ll}\text{a)}&k\geq 4\text{ is even and}\\ \text{b)}&\mathcal{T}\text{ is a conforming triangulation and contains more than a single triangle.}\end{array} (2.61)
Remark 2.26

It is easy to verify that |𝒯|>1\left|\mathcal{T}\right|>1 implies that there exists a mapping 𝔎:𝒞𝒯(η)\mathfrak{K}:\mathcal{C}_{\mathcal{T}}\left(\eta\right) 𝒯\rightarrow\mathcal{T} with 𝔎(𝐳)𝒯𝐳\mathfrak{K}\left(\mathbf{z}\right)\in\mathcal{T}_{\mathbf{z}} and not all vertices of 𝔎(𝐳)\mathfrak{K}\left(\mathbf{z}\right) are η\eta-critical.

For an η\eta-critical point 𝐳𝒞𝒯(η)\mathbf{z}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right), let n𝐳:=|𝒯𝐳|n_{\mathbf{z}}:=\left|\mathcal{T}_{\mathbf{z}}\right| and fix a counterclockwise numbering of the triangles in

𝒯𝐳={Kj𝐳:1jn𝐳}\mathcal{T}_{\mathbf{z}}=\left\{K_{j}^{\mathbf{z}}:1\leq j\leq n_{\mathbf{z}}\right\} (2.62)

such that Kj𝐳K_{j}^{\mathbf{z}} and Kj+1𝐳K_{j+1}^{\mathbf{z}} share an edge for all 1jn𝐳11\leq j\leq n_{\mathbf{z}}-1. With this notation at hand, the functional A𝒯,𝐳A_{\mathcal{T},\mathbf{z}} is given by

A𝒯,𝐳q:=j=1n𝐳(1)jq|Kj𝐳(𝐳)qk1,0(𝒯).A_{\mathcal{T},\mathbf{z}}q:=\sum_{j=1}^{n_{\mathbf{z}}}\left(-1\right)^{j}\left.q\right|_{K_{j}^{\mathbf{z}}}\left(\mathbf{z}\right)\quad\forall q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right).
Lemma 2.27

Let assumption (2.61) be satisfied. There exists a constant η2>0\eta_{2}>0 which only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} such that for any fixed 0η<η20\leq\eta<\eta_{2} and any qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right), there exists some 𝐯q𝐂𝐑k,0(𝒯)\mathbf{v}_{q}\in\mathbf{CR}_{k,0}\left(\mathcal{T}\right) such that

Kdiv𝐯q\displaystyle\int_{K}\operatorname{div}\mathbf{v}_{q} =0K𝒯,\displaystyle=0\quad\forall K\in\mathcal{T}, (2.63)
qdiv𝒯𝐯q\displaystyle q-\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{v}_{q} Mη,k1SV(𝒯)\displaystyle\in M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) (2.64)

and

𝐯q𝐇1(𝒯)CCRlog(k+1)qL2(Ω).\left\|\mathbf{v}_{q}\right\|_{\mathbf{H}^{1}\left(\mathcal{T}\right)}\leq C_{\operatorname*{CR}}\sqrt{\log\left(k+1\right)}\left\|q\right\|_{L^{2}\left(\Omega\right)}. (2.65)

The constant CCRC_{\operatorname*{CR}} depends only on the shape-regularity of the mesh and αΩ\alpha_{\Omega}.

Proof. For a triangle K𝒯K\in\mathcal{T}, we set 𝒞K(η):={𝐳𝒱(K)𝒞𝒯(η)}\mathcal{C}_{K}\left(\eta\right):=\left\{\mathbf{z}\in\mathcal{V}\left(K\right)\cap\mathcal{C}_{\mathcal{T}}\left(\eta\right)\right\} and 𝒞Kactive(η):={𝐳𝒞K(η):K=𝔎(𝐳)}\mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right):=\left\{\mathbf{z}\in\mathcal{C}_{K}\left(\eta\right):K=\mathfrak{K}\left(\mathbf{z}\right)\right\}. Their cardinalities are denoted by nK:=|𝒞K(η)|n_{K}:=\left|\mathcal{C}_{K}\left(\eta\right)\right| and nKactive:=|𝒞Kactive(η)|n_{K}^{\operatorname*{active}}:=\left|\mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right)\right|. Note that 0nKactivenK20\leq n_{K}^{\operatorname*{active}}\leq n_{K}\leq 2 (cf. Rem. 2.26). We number the vertices 𝐕j\mathbf{V}_{j} in KK with the convention {𝐕j:1jnK}=𝒞K(η)\left\{\mathbf{V}_{j}:1\leq j\leq n_{K}\right\}=\mathcal{C}_{K}\left(\eta\right) and {𝐕j:1jnKactive}=𝒞Kactive(η)\left\{\mathbf{V}_{j}:1\leq j\leq n_{K}^{\operatorname*{active}}\right\}=\mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right). The angle in KK at 𝐕j\mathbf{V}_{j} is denoted by αj\alpha_{j}. Let EjE_{j} be the edge in KK opposite to 𝐕j\mathbf{V}_{j} and let 𝐧j\mathbf{n}_{j} denote the outward unit normal vector at EjE_{j}.

In a similar way as for the construction of B~k,ECR\tilde{B}_{k,E}^{\operatorname*{CR}} in Lemma 2.16 (and employing Lemma B.1 instead of Lemma B.2 for (2.21)) there exists a function B~k,KCRCRk,0(𝒯)\tilde{B}_{k,K}^{\operatorname*{CR}}\in\operatorname*{CR}_{k,0}\left(\mathcal{T}\right) with

  1. 1.

    suppB~k,KCR=K,\operatorname*{supp}\tilde{B}_{k,K}^{\operatorname*{CR}}=K,

  2. 2.

    for all K𝒯K\in\mathcal{T}, for all 𝐳𝒱(K)\mathbf{z}\in\mathcal{V}\left(K\right)

    (B~k,KCR|K)(𝐳)=(Bk,KCR|K)(𝐳),\nabla\left(\left.\tilde{B}_{k,K}^{\operatorname*{CR}}\right|_{K}\right)\left(\mathbf{z}\right)=\nabla\left(\left.B_{k,K}^{\operatorname*{CR}}\right|_{K}\right)\left(\mathbf{z}\right), (2.66)
  3. 3.

    for all K𝒯K\in\mathcal{T}

    B~k,KCR|K|K=Bk,KCR|K|K,\left.\left.\tilde{B}_{k,K}^{\operatorname*{CR}}\right|_{K}\right|_{\partial K}=\left.\left.B_{k,K}^{\operatorname*{CR}}\right|_{K}\right|_{\partial K}, (2.67)
  4. 4.

    for all K𝒯K\in\mathcal{T} and any 𝐜2\mathbf{c}\in\mathbb{R}^{2}

    Kdiv𝒯(B~k,KCR𝐜)=0.\int_{K}\operatorname{div}_{\mathcal{T}}\left(\tilde{B}_{k,K}^{\operatorname*{CR}}\mathbf{c}\right)=0. (2.68)
  5. 5.

    The piecewise gradient is bounded by

    B~k,KCRH1(𝒯)Clog(k+1).\left\|\tilde{B}_{k,K}^{\operatorname*{CR}}\right\|_{H^{1}\left(\mathcal{T}\right)}\leq C\sqrt{\log\left(k+1\right)}. (2.69)

For j=1,2j=1,2, we define

𝝍k,KCR,j:={B~k,KCR𝐧jin K,0in Ω\K,\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j}:=\left\{\begin{array}[c]{ll}\tilde{B}_{k,K}^{\operatorname*{CR}}\mathbf{n}_{j}&\text{in }K,\\ 0&\text{in }\Omega\backslash K,\end{array}\right. (2.70)

i.e., we fix 𝐯K:=𝐧1\mathbf{v}_{K}:=\mathbf{n}_{1} and 𝐰K=𝐧2\mathbf{w}_{K}=\mathbf{n}_{2} in (2.5). The divergence of 𝝍k,KCR,j\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j} evaluated at a vertex 𝐕s\mathbf{V}_{s}, s=1,2s=1,2, is given by

div(𝝍k,KCR,j|K)(𝐕s)\displaystyle\operatorname{div}\left(\left.\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j}\right|_{K}\right)\left(\mathbf{V}_{s}\right) =(2.66)div(Bk,KCR𝐧j|K)(𝐕s)=i=13Lk(12λK,i(𝐕s))𝐧jλK,i\displaystyle\overset{\text{(\ref{divtrianglerel})}}{=}\operatorname{div}\left(\left.B_{k,K}^{\operatorname*{CR}}\mathbf{n}_{j}\right|_{K}\right)\left(\mathbf{V}_{s}\right)=-{\displaystyle\sum\limits_{i=1}^{3}}L_{k}^{\prime}\left(1-2\lambda_{K,i}\left(\mathbf{V}_{s}\right)\right)\partial_{\mathbf{n}_{j}}\lambda_{K,i}
=k(k+1)𝐧jλK,s=(2.3)(k+12)|Es||K|×{1j=s,cosα3js.\displaystyle=k\left(k+1\right)\partial_{\mathbf{n}_{j}}\lambda_{K,s}\overset{\text{(\ref{normcomp})}}{=}\binom{k+1}{2}\frac{\left|E_{s}\right|}{\left|K\right|}\times\left\{\begin{array}[c]{ll}-1&j=s,\\ \cos\alpha_{3}&j\neq s.\end{array}\right.

Let qk1,0(𝒯)q\in\mathbb{P}_{k-1,0}\left(\mathcal{T}\right). We choose 𝜹K:=(δK,j)j=12\mbox{\boldmath$\delta$}_{K}:=\left(\delta_{K,j}\right)_{j=1}^{2} by the conditions for s=1,2s=1,2

A𝒯,𝐕s(div𝒯(j=12δK,j𝝍k,KCR,j))=!{A𝒯,𝐕s(q)if 𝐕s𝒞Kactive(η),0otherwise.A_{\mathcal{T},\mathbf{V}_{s}}\left(\operatorname*{div}\nolimits_{\mathcal{T}}\left(\sum_{j=1}^{2}\delta_{K,j}\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j}\right)\right)\overset{!}{=}\left\{\begin{array}[c]{ll}A_{\mathcal{T},\mathbf{V}_{s}}\left(q\right)&\text{if }\mathbf{V}_{s}\in\mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right),\\ 0&\text{otherwise.}\end{array}\right. (2.71)

For s=1,2s=1,2, let s\ell_{s} be defined by Ks𝐕s=KK_{\ell_{s}}^{\mathbf{V}_{s}}=K (cf. (2.62)). Then,

A𝒯,𝐕s(div𝒯(j=12δK,j𝝍k,KCR,j))\displaystyle A_{\mathcal{T},\mathbf{V}_{s}}\left(\operatorname*{div}\nolimits_{\mathcal{T}}\left(\sum_{j=1}^{2}\delta_{K,j}\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j}\right)\right) =(1)sj=12δK,j(div𝝍k,KCR,j|K)(𝐕s)\displaystyle=\left(-1\right)^{\ell_{s}}\sum_{j=1}^{2}\delta_{K,j}\left(\operatorname*{div}\left.\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j}\right|_{K}\right)\left(\mathbf{V}_{s}\right)
=𝐌K𝜹K\displaystyle=\mathbf{M}_{K}\mbox{\boldmath$\delta$}_{K}

for

𝐌K=(1)s+1(k+12)|K|1[|E1||E1|cosα3|E2|cosα3|E2|].\mathbf{M}_{K}=\left(-1\right)^{\ell_{s}+1}\binom{k+1}{2}\left|K\right|^{-1}\left[\begin{array}[c]{cc}\left|E_{1}\right|&-\left|E_{1}\right|\cos\alpha_{3}\\ -\left|E_{2}\right|\cos\alpha_{3}&\left|E_{2}\right|\end{array}\right].

