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

Normal-normal continuous symmetric stress approximation in three-dimensional linear elasticity thanks: Supported by ANID-Chile through FONDECYT project 1230013

Carsten Carstensen Department of Mathematics, Humboldt–Universität zu Berlin, Unter den Linden 6, 10099 Berlin, Germany, email: [email protected]    Norbert Heuer Facultad de Matemáticas, Pontificia Universidad Católica de Chile, Avenida Vicuña Mackenna 4860, Santiago, Chile, email: [email protected]
Abstract

We present a conforming setting for a mixed formulation of linear elasticity with symmetric stress that has normal-normal continuous components across faces of tetrahedral meshes. We provide a stress element for this formulation with 3030 degrees of freedom that correspond to standard boundary conditions. The resulting scheme converges quasi-optimally and is locking free. Numerical experiments illustrate the performance.


AMS Subject Classification: 65N30, 74G15, 74S05

1 Introduction and model problem

Adams & Cockburn [2] and Arnold et al. [4] presented and analyzed, with slightly different degrees of freedom, the first stable element for conforming stress approximations in 3D linear elasticity on tetrahedral meshes. It provides pointwise symmetric piecewise polynomial approximations of degree at least four with continuous normal components across faces. With its 162 degrees of freedom (plus displacement degrees) this element is prohibitively expensive for practical applications. Modifications and generalizations to arbitrary dimension and higher order have been proposed by Hu & Zhang, Hu [16, 15] and Chen & Huang [9]. The difficulty of constructing conforming elements with few degrees of freedom has been the reason for alternative developments, e.g., of discontinuous and non-conforming methods with weak symmetry. But the quest for cheap stable and conforming stress elements goes on, in particular on tetrahedra which provide important geometric flexibility. Hu and Zhang [17] proposed to use low-degree polynomial tensors enriched with edge and face related bubble polynomials to achieve stability, in this way reducing the degrees of freedom to 4848. A variation of this approach and extension to arbitrary dimension is given in [18].

In this paper we present a pointwise symmetric tetrahedral element of quadratic polynomials with 3030 degrees of freedom (they can be reduced to 1212 by static condensation) and provide a variational framework that renders the element conforming. In geometrical terms our element leads to linear normal-normal components on faces that are continuous across elements. This approach goes back to the seminal TDNNS —tangential displacement and normal-normal stress continuous— method by Pechstein & Schöberl [21, 22, 23] that in turn transfers the Hellan–Herrmann–Johnson idea from plate bending to linear elasticity, cf. [12, 13, 20]. In [7] we presented a continuous and discrete framework for this setting in the case of plane elasticity with mixed approximation on triangular meshes. Its advantages are

  • a low number of degrees of freedom related to

  • physical degrees of freedom that allow for the

  • treatment of any boundary condition that is physically relevant, and a

  • variational framework that makes the element conforming.

Furthermore, it is

  • provably stable and locking free.

This paper solves the critical open case of three-dimensional elasticity on tetrahedral elements. We stress the fact that all the advantages listed above remain valid in three dimensions. We expect that the existence of a conforming framework facilitates, for instance, the analysis of a posteriori error estimation. Perhaps more important is the fact that, differently from the previously mentioned H(div)H(\mathrm{div})-conforming methods, the degrees of freedom of our element avoid vertex values and edge moments of stresses. Such degrees do not relate to boundary conditions in engineering applications and are only well posed when making artificially high regularity assumptions.

In order not to be repetitive, we only briefly recall settings and cite proofs from [7], and also refer to [14] for general results on a hybrid framework. Instead, we focus on the new ingredients required for the three-dimensional setting. In particular, we restrict the analysis to homogeneous Dirichlet boundary conditions. The incorporation and analysis of general boundary conditions are similar to the two-dimensional situation treated in [7].

Model problem. We consider a bounded, polyhedral Lipschitz domain Ω3\Omega\subset\mathbb{R}^{3} with boundary Γ:=Ω\Gamma:=\partial\Omega decomposed into non-intersecting relatively open pieces

Γ=closure(Γℎ𝑐Γ𝑠𝑐Γ𝑠𝑠Γ𝑠𝑓).\displaystyle\Gamma=\mathrm{closure}({\Gamma_{\mathit{hc}}}\cup{\Gamma_{\mathit{sc}}}\cup{\Gamma_{\mathit{ss}}}\cup{\Gamma_{\mathit{sf}}}).