We define 𝐫K=(rK,s)s=12\mathbf{r}_{K}=\left(r_{K,s}\right)_{s=1}^{2} by

rK,s:={A𝒯,𝐕s(q)if 𝐕s𝒞Kactive(η),0otherwiser_{K,s}:=\left\{\begin{array}[c]{ll}A_{\mathcal{T},\mathbf{V}_{s}}\left(q\right)&\text{if }\mathbf{V}_{s}\in\mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right),\\ 0&\text{otherwise}\end{array}\right.

so that 𝜹K\mbox{\boldmath$\delta$}_{K} is the solution of

𝐌K𝜹K=𝐫K.\mathbf{M}_{K}\mbox{\boldmath$\delta$}_{K}=\mathbf{r}_{K}.

Observe that

det[|E1||E1|cosα3|E2|cosα3|E2|]=|E1||E2|sin2α3=2|K|sinα3\det\left[\begin{array}[c]{cc}\left|E_{1}\right|&-\left|E_{1}\right|\cos\alpha_{3}\\ -\left|E_{2}\right|\cos\alpha_{3}&\left|E_{2}\right|\end{array}\right]=\left|E_{1}\right|\left|E_{2}\right|\sin^{2}\alpha_{3}=2\left|K\right|\sin\alpha_{3}

and sinα3sinϕ𝒯>0\sin\alpha_{3}\geq\sin\phi_{\mathcal{T}}>0 due to the shape-regularity of the mesh. For the coefficient 𝜹K\mbox{\boldmath$\delta$}_{K} we get explicitly

𝜹K=(1)s+1(k+1)ksinα3[|E2||E1|cosα3|E2|cosα3|E1|]𝐫K\mbox{\boldmath$\delta$}_{K}=\frac{\left(-1\right)^{\ell_{s}+1}}{\left(k+1\right)k\sin\alpha_{3}}\left[\begin{array}[c]{cc}\left|E_{2}\right|&\left|E_{1}\right|\cos\alpha_{3}\\ \left|E_{2}\right|\cos\alpha_{3}&\left|E_{1}\right|\end{array}\right]\mathbf{r}_{K} (2.72)

with an estimate

𝜹KChKk(k+1)𝐫KLem. 2.19CqL2(ωK)if 𝒞Kactive(η),\left\|\mbox{\boldmath$\delta$}_{K}\right\|\leq C\frac{h_{K}}{k\left(k+1\right)}\left\|\mathbf{r}_{K}\right\|\overset{\text{Lem. \ref{LemFunctional}}}{\leq}C\left\|q\right\|_{L^{2}\left(\omega_{K}\right)}\quad\text{if }\mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right)\neq\emptyset,

where CC only depends on the shape-regularity of the mesh. Note that this is the analogue for even kk to (2.44). If 𝒞Kactive(η)=\mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right)=\emptyset, it holds 𝜹K=𝟎\mbox{\boldmath$\delta$}_{K}=\mathbf{0}.

We define the global function

𝐯q:=K𝒯𝒞Kactive(η)j=12δK,j𝝍k,KCR,j.\mathbf{v}_{q}:=\sum_{\begin{subarray}{c}K\in\mathcal{T}\\ \mathcal{C}_{K}^{\operatorname*{active}}\left(\eta\right)\neq\emptyset\end{subarray}}\sum_{j=1}^{2}\delta_{K,j}\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j}. (2.73)

From (2.68) we conclude that 𝐯q\mathbf{v}_{q} satisfies (2.63).

Next, we verify (2.64). Let 𝐲𝒞𝒯(η)\mathbf{y}\in\mathcal{C}_{\mathcal{T}}\left(\eta\right) and recall the notation and convention as in (2.62). Let K𝐲=𝔎(𝐲)K_{\ell}^{\mathbf{y}}=\mathfrak{K}\left(\mathbf{y}\right). Then (2.64) follows from

A𝒯,𝐲(div𝒯𝐯q)\displaystyle A_{\mathcal{T},\mathbf{y}}\left(\operatorname*{div}\nolimits_{\mathcal{T}}\mathbf{v}_{q}\right) =(1)(div𝐯q|K𝐲)(𝐲)=(1)(divj=12δK𝐲,j𝝍k,K𝐲CR,j|K𝐲)(𝐲)\displaystyle=\left(-1\right)^{\ell}\left(\operatorname{div}\left.\mathbf{v}_{q}\right|_{K_{\ell}^{\mathbf{y}}}\right)\left(\mathbf{y}\right)=\left(-1\right)^{\ell}\left(\operatorname{div}\left.\sum_{j=1}^{2}\delta_{K_{\ell}^{\mathbf{y}},j}\mbox{\boldmath$\psi$}_{k,K_{\ell}^{\mathbf{y}}}^{\operatorname*{CR},j}\right|_{K_{\ell}^{\mathbf{y}}}\right)\left(\mathbf{y}\right)
=A𝒯,𝐲(q).\displaystyle=A_{\mathcal{T},\mathbf{y}}\left(q\right).

The estimate

𝝍k,KCR,j𝕃2(K)Clog(k+1)\left\|\nabla\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR},j}\right\|_{\mathbb{L}^{2}\left(K\right)}\leq C\sqrt{\log\left(k+1\right)}

for a constant CC which only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} follows directly from (2.69) and the final estimate (2.65) is derived by repeating the arguments as in the proof of Lemma 2.20 

This lemma allows us to extend Definition 2.21 to the case of even kk by defining the coefficients 𝜹K\mbox{\boldmath$\delta$}_{K} by (2.72) and the functions 𝝍k,KCR\mbox{\boldmath$\psi$}_{k,K}^{\operatorname*{CR}} by (2.70) and set (cf. (2.73)) ΠkCRq:=𝐯q\Pi_{k}^{\operatorname*{CR}}q:=\mathbf{v}_{q}. Since (IΠkCR)qMη,k1SV(𝒯)\left(I-\Pi_{k}^{\operatorname*{CR}}\right)q\in M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right) we may apply the further steps in the proof of Lemma 2.22 to obtain the inf-sup stability for even kk.

Theorem 2.28

Let 𝒯\mathcal{T} be a conforming triangulation satisfies (2.61). Then, the inf-sup constant for the corresponding Crouzeix-Raviart discretization satisfies

c𝒯,kc𝒯(log(k+1))1/2c_{\mathcal{T},k}\geq c_{\mathcal{T}}\left(\log\left(k+1\right)\right)^{-1/2}

for some constant c𝒯c_{\mathcal{T}} which depends only on the shape-regularity of the mesh and αΩ\alpha_{\Omega}.

3 Conclusion

In this paper, we have derived lower bounds for the inf-sup constant for Crouzeix-Raviart elements for the Stokes equation which are explicit with respect to the polynomial degree kk and are independent of the mesh size.

  1. 1.

    The inf-sup constant can be bounded from below by c𝒯,kc𝒯(log(k+1))1/2c_{\mathcal{T},k}\geq c_{\mathcal{T}}\left(\log\left(k+1\right)\right)^{-1/2} if

    1. (a)

      for odd k3k\geq 3,

      1. i.

        𝒯\mathcal{T} has at least one interior point and

      2. ii.

        for k5k\geq 5, 𝒯\mathcal{T} has no acute critical point,

    2. (b)

      for even k4k\geq 4, 𝒯\mathcal{T} contains more than one triangle,

  2. 2.

    If for odd kk, condition 1.a.ii. is not satisfied but a step-by-step construction (2.13) for some L1L\geq 1 is possible, then, c𝒯,kc𝒯(log(k+1))(L+1)/2c_{\mathcal{T},k}\geq c_{\mathcal{T}}\left(\log\left(k+1\right)\right)^{-\left(L+1\right)/2}.

Finally, we compare these findings with some other stable pairs of Stokes elements on triangulations in the literature. The element (𝐒k,0(𝒯),k2,0(𝒯))\left(\mathbf{S}_{k,0}\left(\mathcal{T}\right),\mathbb{P}_{k-2,0}\left(\mathcal{T}\right)\right) has a discrete inf-sup constant which can be estimated from below by Ck3Ck^{-3} (see [31], [33]). The discrete inf-sup constant for the Scott-Vogelius element (𝐒k,0(𝒯),M0,k1SV(𝒯))\left(\mathbf{S}_{k,0}\left(\mathcal{T}\right),M_{0,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right)\right) for k4k\geq 4 can be estimated from below by cΘminkmc\Theta_{\min}k^{-m} for some integer mm sufficiently large (see [32], [37]). The pressure-wired Stokes element (𝐒k,0(𝒯),Mη,k1SV(𝒯))\left(\mathbf{S}_{k,0}\left(\mathcal{T}\right),M_{\eta,k-1}^{\operatorname*{SV}}\left(\mathcal{T}\right)\right) in LABEL:Sauter_eta_wired (again for k4k\geq 4) is a mesh-robust generalization of the Scott-Vogelius element with a lower bound of the inf-sup constant of the form c(Θmin+η)c\left(\Theta_{\min}+\eta\right). In [3], a conforming stable pair (𝐗k(𝒯),Mk1(𝒯))\left(\mathbf{X}_{k}\left(\mathcal{T}\right),M_{k-1}\left(\mathcal{T}\right)\right) of Stokes elements on triangulations is introduced and it is proved that the discrete inf-sup constant can be estimated from below by c/Θ~minc/\tilde{\Theta}_{\min} for a constant cc independent of hh and kk and Θ~min:=min𝐳𝒱Ω(𝒯)\𝒞𝒯Θ(𝐳)\tilde{\Theta}_{\min}:=\min_{\mathbf{z}\in\mathcal{V}_{\partial\Omega}\left(\mathcal{T}\right)\backslash\mathcal{C}_{\mathcal{T}}}\Theta\left(\mathbf{z}\right). However, the implementation requires finite elements for the velocity with C1C^{1} continuity at the triangle vertices and pressures which are continuous in the triangle vertices.

Acknowledgement 3.1

Thanks are due to Benedikt Gräßle, HU Berlin, for fruitful discussions on Lemma 2.13.

I am grateful to my colleagues from TU Vienna, Profs. Joachim Schöberl and Markus Melenk. Joachim showed by numerical experiments that the lower bound of the inf-sup constant in the first arxiv version of the paper, namely k1/4k^{-1/4}, might be too pessimistic and “it should be at least k1/6k^{-1/6}” and Markus raised the suspicion that the interpolation argument might be too pessimistic for Legendre polynomials.

Appendix A The inverse of the matrix 𝐓+𝚫\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell} in (2.37)

Lemma A.1

There exists η2>0\eta_{2}>0 which only depends on the shape-regularity of the mesh and αΩ\alpha_{\Omega} such that for any 0η<η20\leq\eta<\eta_{2} the matrix 𝐓+𝚫\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell} in (2.37) is invertible and there exists a constant CC depending only on the shape-regularity of the mesh and αΩ\alpha_{\Omega} such that (cf. Notation 2.1)

(𝐓+𝚫)1C.\left\|\left(\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell}\right)^{-1}\right\|\leq C.

Note that the matrix 𝐓\mathbf{T}_{\ell} in (2.43) is the same as the matrix 𝐓n,𝜶\mathbf{T}_{n,\mbox{\boldmath$\alpha$}} which has been analysed in [10, (3.36)]. In particular, the formula

det𝐓=sin(j=1n+1α,j,1)j=1n+1sinα,j,1\det\mathbf{T}_{\ell}=\frac{\sin\left(\sum_{j=1}^{n_{\ell}+1}\alpha_{\ell,j,1}\right)}{{\displaystyle\prod\nolimits_{j=1}^{n_{\ell}+1}}\sin\alpha_{\ell,j,1}}

was proved.

Next we show that the sum j=1n+1α,j,1\sum_{j=1}^{n_{\ell}+1}\alpha_{\ell,j,1} is bounded away from 0 and π\pi. The bound j=1n+1α,j,1φ𝒯\sum_{j=1}^{n_{\ell}+1}\alpha_{\ell,j,1}\geq\varphi_{\mathcal{T}} follows from Remark 1.1. Since 𝐳,j\mathbf{z}_{\ell,j}, 1jn1\leq j\leq n_{\ell} are η\eta-critical points the sum of both angles adjacent to E,jE_{\ell,j} at 𝐳,j\mathbf{z}_{\ell,j} satisfy

sin(α,j,2+α,j+1,3)η\sin\left(\alpha_{\ell,j,2}+\alpha_{\ell,j+1,3}\right)\leq\eta\text{. }

We write α,j,2+α,j+1,3=:π+δj\alpha_{\ell,j,2}+\alpha_{\ell,j+1,3}=:\pi+\delta_{j}. From the proof of Lemma 2.10, in particular from the estimate (2.8) we conclude that |δj|c2η\left|\delta_{j}\right|\leq c_{2}\eta.

Since all points 𝐳,j\mathbf{z}_{\ell,j} are edge-connected to the same point 𝐳\mathbf{z}_{\ell}, the number nn_{\ell} is bounded from above by a constant nmaxn_{\max} which only depends on the shape-regularity of the mesh. Hence,

j=1n+1α,j,1\displaystyle\sum_{j=1}^{n_{\ell}+1}\alpha_{\ell,j,1} =j=1n+1(πα,j,2α,j,3)=(n+1)πα,1,3α,n+1,2j=1n(α,j,2+α,j+1,3)\displaystyle=\sum_{j=1}^{n_{\ell}+1}\left(\pi-\alpha_{\ell,j,2}-\alpha_{\ell,j,3}\right)=\left(n_{\ell}+1\right)\pi-\alpha_{\ell,1,3}-\alpha_{\ell,n_{\ell}+1,2}-\sum_{j=1}^{n_{\ell}}\left(\alpha_{\ell,j,2}+\alpha_{\ell,j+1,3}\right)
=(n+1)πα,1,3α,n+1,2j=1n(π+δj)=πα,1,3α,n+1,2j=1nδj\displaystyle=\left(n_{\ell}+1\right)\pi-\alpha_{\ell,1,3}-\alpha_{\ell,n_{\ell}+1,2}-\sum_{j=1}^{n_{\ell}}\left(\pi+\delta_{j}\right)=\pi-\alpha_{\ell,1,3}-\alpha_{\ell,n_{\ell}+1,2}-\sum_{j=1}^{n_{\ell}}\delta_{j}
π2φ𝒯+nmaxc2η.\displaystyle\leq\pi-2\varphi_{\mathcal{T}}+n_{\max}c_{2}\eta.

By adjusting the constant η0\eta_{0} in Lemma 2.10 to η1:=min{η0,φ𝒯/(nmaxc2)}\eta_{1}:=\min\left\{\eta_{0},\varphi_{\mathcal{T}}/\left(n_{\max}c_{2}\right)\right\} it follows that

j=1n+1α,j,1πφ𝒯.\sum_{j=1}^{n_{\ell}+1}\alpha_{\ell,j,1}\leq\pi-\varphi_{\mathcal{T}}\text{.}

By using the trivial estimate 0<sinα,j,110<\sin\alpha_{\ell,j,1}\leq 1, we may conclude that

det𝐓=sin(j=1n+1α,j,1)j=1n+1sinα,j,1sinφ𝒯>0.\det\mathbf{T}_{\ell}=\frac{\sin\left(\sum_{j=1}^{n_{\ell}+1}\alpha_{\ell,j,1}\right)}{{\displaystyle\prod\nolimits_{j=1}^{n_{\ell}+1}}\sin\alpha_{\ell,j,1}}\geq\sin\varphi_{\mathcal{T}}>0. (A.1)

Note that the entries in the matrix 𝐓\mathbf{T}_{\ell} (cf. 2.43) satisfy

|(𝐓)𝐲,𝐳|1sin2φ𝒯\left|\left(\mathbf{T}_{\ell}\right)_{\mathbf{y},\mathbf{z}}\right|\leq\frac{1}{\sin^{2}\varphi_{\mathcal{T}}} (A.2)

and hence the Frobenius norm 𝐓F\left\|\mathbf{T}_{\ell}\right\|_{\operatorname{F}} can be estimated by

𝐓F3nsin2φ𝒯3nmaxsin2φ𝒯.\left\|\mathbf{T}_{\ell}\right\|_{\operatorname{F}}\leq\frac{3n_{\ell}}{\sin^{2}\varphi_{\mathcal{T}}}\leq\frac{3n_{\max}}{\sin^{2}\varphi_{\mathcal{T}}}.

It is well known that 𝐓𝐓F\left\|\mathbf{T}_{\ell}\right\|\leq\left\|\mathbf{T}_{\ell}\right\|_{\operatorname{F}} and hence the bound on 𝐓\left\|\mathbf{T}_{\ell}\right\| follows.

We combine (A.1), (A.2), and nnmaxn_{\ell}\leq n_{\max} to obtain by Cramer’s rule that there exists a constant CC which only depends on the shape-regularity such that

|(𝐓1)𝐲,𝐳|C.\left|\left(\mathbf{T}_{\ell}^{-1}\right)_{\mathbf{y},\mathbf{z}}\right|\leq C.

By the same arguments as before we conclude that 𝐓1C~\left\|\mathbf{T}_{\ell}^{-1}\right\|\leq\tilde{C} for a constant C~\tilde{C} which only depends on the shape-regularity of the mesh.

Next we estimate 𝚫\left\|\mbox{\boldmath$\Delta$}_{\ell}\right\|. Since 𝚫\mbox{\boldmath$\Delta$}_{\ell} is diagonal it suffices to estimate the diagonal entries

|sin(α,j,2+α,j+1,3)sinα,j,2sinα,j+1,3|=|sin(π+δ)sinα,j,2sinα,j+1,3|=|sinδsinα,j,2sinα,j+1,3|c2ηsin2φ𝒯.\left|\frac{\sin\left(\alpha_{\ell,j,2}+\alpha_{\ell,j+1,3}\right)}{\sin\alpha_{\ell,j,2}\sin\alpha_{\ell,j+1,3}}\right|=\left|\frac{\sin\left(\pi+\delta_{\ell}\right)}{\sin\alpha_{\ell,j,2}\sin\alpha_{\ell,j+1,3}}\right|=\left|\frac{\sin\delta_{\ell}}{\sin\alpha_{\ell,j,2}\sin\alpha_{\ell,j+1,3}}\right|\leq\frac{c_{2}\eta}{\sin^{2}\varphi_{\mathcal{T}}}.

We write 𝐓+𝚫=𝐓(𝐈+𝐓1𝚫)\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell}=\mathbf{T}_{\ell}\left(\mathbf{I}+\mathbf{T}_{\ell}^{-1}\mbox{\boldmath$\Delta$}_{\ell}\right) and obtain

𝐓1𝚫𝐓1𝚫C~c2ηsin2φ𝒯.\left\|\mathbf{T}_{\ell}^{-1}\mbox{\boldmath$\Delta$}_{\ell}\right\|\leq\left\|\mathbf{T}_{\ell}^{-1}\right\|\left\|\mbox{\boldmath$\Delta$}_{\ell}\right\|\leq\tilde{C}\frac{c_{2}\eta}{\sin^{2}\varphi_{\mathcal{T}}}.

Next, we adjust the upper bound η1\eta_{1} by setting η2:=min{η1,sin2φ𝒯2C~c2}\eta_{2}:=\min\left\{\eta_{1},\frac{\sin^{2}\varphi_{\mathcal{T}}}{2\tilde{C}c_{2}}\right\} to obtain 𝐓1𝚫1/2\left\|\mathbf{T}_{\ell}^{-1}\mbox{\boldmath$\Delta$}_{\ell}\right\|\leq 1/2 with implies the invertibility of 𝐓+𝚫\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell} with bound

(𝐓+𝚫)12C~.\left\|\left(\mathbf{T}_{\ell}+\mbox{\boldmath$\Delta$}_{\ell}\right)^{-1}\right\|\leq 2\tilde{C}.

 

Appendix B Estimate of the H1/2H^{1/2} norm of traces of non-conforming Crouzeix-Raviart functions

In this appendix, we prove the norm estimate (2.69) for B~k,KCR\tilde{B}_{k,K}^{\operatorname*{CR}} and (2.19) for B~k,ECR\tilde{B}_{k,E}^{\operatorname*{CR}}. We first introduce some norms and semi-norms on the unit interval I:=[1,1]I:=\left[-1,1\right] in a formal way:

|u|H1/2(I):=(1111|u(s)u(t)st|2𝑑s𝑑t)1/2,uH1/2(I):=(uL2(I)2+|u|H1/2(I)2)1/2,|u|H(0,1/2(I):=(11|u(s)|21+s𝑑s)1/2,uH(0,1/2(I):=(uH1/2(I)2+|u|H(0,1/2(I)2)1/2,|u|H,0)1/2(I):=(11|u(s)|21s𝑑s)1/2,uH,0)1/2(I):=(uH1/2(I)2+|u|H,0)1/2(I)2)1/2,|u|H001/2(I):=(|u|H(0,1/2(I)2+|u|H,0)1/2(I)2)1/2,uH001/2(I):=(uH1/2(I)2+|u|H001/2(I)2)1/2.\begin{array}[c]{ll}\left|u\right|_{H^{1/2}\left(I\right)}:=\left({\displaystyle\int_{-1}^{1}}{\displaystyle\int_{-1}^{1}}\left|\dfrac{u\left(s\right)-u\left(t\right)}{s-t}\right|^{2}dsdt\right)^{1/2},&\left\|u\right\|_{H^{1/2}\left(I\right)}:=\left(\left\|u\right\|_{L^{2}\left(I\right)}^{2}+\left|u\right|_{H^{1/2}\left(I\right)}^{2}\right)^{1/2},\\ \left|u\right|_{H_{(0,}^{1/2}\left(I\right)}:=\left({\displaystyle\int_{-1}^{1}}\dfrac{\left|u\left(s\right)\right|^{2}}{1+s}ds\right)^{1/2},&\left\|u\right\|_{H_{(0,}^{1/2}\left(I\right)}:=\left(\left\|u\right\|_{H^{1/2}\left(I\right)}^{2}+\left|u\right|_{H_{(0,}^{1/2}\left(I\right)}^{2}\right)^{1/2},\\ \left|u\right|_{H_{,0)}^{1/2}\left(I\right)}:=\left({\displaystyle\int_{-1}^{1}}\dfrac{\left|u\left(s\right)\right|^{2}}{1-s}ds\right)^{1/2},&\left\|u\right\|_{H_{,0)}^{1/2}\left(I\right)}:=\left(\left\|u\right\|_{H^{1/2}\left(I\right)}^{2}+\left|u\right|_{H_{,0)}^{1/2}\left(I\right)}^{2}\right)^{1/2},\\ \left|u\right|_{H_{00}^{1/2}\left(I\right)}:=\left(\left|u\right|_{H_{(0,}^{1/2}\left(I\right)}^{2}+\left|u\right|_{H_{,0)}^{1/2}\left(I\right)}^{2}\right)^{1/2},&\left\|u\right\|_{H_{00}^{1/2}\left(I\right)}:=\left(\left\|u\right\|_{H^{1/2}\left(I\right)}^{2}+\left|u\right|_{H_{00}^{1/2}\left(I\right)}^{2}\right)^{1/2}.\end{array} (B.1)
Lemma B.1

Let k4k\geq 4 be even and K𝒯K\in\mathcal{T}. Let B~k,KCR|K\left.\tilde{B}_{k,K}^{\operatorname*{CR}}\right|_{\partial K} be defined by (2.67). Then there exists an absolute constant CC depending only on the shape regularity of 𝒯\mathcal{T} such that

B~k,KCRH1/2(K)Clog(k+1).\left\|\tilde{B}_{k,K}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial K\right)}\leq C\sqrt{\log\left(k+1\right)}.