Boundary conditions are of hard clamped (“ℎ𝑐\mathit{hc}”), soft clamped (“𝑠𝑐\mathit{sc}”), simply supported (“𝑠𝑠\mathit{ss}”) and traction (“𝑠𝑓\mathit{sf}” for stress free) types on the corresponding boundary pieces. Individual sets are either empty or unions of connected sets of positive measure, subject to the validity of Korn’s inequality. For the discretization it has to be assumed that the boundary pieces are conforming with respect to the faces of the tetrahedral elements, that is, interfaces between different boundary pieces do not cut faces.

The model problem of linear elasticity with general boundary conditions reads

σ=2με(u)+λ(divu)idanddivσ=finΩ,\displaystyle\sigma=2\mu\varepsilon(u)+\lambda(\operatorname{div}u)\operatorname{\mathrm{id}}\quad\text{and}\quad-\operatorname{div}\sigma=f\quad\text{in}\ \Omega,
u\displaystyle u =gD\displaystyle={g_{D}} onΓℎ𝑐,\displaystyle\text{on}\ {\Gamma_{\mathit{hc}}},\qquad un=gD,n,πt(σn)=gN,t\displaystyle u\cdot n={g_{D,n}},\ \pi_{t}(\sigma n)={g_{N,t}}\quad onΓ𝑠𝑐,\displaystyle\text{on}\ {\Gamma_{\mathit{sc}}}, (1)
πtu\displaystyle\pi_{t}u =gD,t,nσn=gN,n\displaystyle={g_{D,t}},\ n\cdot\sigma n={g_{N,n}}\quad onΓ𝑠𝑠,\displaystyle\text{on}\ {\Gamma_{\mathit{ss}}}, σn=gN\displaystyle\sigma n={g_{N}} onΓ𝑠𝑓.\displaystyle\text{on}\ {\Gamma_{\mathit{sf}}}.

Here, ε(u)\varepsilon(u) and id\operatorname{\mathrm{id}} are the strain and identity tensors, respectively, and μ,λ>0\mu,\lambda>0 are the Lamé parameters. We assume that μ\mu is fixed but allow for general λ\lambda, corresponding to an unrestricted Poisson ratio ν(0,1/2)\nu\in(0,1/2). For short, we write σ=ε(u)\sigma=\mathbb{C}\varepsilon(u) with elasticity tensor \mathbb{C}, and let 𝔸:=1\mathbb{A}:=\mathbb{C}^{-1} be the compliance tensor. Furthermore, nn is the exterior unit normal vector on Γ\Gamma, and πtu:=(u(un)n)|Γ\pi_{t}u:=(u-(u\cdot n)n)|_{\Gamma} is the tangential trace on Γ\Gamma of vector fields uu. Boundary data have to be consistent with the existence of a function uH1(Ω;3)u\in H^{1}(\Omega;\mathbb{R}^{3}) and tensor σH(div,Ω;3×3)\sigma\in H(\operatorname{div},\Omega;\mathbb{R}^{3\times 3}) that satisfy (1) as appropriate traces. Note that f,gD,gD,t,gN,gN,tf,{g_{D}},{g_{D,t}},{g_{N}},{g_{N,t}} are vector-valued and gD,n,gN,n{g_{D,n}},{g_{N,n}} are scalar functions.

As mentioned before, our forthcoming variational formulation and mixed finite element discretization are compatible with the general boundary conditions (1). Only for ease of presentation will we restrict ourselves to homogeneous Dirichlet conditions, thus consider

σ=ε(u),divσ=finΩ,u|Γ=0,\displaystyle\sigma=\mathbb{C}\varepsilon(u),\quad-\operatorname{div}\sigma=f\quad\text{in}\ \Omega,\qquad u|_{\Gamma}=0, (2)

and refer to [7] for details concerning general boundary conditions.