Proof. We first prove the estimate for the reference element K^\widehat{K}. By construction (see (2.67)) the function B~k,K^CR\tilde{B}_{k,\widehat{K}}^{\operatorname*{CR}} coincides with Bk,K^CRB_{k,\widehat{K}}^{\operatorname*{CR}} on K^\partial\widehat{K}. Let the vertices of K^\widehat{K} be numbered counterclockwise and denoted by 𝐳^i\widehat{\mathbf{z}}_{i}, 1i31\leq i\leq 3. The edge opposite to 𝐳^i\widehat{\mathbf{z}}_{i} is E^i=[𝐳^i+1,𝐳^i1]\widehat{E}_{i}=\left[\widehat{\mathbf{z}}_{i+1},\widehat{\mathbf{z}}_{i-1}\right] (with cyclic numbering convention 𝐳3+1:=𝐳1\mathbf{z}_{3+1}:=\mathbf{z}_{1} and 𝐳11:=𝐳3\mathbf{z}_{1-1}:=\mathbf{z}_{3}). We choose the pullbacks to [1,1]\left[-1,1\right] by

ϕi(s):=𝐳^i+1+s(𝐳^i1𝐳^i+1)i=1,2,3\phi_{i}\left(s\right):=\widehat{\mathbf{z}}_{i+1}+s\left(\widehat{\mathbf{z}}_{i-1}-\widehat{\mathbf{z}}_{i+1}\right)\quad i=1,2,3 (B.2)

and observe B~k,K^CR|E^iϕi=Lk\left.\tilde{B}_{k,\widehat{K}}^{\operatorname*{CR}}\right|_{\widehat{E}_{i}}\circ\phi_{i}=L_{k}. From [25] (see also [4, p 1870]) we deduce that the H1/2(K^)H^{1/2}\left(\partial\widehat{K}\right) norm is equivalent to

|v|H1/2(K^):=(i=13(viH1/2(I)2+|di|H(0,1/2(I)2))1/2,\left|\kern-1.00006pt\left|\kern-1.00006pt\left|v\right|\kern-1.00006pt\right|\kern-1.00006pt\right|_{H^{1/2}\left(\partial\widehat{K}\right)}:=\left(\sum_{i=1}^{3}\left(\left\|v_{i}\right\|_{H^{1/2}\left(I\right)}^{2}+\left|d_{i}\right|_{H_{(0,}^{1/2}\left(I\right)}^{2}\right)\right)^{1/2}, (B.3)

where for vH1/2(K^)v\in H^{1/2}\left(\partial\widehat{K}\right):

vi:=v|Eiϕi,di(s):=vi1(s)vi+1(s).v_{i}:=\left.v\right|_{E_{i}}\circ\phi_{i},\quad d_{i}\left(s\right):=v_{i-1}\left(s\right)-v_{i+1}\left(-s\right). (B.4)

In our application (and even kk) we have vi1(s)=Lk(s)=Lk(s)=vi+1(s)v_{i-1}\left(s\right)=L_{k}\left(s\right)=L_{k}\left(-s\right)=v_{i+1}\left(-s\right) so that di=0d_{i}=0. We use LkL2(I)=2/(2k+1)\left\|L_{k}\right\|_{L^{2}\left(I\right)}=\sqrt{2/\left(2k+1\right)} to get

|Lk|H1/2(K^)(62k+1+3|Lk|H1/2(I)2)1/2.\left|\kern-1.00006pt\left|\kern-1.00006pt\left|L_{k}\right|\kern-1.00006pt\right|\kern-1.00006pt\right|_{H^{1/2}\left(\partial\widehat{K}\right)}\leq\left(\frac{6}{2k+1}+3\left|L_{k}\right|_{H^{1/2}\left(I\right)}^{2}\right)^{1/2}. (B.5)

This integral can be evaluated analytically for v=Lkv=L_{k} (see Lem. C.2) and we obtain

|Lk|H1/2(I)=2(=1k1)1/2Clog(k+1)k=1,2,\left|L_{k}\right|_{H^{1/2}\left(I\right)}=2\left(\sum_{\ell=1}^{k}\frac{1}{\ell}\right)^{1/2}\leq\sqrt{C\log\left(k+1\right)}\qquad\forall k=1,2,\ldots (B.6)

for a generic constant C>0C>0. This leads to the final estimate on the reference element:

B~k,K^CRH1/2(K^)\displaystyle\left\|\tilde{B}_{k,\widehat{K}}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial\widehat{K}\right)} C|B~k,K^CR|H1/2(K^)(B.5), (B.6)C(12k+1+log(k+1))1/2\displaystyle\leq C\left|\kern-1.00006pt\left|\kern-1.00006pt\left|\tilde{B}_{k,\widehat{K}}^{\operatorname*{CR}}\right|\kern-1.00006pt\right|\kern-1.00006pt\right|_{H^{1/2}\left(\partial\widehat{K}\right)}\overset{\text{(\ref{Lkhalfway}), (\ref{LkHalfEst})}}{\leq}C\left(\frac{1}{2k+1}+\log\left(k+1\right)\right)^{1/2}
Clog(k+1).\displaystyle\leq C\sqrt{\log\left(k+1\right)}.

For a triangle K𝒯K\in\mathcal{T}, let ϕK:K^K\phi_{K}:\widehat{K}\rightarrow K be an affine pullback and set B~k,K^CR=B~k,KCRϕK\tilde{B}_{k,\widehat{K}}^{\operatorname*{CR}}=\tilde{B}_{k,K}^{\operatorname*{CR}}\circ\phi_{K}. Then, the transformation rule for integrals yields

B~k,KCRH1/2(K)\displaystyle\left\|\tilde{B}_{k,K}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial K\right)} =(B~k,KCRL2(K)2+|B~k,KCR|H1/2(K)2)1/2\displaystyle=\left(\left\|\tilde{B}_{k,K}^{\operatorname*{CR}}\right\|_{L^{2}\left(\partial K\right)}^{2}+\left|\tilde{B}_{k,K}^{\operatorname*{CR}}\right|_{H^{1/2}\left(\partial K\right)}^{2}\right)^{1/2}
=C(hKB~k,K^CRL2(K^)2+|B~k,K^CR|H1/2(K^)2)1/2,\displaystyle=C\left(h_{K}\left\|\tilde{B}_{k,\widehat{K}}^{\operatorname*{CR}}\right\|_{L^{2}\left(\partial\widehat{K}\right)}^{2}+\left|\tilde{B}_{k,\widehat{K}}^{\operatorname*{CR}}\right|_{H^{1/2}\left(\partial\widehat{K}\right)}^{2}\right)^{1/2}, (B.7)

where CC only depends on the shape regularity of the mesh. The leads to the claim.  

The case of odd k5k\geq 5 is considered in the following lemma.

Lemma B.2

Let k5k\geq 5 be odd. For EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right), let the function B~k,ECR\tilde{B}_{k,E}^{\operatorname*{CR}} be as in the proof of Lemma 2.16. Then there exists a generic constant CC depending only on the shape regularity of 𝒯\mathcal{T} such that for any K𝒯EK\in\mathcal{T}_{E}, it holds

B~k,ECRH1/2(K)Clog(k+1).\left\|\tilde{B}_{k,E}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial K\right)}\leq C\sqrt{\log\left(k+1\right)}.

Proof. Let EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right) and K𝒯EK\in\mathcal{T}_{E}. Similarly as in (B.7) we have

B~k,ECRH1/2(K)CB~k,E^CRH1/2(K^)\left\|\tilde{B}_{k,E}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial K\right)}\leq C\left\|\tilde{B}_{k,\widehat{E}}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial\widehat{K}\right)}

with B~k,E^CR=B~k,ECRϕK\tilde{B}_{k,\widehat{E}}^{\operatorname*{CR}}=\tilde{B}_{k,E}^{\operatorname*{CR}}\circ\phi_{K} and affine pullback ϕK:K^K\phi_{K}:\widehat{K}\rightarrow K. Number the vertices in K^\widehat{K} counterclockwise 𝐳^i\widehat{\mathbf{z}}_{i}, 1i31\leq i\leq 3, such that 𝐳^3\widehat{\mathbf{z}}_{3} is opposite to E^:=ϕK1(E)\widehat{E}:=\phi_{K}^{-1}\left(E\right). The edge in K^\partial\widehat{K} opposite to 𝐳^i\widehat{\mathbf{z}}_{i} is denoted by E^i\widehat{E}_{i} and this implies E^=E^3\widehat{E}=\widehat{E}_{3}. The edgewise pullbacks ϕi\phi_{i} are defined as in (B.2). We employ the equivalence of the H1/2(K^)H^{1/2}\left(\partial\widehat{K}\right) norm with ||||||H1/2(K^)\left|\kern-1.00006pt\left|\kern-1.00006pt\left|\cdot\right|\kern-1.00006pt\right|\kern-1.00006pt\right|_{H^{1/2}\left(\partial\widehat{K}\right)} (see (B.3)) and obtain

B~k,E^CRH1/2(K^)C(i=13(viH1/2(I)2+|di|H(0,1/2(I)2))1/2\left\|\tilde{B}_{k,\widehat{E}}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial\widehat{K}\right)}\leq C\left(\sum_{i=1}^{3}\left(\left\|v_{i}\right\|_{H^{1/2}\left(I\right)}^{2}+\left|d_{i}\right|_{H_{(0,}^{1/2}\left(I\right)}^{2}\right)\right)^{1/2}

with

vi:=B~k,E^CR|E^iϕiand di(s):=vi1(s)vi+1(s).v_{i}:=\left.\tilde{B}_{k,\widehat{E}}^{\operatorname*{CR}}\right|_{\widehat{E}_{i}}\circ\phi_{i}\quad\text{and\quad}d_{i}\left(s\right):=v_{i-1}\left(s\right)-v_{i+1}\left(-s\right)\text{.}

Note that

v1(s)=Lk(s),v2(s)=Lk(s),v3=Lk1w~3v_{1}\left(s\right)=L_{k}\left(-s\right),\quad v_{2}\left(s\right)=L_{k}\left(s\right),\quad v_{3}=L_{k-1}-\tilde{w}_{3} (B.8)

with w~3:=Lk1(1)ψ~kLk1(1)ψ~k+\tilde{w}_{3}:=L_{k-1}^{\prime}\left(-1\right)\tilde{\psi}_{k}^{-}-L_{k-1}^{\prime}\left(1\right)\tilde{\psi}_{k}^{+} and ψ~k±\tilde{\psi}_{k}^{\pm} as in (2.20). The antisymmetry of the Legendre polynomial for odd kk implies d3=0d_{3}=0 and

B~k,E^CRH1/2(K^)C(LkH1/2(I)2+Lk1H1/2(I)2+w~3H1/2(I)2+i=12|di|H(0,1/2(I)2)1/2.\left\|\tilde{B}_{k,\widehat{E}}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial\widehat{K}\right)}\leq C\left(\left\|L_{k}\right\|_{H^{1/2}\left(I\right)}^{2}+\left\|L_{k-1}\right\|_{H^{1/2}\left(I\right)}^{2}+\left\|\tilde{w}_{3}\right\|_{H^{1/2}\left(I\right)}^{2}+\sum_{i=1}^{2}\left|d_{i}\right|_{H_{(0,}^{1/2}\left(I\right)}^{2}\right)^{1/2}. (B.9)

The estimates LjH1/2(I)2Clog(k+1)\left\|L_{j}\right\|_{H^{1/2}\left(I\right)}^{2}\leq C\log\left(k+1\right) for j{k1,k}j\in\left\{k-1,k\right\} follow from (B.6). For the last term, we employ

|di|H(0,1/2(I)\displaystyle\left|d_{i}\right|_{H_{(0,}^{1/2}\left(I\right)} (B.4), (B.8)|Lk+Lk1|H(0,1/2(I)+|w~3|H(0,1/2(I)\displaystyle\overset{\text{(\ref{defdi}), (\ref{defvineu})}}{\leq}\left|L_{k}+L_{k-1}\right|_{H_{(0,}^{1/2}\left(I\right)}+\left|\tilde{w}_{3}\right|_{H_{(0,}^{1/2}\left(I\right)}
(B.1)|Lk+Lk1|H(0,1/2(I)+|w~3|H001/2(I).\displaystyle\overset{\text{(\ref{halfseminorms})}}{\leq}\left|L_{k}+L_{k-1}\right|_{H_{(0,}^{1/2}\left(I\right)}+\left|\tilde{w}_{3}\right|_{H_{00}^{1/2}\left(I\right)}.

The last term can be estimated by taking into account |Lk(±1)|=(k+12)\left|L_{k}^{\prime}\left(\pm 1\right)\right|=\binom{k+1}{2} (obtained, e.g., by evaluating and differentiating [17, 18.5.8 for the choice α=β=0\alpha=\beta=0 and 18.5.10 for the choice λ=1/2.\lambda=1/2.] at ±1\pm 1):

w~3H001/2(I)C(k+1)2(ψ~kH001/2(I)+ψ~k+H001/2(I))[3, (A.5)]C.\left\|\tilde{w}_{3}\right\|_{H_{00}^{1/2}\left(I\right)}\leq C\left(k+1\right)^{2}\left(\left\|\tilde{\psi}_{k}^{-}\right\|_{H_{00}^{1/2}\left(I\right)}+\left\|\tilde{\psi}_{k}^{+}\right\|_{H_{00}^{1/2}\left(I\right)}\right)\overset{\text{\cite[cite]{[\@@bibref{}{Ainsworth_parker_I}{}{}, (A.5)]}}}{\leq}C.

For the third term in (B.9), we simply employ w~3H1/2(I)w~3H001/2(I)\left\|\tilde{w}_{3}\right\|_{H^{1/2}\left(I\right)}\leq\left\|\tilde{w}_{3}\right\|_{H_{00}^{1/2}\left(I\right)} so that

B~k,E^CRH1/2(K^)C(log(k+1)+|Lk+Lk1|H(0,1/2(I)).\left\|\tilde{B}_{k,\widehat{E}}^{\operatorname*{CR}}\right\|_{H^{1/2}\left(\partial\widehat{K}\right)}\leq C\left(\sqrt{\log\left(k+1\right)}+\left|L_{k}+L_{k-1}\right|_{H_{(0,}^{1/2}\left(I\right)}\right).

The last integral can be evaluated analytically: By using the recurrence relation (cf. [17, 18.9.1])

Lk(x)=2k1kxLk1(x)k1kLk2(x)L_{k}\left(x\right)=\frac{2k-1}{k}xL_{k-1}\left(x\right)-\frac{k-1}{k}L_{k-2}\left(x\right)

we obtain

Lk(x)+Lk1(x)=2k1k(x+1)Lk1(x)k1k(Lk1(x)+Lk2(x)).L_{k}\left(x\right)+L_{k-1}\left(x\right)=\frac{2k-1}{k}\left(x+1\right)L_{k-1}\left(x\right)-\frac{k-1}{k}\left(L_{k-1}\left(x\right)+L_{k-2}\left(x\right)\right). (B.10)

This and the orthogonality properties of the Legendre polynomials lead to

Ik\displaystyle I_{k} :=|Lk+Lk1|H(0,1/2(I)2=11(2k1k(x+1)Lk1(x)k1k(Lk1(x)+Lk2(x)))2x+1𝑑x\displaystyle:=\left|L_{k}+L_{k-1}\right|_{H_{(0,}^{1/2}\left(I\right)}^{2}=\int_{-1}^{1}\frac{\left(\frac{2k-1}{k}\left(x+1\right)L_{k-1}\left(x\right)-\frac{k-1}{k}\left(L_{k-1}\left(x\right)+L_{k-2}\left(x\right)\right)\right)^{2}}{x+1}dx
=(2k1k)211(x+1)Lk12(x)𝑑x4(k1)k2+(k1k)2Ik1.\displaystyle=\left(\frac{2k-1}{k}\right)^{2}\int_{-1}^{1}\left(x+1\right)L_{k-1}^{2}\left(x\right)dx-\frac{4\left(k-1\right)}{k^{2}}+\left(\frac{k-1}{k}\right)^{2}I_{k-1}. (B.11)

The integral in (B.11) will be evaluated in (C.1). We obtain

(2k1k)211(x+1)Lk12(x)𝑑x=2(2k1)k2\left(\frac{2k-1}{k}\right)^{2}\int_{-1}^{1}\left(x+1\right)L_{k-1}^{2}\left(x\right)dx=\frac{2\left(2k-1\right)}{k^{2}}

and the explicit recursion formula

Ik=2k2+(k1k)2Ik1I_{k}=\frac{2}{k^{2}}+\left(\frac{k-1}{k}\right)^{2}I_{k-1}

with starting value

I1=11(L1(x)+L0(x))2x+1𝑑x=11(1+x)2x+1𝑑x=2.I_{1}=\int_{-1}^{1}\frac{\left(L_{1}\left(x\right)+L_{0}\left(x\right)\right)^{2}}{x+1}dx=\int_{-1}^{1}\frac{\left(1+x\right)^{2}}{x+1}dx=2.

It is easy to verify that Ik=2/kI_{k}=2/k solves the recursion and hence,

|Lk+Lk1|H(0,1/2(I)=2/k.\left|L_{k}+L_{k-1}\right|_{H_{(0,}^{1/2}\left(I\right)}=\sqrt{2/k}.

From this, the assertion follows.  

Appendix C Analytic evaluation of some integrals involving Legendre polynomials

In the proof of Lemma 2.20 some integrals over Legendre polynomials appear and we present here their explicit evaluation.

Lemma C.1

For k0k\geq 0, it holds

11Lk2(t)=11(t+1)Lk2(t)𝑑t=22k+1\int_{-1}^{1}L_{k}^{2}\left(t\right)=\int_{-1}^{1}\left(t+1\right)L_{k}^{2}\left(t\right)dt=\frac{2}{2k+1} (C.1)

and

11(Lk(t))2𝑑t=11(t+1)(Lk(t))2𝑑t=k(k+1).\int_{-1}^{1}\left(L_{k}^{\prime}\left(t\right)\right)^{2}dt=\int_{-1}^{1}\left(t+1\right)\left(L_{k}^{\prime}\left(t\right)\right)^{2}dt=k\left(k+1\right). (C.2)

Proof. The relation 11Lk2(t)𝑑t=2/(2k+1)\int_{-1}^{1}L_{k}^{2}\left(t\right)dt=2/\left(2k+1\right) follows from [23, 7.221(1)].

For k=0k=0, these relations follows from L0(t)=1L_{0}\left(t\right)=1. Let k1k\geq 1. The recurrence relation in [17, 18.9.1, Table 18.9.1] imply

(t+1)Lk(t)=k+12k+1Lk+1(t)+Lk(t)+k2k+1Lk1(t).\left(t+1\right)L_{k}\left(t\right)=\frac{k+1}{2k+1}L_{k+1}\left(t\right)+L_{k}\left(t\right)+\frac{k}{2k+1}L_{k-1}\left(t\right). (C.3)

Substituting (t+1)Lk(t)\left(t+1\right)L_{k}\left(t\right) under the integral in (C.1) by this and taking into account the orthogonality relations of the Legendre polynomials leads to

11(t+1)Lk2(t)𝑑t=11Lk2(t)𝑑t=22k+1.\int_{-1}^{1}\left(t+1\right)L_{k}^{2}\left(t\right)dt=\int_{-1}^{1}L_{k}^{2}\left(t\right)dt=\frac{2}{2k+1}.

For the second integral (C.2) we employ integration by parts:

11(t+1)(Lk(t))2𝑑t=(t+1)Lk(t)Lk(t)|1111gk(t)Lk(t)𝑑tfor g(t):=((t+1)Lk(t)).\int_{-1}^{1}\left(t+1\right)\left(L_{k}^{\prime}\left(t\right)\right)^{2}dt=\left.\left(t+1\right)L_{k}^{\prime}\left(t\right)L_{k}\left(t\right)\right|_{-1}^{1}-\int_{-1}^{1}g_{k}\left(t\right)L_{k}\left(t\right)dt\quad\text{for }g\left(t\right):=\left(\left(t+1\right)L_{k}^{\prime}\left(t\right)\right)^{\prime}. (C.4)

Since gk1g\in\mathbb{P}_{k-1} the orthogonality properties of Legendre polynomials imply that the integral in the right-hand side of (C.4) is zero. By using Lk(±1)=(±1)kL_{k}\left(\pm 1\right)=\left(\pm 1\right)^{k} (cf. [17, Table 18.6.1]) and Lk(±1)=(±1)k+1(k+12)L_{k}^{\prime}\left(\pm 1\right)=\left(\pm 1\right)^{k+1}\binom{k+1}{2} (cf. [17, combine 18.9.15 with Table 18.6.1]) we get

11(t+1)(Lk(t))2𝑑t=2Lk(1)Lk(1)=k(k+1).\int_{-1}^{1}\left(t+1\right)\left(L_{k}^{\prime}\left(t\right)\right)^{2}dt=2L_{k}^{\prime}\left(1\right)L_{k}\left(1\right)=k\left(k+1\right).

Finally

11(Lk(t))2𝑑t=Lk(t)Lk(t)|t=1111Lk′′(t)Lk(t)𝑑t.\int_{-1}^{1}\left(L_{k}^{\prime}\left(t\right)\right)^{2}dt=\left.L_{k}^{\prime}\left(t\right)L_{k}\left(t\right)\right|_{t=-1}^{1}-\int_{-1}^{1}L_{k}^{\prime\prime}\left(t\right)L_{k}\left(t\right)dt.

The last integral is zero due the orthogonality of the Legendre polynomials. The endpoint values of LkL_{k} and LkL_{k}^{\prime} lead to the assertion.  

In the final part of this section, we will compute the value of |Lk|H1/2(I)\left|L_{k}\right|_{H^{1/2}\left(I\right)} explicitly. We set

Ik(s):=|Lk|Hs(I)2=1111(Lk(x)Lk(y))2|xy|1+2s𝑑y𝑑x.I_{k}\left(s\right):=\left|L_{k}\right|_{H^{s}\left(I\right)}^{2}=\int_{-1}^{1}\int_{-1}^{1}\frac{\left(L_{k}\left(x\right)-L_{k}\left(y\right)\right)^{2}}{\left|x-y\right|^{1+2s}}dydx.
Lemma C.2

It holds

Ik(1/2)=4=1k1.I_{k}\left(1/2\right)=4\sum_{\ell=1}^{k}\frac{1}{\ell}.

Proof. We write

(Lk(y)Lk(x))2==22kκ(x)!(yx)\left(L_{k}\left(y\right)-L_{k}\left(x\right)\right)^{2}=\sum_{\ell=2}^{2k}\frac{\kappa_{\ell}\left(x\right)}{\ell!}\left(y-x\right)^{\ell}

with

κ(x)\displaystyle\kappa_{\ell}\left(x\right) :=ddy((Lk(y)Lk(x))2)|yx\displaystyle:=\left.\frac{d^{\ell}}{dy^{\ell}}\left(\left(L_{k}\left(y\right)-L_{k}\left(x\right)\right)^{2}\right)\right|_{y\leftarrow x}
=r=0(r)(drdyr(Lk(y)Lk(x)))drdyr(Lk(y)Lk(x))|yx\displaystyle=\sum_{r=0}^{\ell}\binom{\ell}{r}\left.\left(\frac{d^{r}}{dy^{r}}\left(L_{k}\left(y\right)-L_{k}\left(x\right)\right)\right)\frac{d^{\ell-r}}{dy^{\ell-r}}\left(L_{k}\left(y\right)-L_{k}\left(x\right)\right)\right|_{y\leftarrow x}
=r=11(r)(drdyr(Lk(y)Lk(x)))drdyr(Lk(y)Lk(x))|yx\displaystyle=\sum_{r=1}^{\ell-1}\binom{\ell}{r}\left.\left(\frac{d^{r}}{dy^{r}}\left(L_{k}\left(y\right)-L_{k}\left(x\right)\right)\right)\frac{d^{\ell-r}}{dy^{\ell-r}}\left(L_{k}\left(y\right)-L_{k}\left(x\right)\right)\right|_{y\leftarrow x}
=r=11(r)Lk(r)(x)Lk(r)(x)=(Lk2)()(x)2Lk(x)Lk()(x).\displaystyle=\sum_{r=1}^{\ell-1}\binom{\ell}{r}L_{k}^{\left(r\right)}\left(x\right)L_{k}^{\left(\ell-r\right)}\left(x\right)=\left(L_{k}^{2}\right)^{\left(\ell\right)}\left(x\right)-2L_{k}\left(x\right)L_{k}^{\left(\ell\right)}\left(x\right).

Hence,

Ik(s):=IkI(s)IkII(s)I_{k}\left(s\right):=I_{k}^{\operatorname*{I}}\left(s\right)-I_{k}^{\operatorname*{II}}\left(s\right)

with

IkI(s)\displaystyle I_{k}^{\operatorname*{I}}\left(s\right) ==22k1!1111(Lk2)()(x)(yx)|xy|1+2s𝑑y𝑑x,\displaystyle=\sum_{\ell=2}^{2k}\frac{1}{\ell!}\int_{-1}^{1}\int_{-1}^{1}\frac{\left(L_{k}^{2}\right)^{\left(\ell\right)}\left(x\right)\left(y-x\right)^{\ell}}{\left|x-y\right|^{1+2s}}dydx,
IkII(s)\displaystyle I_{k}^{\operatorname*{II}}\left(s\right) =2=22k1!1111Lk(x)Lk()(x)(yx)|xy|1+2s𝑑y𝑑x.\displaystyle=2\sum_{\ell=2}^{2k}\frac{1}{\ell!}\int_{-1}^{1}\int_{-1}^{1}\frac{L_{k}\left(x\right)L_{k}^{\left(\ell\right)}\left(x\right)\left(y-x\right)^{\ell}}{\left|x-y\right|^{1+2s}}dydx.

We perform the integration with respect to yy explicitly and get

IkI(s)\displaystyle I_{k}^{\operatorname*{I}}\left(s\right) ==22k1!11(Lk2)()(x)w,s(x)𝑑x,\displaystyle=\sum_{\ell=2}^{2k}\frac{1}{\ell!}\int_{-1}^{1}\left(L_{k}^{2}\right)^{\left(\ell\right)}\left(x\right)w_{\ell,s}\left(x\right)dx,
IkII(s)\displaystyle I_{k}^{\operatorname*{II}}\left(s\right) =2=22k1!11Lk(x)Lk()(x)w,s(x)𝑑x\displaystyle=2\sum_{\ell=2}^{2k}\frac{1}{\ell!}\int_{-1}^{1}L_{k}\left(x\right)L_{k}^{\left(\ell\right)}\left(x\right)w_{\ell,s}\left(x\right)dx

with

w,s(x):=(1)(1+x)2s+(1x)2s2s.w_{\ell,s}\left(x\right):=\frac{\left(-1\right)^{\ell}\left(1+x\right)^{\ell-2s}+\left(1-x\right)^{\ell-2s}}{\ell-2s}.

From now on we restrict to the case s=1/2s=1/2. Since w,1/2Lk()w_{\ell,1/2}L_{k}^{\left(\ell\right)} is a polynomial of maximal degree k1k-1 the second integral vanishes: IkII(1/2)=0I_{k}^{\operatorname*{II}}\left(1/2\right)=0. We apply recursively integration by parts and use the orthogonality of the Legendre polynomials to obtain

Ik(1/2)=IkI(1/2)==22km=02k(1)m2m+2(m+)(m+1)1(m+1)!(Lk2)(m+1)(1).I_{k}\left(1/2\right)=I_{k}^{\operatorname*{I}}\left(1/2\right)=\sum_{\ell=2}^{2k}\sum_{m=0}^{2k-\ell}\frac{\left(-1\right)^{m}2^{m+2}}{\left(m+\ell\right)\left(m+\ell-1\right)}\frac{1}{\left(m+1\right)!}\left(L_{k}^{2}\right)^{\left(m+1\right)}\left(1\right).

We interchange the ordering of the summation, introduce the new variable t=+m2t=\ell+m-2, and obtain

Ik(1/2)\displaystyle I_{k}\left(1/2\right) =m=02k2=22km(1)m2m+2(m+)(m+1)1(m+1)!(Lk2)(m+1)(1)\displaystyle=\sum_{m=0}^{2k-2}\sum_{\ell=2}^{2k-m}\frac{\left(-1\right)^{m}2^{m+2}}{\left(m+\ell\right)\left(m+\ell-1\right)}\frac{1}{\left(m+1\right)!}\left(L_{k}^{2}\right)^{\left(m+1\right)}\left(1\right)
=m=02k2(1)m2m+2(m+1)!(Lk2)(m+1)(1)t=m2k21(t+2)(t+1).\displaystyle=\sum_{m=0}^{2k-2}\frac{\left(-1\right)^{m}2^{m+2}}{\left(m+1\right)!}\left(L_{k}^{2}\right)^{\left(m+1\right)}\left(1\right)\sum_{t=m}^{2k-2}\frac{1}{\left(t+2\right)\left(t+1\right)}.

By using a telescoping sum argument, it is easy to verify that the inner sum equals

t=m2k21(t+2)(t+1)=t=m2k2(1t+11t+2)=1m+112k.\sum_{t=m}^{2k-2}\frac{1}{\left(t+2\right)\left(t+1\right)}=\sum_{t=m}^{2k-2}\left(\frac{1}{t+1}-\frac{1}{t+2}\right)=\frac{1}{m+1}-\frac{1}{2k}.

Hence,

Ik(1/2)=1kIkIII(1/2)2IkIV(1/2),I_{k}\left(1/2\right)=\frac{1}{k}I_{k}^{\operatorname*{III}}\left(1/2\right)-2I_{k}^{\operatorname*{IV}}\left(1/2\right), (C.5)

for

IkIII(1/2)\displaystyle I_{k}^{\operatorname*{III}}\left(1/2\right) :=m=02k2(1)m2m+1(m+1)!(Lk2)(m+1)(1),\displaystyle:=-\sum_{m=0}^{2k-2}\frac{\left(-1\right)^{m}2^{m+1}}{\left(m+1\right)!}\left(L_{k}^{2}\right)^{\left(m+1\right)}\left(1\right),
IkIV(1/2)\displaystyle I_{k}^{\operatorname*{IV}}\left(1/2\right) :=m=02k2(1)m+12m+1(m+1)!(m+1)(Lk2)(m+1)(1).\displaystyle:=\sum_{m=0}^{2k-2}\frac{\left(-1\right)^{m+1}2^{m+1}}{\left(m+1\right)!\left(m+1\right)}\left(L_{k}^{2}\right)^{\left(m+1\right)}\left(1\right).

For IkIIII_{k}^{\operatorname*{III}} and k1k\geq 1 we get

IkIII(1/2)\displaystyle I_{k}^{\operatorname*{III}}\left(1/2\right) =m=12k1(1)m2mm!(Lk2)(m)(1)\displaystyle=\sum_{m=1}^{2k-1}\frac{\left(-1\right)^{m}2^{m}}{m!}\left(L_{k}^{2}\right)^{\left(m\right)}\left(1\right)
=Lk2(1)22k(2k)!(Lk2)(2k)(1)+m=02k(1)m2mm!(Lk2)(m)(1).\displaystyle=-L_{k}^{2}\left(1\right)-\frac{2^{2k}}{\left(2k\right)!}\left(L_{k}^{2}\right)^{\left(2k\right)}\left(1\right)+\sum_{m=0}^{2k}\frac{\left(-1\right)^{m}2^{m}}{m!}\left(L_{k}^{2}\right)^{\left(m\right)}\left(1\right). (C.6)

We use the endpoint formula Lk(m)(±1)=0L_{k}^{\left(m\right)}\left(\pm 1\right)=0 for m>km>k and for m{0,1,,k}:m\in\left\{0,1,\ldots,k\right\}:

Lk(m)(±1)=[1, 22.5.37, 22.4.2](±1)k+m(2m1)!!(k+mkm)=(±1)k+m1(2m)!!(k+m)!(km)!.L_{k}^{\left(m\right)}\left(\pm 1\right)\overset{\text{\cite[cite]{[\@@bibref{}{Abramowitz}{}{}, 22.5.37, 22.4.2]}}}{=}\left(\pm 1\right)^{k+m}\left(2m-1\right)!!\binom{k+m}{k-m}=\left(\pm 1\right)^{k+m}\frac{1}{\left(2m\right)!!}\frac{\left(k+m\right)!}{\left(k-m\right)!}.

This leads to Lk2(1)=1L_{k}^{2}\left(1\right)=1 and the Leibniz rule for differentiation yields

22k(2k)!(Lk2)(2k)(1)\displaystyle\frac{2^{2k}}{\left(2k\right)!}\left(L_{k}^{2}\right)^{\left(2k\right)}\left(1\right) =22k(2k)!=02k(2k)Lk()(1)Lk(2k)(1)=22k(2k)!(2kk)(Lk(k)(1))2\displaystyle=\frac{2^{2k}}{\left(2k\right)!}\sum_{\ell=0}^{2k}\binom{2k}{\ell}L_{k}^{\left(\ell\right)}\left(1\right)L_{k}^{\left(2k-\ell\right)}\left(1\right)=\frac{2^{2k}}{\left(2k\right)!}\binom{2k}{k}\left(L_{k}^{\left(k\right)}\left(1\right)\right)^{2}
=22k(2k)!(2kk)((2k)!(2k)!!)2=22k(2k)!(2k)!(k!)2(2k)!222k(k)!2=(2k)!2(k)!4.\displaystyle=\frac{2^{2k}}{\left(2k\right)!}\binom{2k}{k}\left(\frac{\left(2k\right)!}{\left(2k\right)!!}\right)^{2}=\frac{2^{2k}}{\left(2k\right)!}\frac{\left(2k\right)!}{\left(k!\right)^{2}}\frac{\left(2k\right)!^{2}}{2^{2k}\left(k\right)!^{2}}=\frac{\left(2k\right)!^{2}}{\left(k\right)!^{4}}.

The last sum in (C.6) is the Taylor expansion of Lk2L_{k}^{2} about x=1x=1, evaluated at x=1x=-1, i.e.,

m=02k(1)m2mm!(Lk2)(m)(1)=(Lk2)(1)=1.\sum_{m=0}^{2k}\frac{\left(-1\right)^{m}2^{m}}{m!}\left(L_{k}^{2}\right)^{\left(m\right)}\left(1\right)=\left(L_{k}^{2}\right)\left(-1\right)=1.

Hence,

IkIII(1/2)=(2k)!2(k)!4.I_{k}^{\operatorname*{III}}\left(1/2\right)=-\frac{\left(2k\right)!^{2}}{\left(k\right)!^{4}}. (C.7)

For the quantity IkIVI_{k}^{\operatorname*{IV}} we obtain by similar arguments

IkIV(1/2)\displaystyle I_{k}^{\operatorname*{IV}}\left(1/2\right) =m=12k1(1)m2mm!m(Lk2)(m)(1)\displaystyle=\sum_{m=1}^{2k-1}\frac{\left(-1\right)^{m}2^{m}}{m!m}\left(L_{k}^{2}\right)^{\left(m\right)}\left(1\right)
=22k1(2k)!k(Lk2)(2k)(1)111s1m=12k1m!(Lk2)(m)(1)(s1)mds\displaystyle=-\frac{2^{2k-1}}{\left(2k\right)!k}\left(L_{k}^{2}\right)^{\left(2k\right)}\left(1\right)-\int_{-1}^{1}\frac{1}{s-1}\sum_{m=1}^{2k}\frac{1}{m!}\left(L_{k}^{2}\right)^{\left(m\right)}\left(1\right)\left(s-1\right)^{m}ds
=12k(2k)!2(k)!411Lk2(s)Lk2(1)s1𝑑s.\displaystyle=-\frac{1}{2k}\frac{\left(2k\right)!^{2}}{\left(k\right)!^{4}}-\int_{-1}^{1}\frac{L_{k}^{2}\left(s\right)-L_{k}^{2}\left(1\right)}{s-1}ds.

The combination of this with (C.5), (C.7) leads to

Ik(1/2)=211Lk2(s)1s1𝑑s=211Lk(s)1s1(Lk(s)+1)𝑑s.I_{k}\left(1/2\right)=2\int_{-1}^{1}\frac{L_{k}^{2}\left(s\right)-1}{s-1}ds=2\int_{-1}^{1}\frac{L_{k}\left(s\right)-1}{s-1}\left(L_{k}\left(s\right)+1\right)ds.

Since Lk(s)1s1\frac{L_{k}\left(s\right)-1}{s-1} is a polynomial of maximal degree k1k-1 the orthogonality of LkL_{k} leads to

Ik(1/2)=211Lk(s)1s1𝑑s.I_{k}\left(1/2\right)=2\int_{-1}^{1}\frac{L_{k}\left(s\right)-1}{s-1}ds. (C.8)

We employ the recursion formula in [17, Table 18.9.1] for k2k\geq 2

Lk(s)=2k1ksLk1(s)k1kLk2(s)L_{k}\left(s\right)=\frac{2k-1}{k}sL_{k-1}\left(s\right)-\frac{k-1}{k}L_{k-2}\left(s\right)

so that

Ik(1/2)\displaystyle I_{k}\left(1/2\right) =2112k1k(sLk1(s)1)k1k(Lk2(s)1)s1𝑑s\displaystyle=2\int_{-1}^{1}\frac{\frac{2k-1}{k}\left(sL_{k-1}\left(s\right)-1\right)-\frac{k-1}{k}\left(L_{k-2}\left(s\right)-1\right)}{s-1}ds
=2k1k211(Lk1(s)+Lk1(s)1s1)𝑑sk1kIk2(1/2)\displaystyle=\frac{2k-1}{k}2\int_{-1}^{1}\left(L_{k-1}\left(s\right)+\frac{L_{k-1}\left(s\right)-1}{s-1}\right)ds-\frac{k-1}{k}I_{k-2}\left(1/2\right)
=2k1kIk1(1/2)k1kIk2(1/2).\displaystyle=\frac{2k-1}{k}I_{k-1}\left(1/2\right)-\frac{k-1}{k}I_{k-2}\left(1/2\right). (C.9)

From (C.8) and L0(s)=1L_{0}\left(s\right)=1, L1(s)=sL_{1}\left(s\right)=s, L2(s)=(3s21)/2L_{2}\left(s\right)=\left(3s^{2}-1\right)/2 we get

I0(1/2)=0,I1(1/2)=211L1(s)1s1𝑑s=4,I2(1/2)=211L2(s)1s1𝑑s=6.\begin{array}[c]{l}I_{0}\left(1/2\right)=0,\\ I_{1}\left(1/2\right)=2\int_{-1}^{1}\frac{L_{1}\left(s\right)-1}{s-1}ds=4,\\ I_{2}\left(1/2\right)=2\int_{-1}^{1}\frac{L_{2}\left(s\right)-1}{s-1}ds=6.\end{array} (C.10)

Now, it is easy to verify that Ik=4=1k1I_{k}=4\sum_{\ell=1}^{k}\frac{1}{\ell} satisfies the recursion (C.9) and the initial value conditions (C.10).  

Appendix D Norm equivalence for Crouzeix-Raviart spaces

It is well known that for V:=H01(Ω)+CRk,0(𝒯)V:=H_{0}^{1}\left(\Omega\right)+\operatorname*{CR}_{k,0}\left(\mathcal{T}\right), the norms (𝒯uL2(Ω)2+uL2(Ω)2)1/2\left(\left\|\nabla_{\mathcal{T}}u\right\|_{L^{2}\left(\Omega\right)}^{2}+\left\|u\right\|_{L^{2}\left(\Omega\right)}^{2}\right)^{1/2} and 𝒯uL2(Ω)\left\|\nabla_{\mathcal{T}}u\right\|_{L^{2}\left(\Omega\right)} are equivalent. In this section, we state estimates for the constants in these equivalencies – the proof is a repetition of the well-known arguments for the case k=1k=1 (see, e.g., [19, Lem. 36.6]).

Theorem D.1

There exists a constant C>0C>0 depending only on the shape-regularity of the mesh and the domain Ω\Omega such that

uH1(𝒯)(𝒯uL2(Ω)2+uL2(Ω)2)1/2CuH1(𝒯)uV.\left\|u\right\|_{H^{1}\left(\mathcal{T}\right)}\leq\left(\left\|\nabla_{\mathcal{T}}u\right\|_{L^{2}\left(\Omega\right)}^{2}+\left\|u\right\|_{L^{2}\left(\Omega\right)}^{2}\right)^{1/2}\leq C\left\|u\right\|_{H^{1}\left(\mathcal{T}\right)}\quad\forall u\in V.

In particular CC is independent of the polynomial degree k1k\geq 1 and the mesh size h𝒯h_{\mathcal{T}}.

Proof. We prove this result only under the regularity assumption that the Poisson problem:

find ϕH01(Ω)s.t. (ϕ,v)𝐋2(Ω)=(f,v)L2(Ω)vH01(Ω)\text{find }\phi\in H_{0}^{1}\left(\Omega\right)\quad\text{s.t.\quad}\left(\nabla\phi,\nabla v\right)_{\mathbf{L}^{2}\left(\Omega\right)}=\left(f,v\right)_{L^{2}\left(\Omega\right)}\quad\forall v\in H_{0}^{1}\left(\Omega\right)

is H2H^{2} regular. For less regularity we refer to [19, Lem. 36.6]. For uVu\in V, we have

uL2(Ω)=supvL2(Ω)\{0}(u,v)L2(Ω)vL2(Ω).\left\|u\right\|_{L^{2}\left(\Omega\right)}=\sup_{v\in L^{2}\left(\Omega\right)\backslash\left\{0\right\}}\frac{\left(u,v\right)_{L^{2}\left(\Omega\right)}}{\left\|v\right\|_{L^{2}\left(\Omega\right)}}. (D.1)

For vL2(Ω)v\in L^{2}\left(\Omega\right), there exists some 𝐰𝐇1(Ω)\mathbf{w}\in\mathbf{H}^{1}\left(\Omega\right) such that div𝐰=v\operatorname*{div}\mathbf{w}=v and 𝐰𝐇1(Ω)CΩvL2(Ω)\left\|\mathbf{w}\right\|_{\mathbf{H}^{1}\left(\Omega\right)}\leq C_{\Omega}\left\|v\right\|_{L^{2}\left(\Omega\right)} for a constant CΩC_{\Omega} which only depends on Ω\Omega. Hence,

(u,v)L2(Ω)=(u,div𝐰)L2(Ω)=(𝒯u,𝐰)L2(Ω)+K𝒯K𝐰,𝐧Ku,\left(u,v\right)_{L^{2}\left(\Omega\right)}=\left(u,\operatorname*{div}\mathbf{w}\right)_{L^{2}\left(\Omega\right)}=-\left(\nabla_{\mathcal{T}}u,\mathbf{w}\right)_{L^{2}\left(\Omega\right)}+\sum_{K\in\mathcal{T}}\int_{\partial K}\left\langle\mathbf{w},\mathbf{n}_{K}\right\rangle u,

where 𝐧K\mathbf{n}_{K} is the unit normal vector pointing to the exterior of KK. Next, we rewrite the sum over the triangle boundaries as a sum over the edges. For EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right) we fix the direction of a unit vector 𝐧E\mathbf{n}_{E} which is orthogonal to EE and for EΩ(𝒯)E\in\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right) let 𝐧E\mathbf{n}_{E} denote the unit vector, orthogonal to EE, pointing to the exterior of Ω\Omega. Then

(u,v)L2(Ω)=(u,div𝐰)L2(Ω)=(𝒯u,𝐰)L2(Ω)EΩ(𝒯)E𝐰,𝐧E[u]E+EΩ(𝒯)E𝐰,𝐧Eu.\left(u,v\right)_{L^{2}\left(\Omega\right)}=\left(u,\operatorname*{div}\mathbf{w}\right)_{L^{2}\left(\Omega\right)}=-\left(\nabla_{\mathcal{T}}u,\mathbf{w}\right)_{L^{2}\left(\Omega\right)}-\sum_{E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)}\int_{E}\left\langle\mathbf{w},\mathbf{n}_{E}\right\rangle\left[u\right]_{E}+\sum_{E\in\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right)}\int_{E}\left\langle\mathbf{w},\mathbf{n}_{E}\right\rangle u.

Let KE𝒯EK_{E}\in\mathcal{T}_{E} be fixed and let 𝐪E(0(KE))2\mathbf{q}_{E}\in\left(\mathbb{P}_{0}\left(K_{E}\right)\right)^{2} be the function with constant value 1|E|E𝐰\frac{1}{\left|E\right|}\int_{E}\mathbf{w}. The orthogonality conditions of the Crouzeix-Raviart elements across edges (see (1.9b)) imply

(u,v)L2(Ω)\displaystyle\left(u,v\right)_{L^{2}\left(\Omega\right)} =(𝒯u,𝐰)L2(Ω)EΩ(𝒯)E𝐰𝐪E,𝐧E[u]E+EΩ(𝒯)E𝐰𝐪E,𝐧Eu\displaystyle=-\left(\nabla_{\mathcal{T}}u,\mathbf{w}\right)_{L^{2}\left(\Omega\right)}-\sum_{E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)}\int_{E}\left\langle\mathbf{w}-\mathbf{q}_{E},\mathbf{n}_{E}\right\rangle\left[u\right]_{E}+\sum_{E\in\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right)}\int_{E}\left\langle\mathbf{w}-\mathbf{q}_{E},\mathbf{n}_{E}\right\rangle u (D.2)
𝒯u𝐋2(Ω)𝐰𝐋2(Ω)\displaystyle\leq\left\|\nabla_{\mathcal{T}}u\right\|_{\mathbf{L}^{2}\left(\Omega\right)}\left\|\mathbf{w}\right\|_{\mathbf{L}^{2}\left(\Omega\right)}
+EΩ(𝒯)[u]EL2(E)𝐰𝐪E𝐋2(E)+EΩ(𝒯)uL2(E)𝐰𝐪E𝐋2(E).\displaystyle+\sum_{E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right)}\left\|\left[u\right]_{E}\right\|_{L^{2}\left(E\right)}\left\|\mathbf{w}-\mathbf{q}_{E}\right\|_{\mathbf{L}^{2}\left(E\right)}+\sum_{E\in\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right)}\left\|u\right\|_{L^{2}\left(E\right)}\left\|\mathbf{w}-\mathbf{q}_{E}\right\|_{\mathbf{L}^{2}\left(E\right)}.

We employ first a weighted trace inequality (see, e.g., [18, Lem. 12.15]) and then a Poincaré-Steklov estimate (see, e.g., [19, (12.17) for p=2p=2 and s=1s=1.]) to get for hE:=|E|h_{E}:=\left|E\right|

𝐰𝐪E𝐋2(E)C(hE1/2𝐰𝐪E𝐋2(KE)+hE1/2(𝐰𝐪E)𝕃2(KE))ChKE1/2𝐰𝕃2(KE),\left\|\mathbf{w}-\mathbf{q}_{E}\right\|_{\mathbf{L}^{2}\left(E\right)}\leq C\left(h_{E}^{-1/2}\left\|\mathbf{w}-\mathbf{q}_{E}\right\|_{\mathbf{L}^{2}\left(K_{E}\right)}+h_{E}^{1/2}\left\|\nabla\left(\mathbf{w}-\mathbf{q}_{E}\right)\right\|_{\mathbb{L}^{2}\left(K_{E}\right)}\right)\leq Ch_{K_{E}}^{1/2}\left\|\nabla\mathbf{w}\right\|_{\mathbb{L}^{2}\left(K_{E}\right)}, (D.3)

where CC only depends on the shape-regularity of the mesh.

Next we estimate the jump of uu across EE. For EΩ(𝒯)E\in\mathcal{E}_{\Omega}\left(\mathcal{T}\right), we define uE0(𝒯E)u_{E}\in\mathbb{P}_{0}\left(\mathcal{T}_{E}\right) as the function with constant value 1|E|Eu|K\frac{1}{\left|E\right|}\int_{E}\left.u\right|_{K} on K𝒯EK\in\mathcal{T}_{E} and observe [uE]E=0\left[u_{E}\right]_{E}=0. Hence,

[u]EL2(E)\displaystyle\left\|\left[u\right]_{E}\right\|_{L^{2}\left(E\right)} =[uuE]EL2(E)K𝒯E(uuE)|KL2(E)\displaystyle=\left\|\left[u-u_{E}\right]_{E}\right\|_{L^{2}\left(E\right)}\leq\sum_{K\in\mathcal{T}_{E}}\left\|\left.\left(u-u_{E}\right)\right|_{K}\right\|_{L^{2}\left(E\right)}
K𝒯E(hE1/2uuEL2(K)+hE1/2u𝐋2(K))CK𝒯EhK1/2u𝐋2(K),\displaystyle\leq\sum_{K\in\mathcal{T}_{E}}\left(h_{E}^{-1/2}\left\|u-u_{E}\right\|_{L^{2}\left(K\right)}+h_{E}^{1/2}\left\|\nabla u\right\|_{\mathbf{L}^{2}\left(K\right)}\right)\leq C\sum_{K\in\mathcal{T}_{E}}h_{K}^{1/2}\left\|\nabla u\right\|_{\mathbf{L}^{2}\left(K\right)},

for a constant CC which only depends on the shape-regularity of the mesh. For EΩ(𝒯)E\in\mathcal{E}_{\partial\Omega}\left(\mathcal{T}\right) the estimate uL2(E)hK1/2u𝐋2(K)\left\|u\right\|_{L^{2}\left(E\right)}\leq h_{K}^{1/2}\left\|\nabla u\right\|_{\mathbf{L}^{2}\left(K\right)} for K𝒯EK\in\mathcal{T}_{E} follows in a similar fashion. The combination of (D.2) with (D.3) and the two trace estimates for uu leads to

(u,v)L2(Ω)\displaystyle\left(u,v\right)_{L^{2}\left(\Omega\right)} 𝒯u𝐋2(Ω)𝐰𝐋2(Ω)+CE(𝒯)𝒯u𝐋2(ωE)𝐰𝐋2(ωE)\displaystyle\leq\left\|\nabla_{\mathcal{T}}u\right\|_{\mathbf{L}^{2}\left(\Omega\right)}\left\|\mathbf{w}\right\|_{\mathbf{L}^{2}\left(\Omega\right)}+C\sum_{E\in\mathcal{E}\left(\mathcal{T}\right)}\left\|\nabla_{\mathcal{T}}u\right\|_{\mathbf{L}^{2}\left(\omega_{E}\right)}\left\|\nabla\mathbf{w}\right\|_{\mathbf{L}^{2}\left(\omega_{E}\right)}
C𝒯u𝐋2(Ω)𝐰𝐇1(Ω)CCΩ𝒯u𝐋2(Ω)qL2(Ω).\displaystyle\leq C\left\|\nabla_{\mathcal{T}}u\right\|_{\mathbf{L}^{2}\left(\Omega\right)}\left\|\mathbf{w}\right\|_{\mathbf{H}^{1}\left(\Omega\right)}\leq CC_{\Omega}\left\|\nabla_{\mathcal{T}}u\right\|_{\mathbf{L}^{2}\left(\Omega\right)}\left\|q\right\|_{L^{2}\left(\Omega\right)}.

Using this estimate in (D.1) finishes the proof.  

References

  • [1] M. Abramowitz and I. A. Stegun. Handbook of Mathematical Functions. Applied Mathematics Series 55. National Bureau of Standards, U.S. Department of Commerce, 1972.
  • [2] M. Ainsworth and S. Jiang. Preconditioning the mass matrix for high order finite element approximation on triangles. SIAM J. Numer. Anal., 57(1):355–377, 2019.
  • [3] M. Ainsworth and C. Parker. Mass conserving mixed hphp-FEM approximations to Stokes flow. Part I: uniform stability. SIAM J. Numer. Anal., 59(3):1218–1244, 2021.
  • [4] M. Ainsworth and C. Parker. Mass conserving mixed hphp-FEM approximations to Stokes flow. Part II: optimal convergence. SIAM J. Numer. Anal., 59(3):1245–1272, 2021.
  • [5] M. Ainsworth and R. Rankin. Fully computable bounds for the error in nonconforming finite element approximations of arbitrary order on triangular elements. SIAM J. Numer. Anal., 46(6):3207–3232, 2008.
  • [6] I. Babuška, A. Craig, J. Mandel, and J. Pitkäranta. Efficient preconditioning for the pp-version finite element method in two dimensions. SIAM J. Numer. Anal., 28(3):624–661, 1991.
  • [7] Á. Baran and G. Stoyan. Gauss-Legendre elements: a stable, higher order non-conforming finite element family. Computing, 79(1):1–21, 2007.
  • [8] C. Bernardi and G. Raugel. Analysis of some finite elements for the Stokes problem. Math. Comp., 44(169):71–79, 1985.
  • [9] S. C. Brenner. Forty years of the Crouzeix-Raviart element. Numer. Methods Partial Differential Equations, 31(2):367–396, 2015.
  • [10] C. Carstensen and S. Sauter. Critical functions and inf-sup stability of Crouzeix-Raviart elements. Comput. Math. Appl., 108:12–23, 2022.
  • [11] C. Carstensen and S. A. Sauter. Crouzeix-Raviart triangular elements are inf-sup stable. Math. Comp., 91(337):2041–2057, 2022.
  • [12] Y. Cha, M. Lee, and S. Lee. Stable nonconforming methods for the Stokes problem. Appl. Math. Comput., 114(2-3):155–174, 2000.
  • [13] P. G. Ciarlet, P. Ciarlet, Jr., S. A. Sauter, and C. Simian. Intrinsic finite element methods for the computation of fluxes for Poisson’s equation. Numer. Math., 132(3):433–462, 2016.
  • [14] P. Ciarlet, Jr., C. F. Dunkl, and S. A. Sauter. A family of Crouzeix-Raviart finite elements in 3D. Anal. Appl. (Singap.), 16(5):649–691, 2018.
  • [15] M. Crouzeix and R. Falk. Nonconforming finite elements for Stokes problems. Math. Comp., 186:437–456, 1989.
  • [16] M. Crouzeix and P.-A. Raviart. Conforming and nonconforming finite element methods for solving the stationary Stokes equations. I. Rev. Française Automat. Informat. Recherche Opérationnelle Sér. Rouge, 7(R-3):33–75, 1973.
  • [17] NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/, Release 1.0.13 of 2016-09-16. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller and B. V. Saunders, eds.
  • [18] A. Ern and J.-L. Guermond. Finite elements I—Approximation and interpolation. Springer, Cham, 2021.
  • [19] A. Ern and J.-L. Guermond. Finite elements II—Galerkin approximation, elliptic and mixed PDEs. Springer, Cham, 2021.
  • [20] M. Fortin. A three-dimensional quadratic nonconforming element. Numer. Math., 46(2):269–279, 1985.
  • [21] M. Fortin and M. Soulie. A nonconforming quadratic finite element on triangles. International Journal for Numerical Methods in Engineering, 19:505–520, 1983.
  • [22] V. Girault and P. Raviart. Finite Element Methods for Navier-Stokes Equations. Springer, Berlin, 1986.
  • [23] I. S. Gradshteyn and I. Ryzhik. Table of Integrals, Series, and Products. Academic Press, New York, London, 1965.
  • [24] B. Gräßle, N.-E. Bohne, and S. A. Sauter. The pressure-wired stokes element: a mesh-robust version of the Scott-Vogelius element, (extended edition), arXiv: 2212.09673, 2022.
  • [25] P. Grisvard. Elliptic Problems in Nonsmooth Domains. Pitman, Boston, 1985.
  • [26] J. Guzmán and M. Neilan. Conforming and divergence-free Stokes elements on general triangular meshes. Math. Comp., 83(285):15–36, 2014.
  • [27] J. Guzmán and L. R. Scott. The Scott-Vogelius finite elements revisited. Math. Comp., 88(316):515–529, 2019.
  • [28] G. Matthies and L. Tobiska. Inf-sup stable non-conforming finite elements of arbitrary order on triangles. Numer. Math., 102(2):293–309, 2005.
  • [29] W. McLean. Strongly Elliptic Systems and Boundary Integral Equations. Cambridge, Univ. Press, 2000.
  • [30] S. Sauter and C. Torres. On the Inf-Sup Stabillity of Crouzeix-Raviart Stokes Elements in 3D. Math. Comp., (https://doi.org/10.1090/mcom/3793), 2022.
  • [31] C. Schwab and M. Suri. Mixed hphp finite element methods for Stokes and non-Newtonian flow. Comput. Methods Appl. Mech. Engrg., 175(3-4):217–241, 1999.
  • [32] L. R. Scott and M. Vogelius. Norm estimates for a maximal right inverse of the divergence operator in spaces of piecewise polynomials. RAIRO Modél. Math. Anal. Numér., 19(1):111–143, 1985.
  • [33] R. Stenberg and M. Suri. Mixed hphp finite element methods for problems in elasticity and Stokes flow. Numer. Math., 72(3):367–389, 1996.
  • [34] G. Stoyan and Á. Baran. Crouzeix-Velte decompositions for higher-order finite elements. Comput. Math. Appl., 51(6-7):967–986, 2006.
  • [35] A. Veeser and P. Zanotti. Quasi-optimal nonconforming methods for symmetric elliptic problems. I—Abstract theory. SIAM J. Numer. Anal., 56(3):1621–1642, 2018.
  • [36] A. Veeser and P. Zanotti. Quasi-optimal nonconforming methods for symmetric elliptic problems. II—Overconsistency and classical nonconforming elements. SIAM J. Numer. Anal., 57(1):266–292, 2019.
  • [37] M. Vogelius. A right-inverse for the divergence operator in spaces of piecewise polynomials. Numerische Mathematik, 41(1):19–37, 1983.
  • [38] T. Warburton and J. S. Hesthaven. On the constants in hphp-finite element trace inverse inequalities. Comput. Methods Appl. Mech. Engrg., 192(25):2765–2773, 2003.