Overview. The remainder is structured as follows. In Section 2 we introduce our notation for spaces and norms. At the center is the definition of skeleton trace spaces of tangential and normal displacements, their characterization (Propositions 12) and density in the space of displacement traces (Proposition 4). The proof of the density result is based on a 3D version of Tartar’s result that says that the set of CC^{\infty}-functions with compact support that vanish in a neighborhood of a point xΩ2x\in\Omega\subset\mathbb{R}^{2} are dense in H01(Ω)H^{1}_{0}(\Omega). This is Proposition 3. Having these results at hand, we follow the canonical procedure from [7] to define the space Hnn(div,𝒯;SS)H_{nn}(\operatorname{div},\mathcal{T};\SS) of symmetric piecewise H(div)H(\mathrm{div})-tensors with normal-normal continuous traces on polyhedral meshes and a related displacement trace operator. Proposition 5 summarizes corresponding conformity and norm relations. We note that not all the results on the trace spaces are required to prove Proposition 5. We present them for their general relevance and possible future reference. Section 3 presents the mixed variational formulation of the model problem that defines the stress as an element of Hnn(div,𝒯;SS)H_{nn}(\operatorname{div},\mathcal{T};\SS). Theorem 6 establishes its Poisson-ratio robust well-posedness. Section 4 is devoted to the discrete setting and analysis. In §4.1 we introduce the stress element and disclose its degrees of freedom. The corresponding conforming piecewise polynomial approximation space Pnn2(𝒯)P^{2}_{nn}(\mathcal{T}) is the subject of §4.2, with approximation properties shown by Proposition 10. The resulting mixed finite element scheme is presented in §4.3, and shown to be locking free with expected convergence orders by Theorem 12. In Section 5 we present a numerical example that illustrates the locking-free convergence of our method.

Throughout, aba\lesssim b stands for aCba\leq Cb with a generic constant C>0C>0 that is independent of involved functions and mesh 𝒯\mathcal{T}. Relation aba\gtrsim b means that bab\lesssim a.

2 Analytical framework

In this section we introduce and analyze spaces and trace operators that provide the basis of our mixed formulation. The next subsection presents some notation and collects results for trace spaces of vector functions. In §2.2 we define the stress space of pointwise symmetric tensors with continuous normal-normal traces and establish duality relations with a trace space.

2.1 Notation, spaces and trace operators

For sub-domains and surfaces ωΩ¯\omega\subset\overline{\Omega} we use standard Lebesgue and Sobolev spaces Lq(ω;U)L_{q}(\omega;U) (q>1q>1) and Hs(ω;U)H^{s}(\omega;U) (s>0s>0) of UU-valued functions on ω\omega with U{,3,SS}U\in\{\mathbb{R},\mathbb{R}^{3},\SS\}. Here, SS3×3\SS\subset\mathbb{R}^{3\times 3} denotes the space of symmetric constant tensors. The generic L2(ω;U)L_{2}(\omega;U)-inner product and norm are (,)ω(\cdot\hskip 1.42262pt,\cdot)_{\omega} and ω=0,ω\|\cdot\|_{\omega}=\|\cdot\|_{0,\omega}, respectively. For integer ss, s,ω\|\cdot\|_{s,\omega} denotes the standard norm in Hs(ω;U)H^{s}(\omega;U), except for s=1s=1 and U=3U=\mathbb{R}^{3} in which case 1,ω:=(ω2+ε()ω2)1/2\|\cdot\|_{1,\omega}:=\Bigl{(}\|\cdot\|_{\omega}^{2}+\|\varepsilon(\cdot)\|_{\omega}^{2}\Bigr{)}^{1/2} with symmetric gradient ε():=(()+())/2\varepsilon(\cdot):=(\nabla(\cdot)+\nabla(\cdot)^{\top})/2. For non-integer s>0s>0, ω,s\|\cdot\|_{\omega,s} and ||ω,s|\cdot|_{\omega,s} denote the Sobolev-Slobodeckij norms and seminorms, cf. [1]. Given s,r0s,r\geq 0 and ωΩ\omega\subset\Omega, we need the space

Hs,r(div,ω;SS):={τHs(ω;SS);divτHr(ω;3}H^{s,r}(\operatorname{div},\omega;\SS):=\{\tau\in H^{s}(\omega;\SS);\;\operatorname{div}\tau\in H^{r}(\omega;\mathbb{R}^{3}\}

with row-wise application divτ\operatorname{div}\tau of the divergence operator, and denote H(div,ω;SS):=H(\operatorname{div},\omega;\SS):=H0,0(div,ω;SS)H^{0,0}(\operatorname{div},\omega;\SS) (identifying H0(Ω;U)H^{0}(\Omega;U) with L2(Ω;U)L_{2}(\Omega;U)). The latter space is provided with the graph norm div,ω:=(ω2+divω2)1/2\|\cdot\|_{\operatorname{div},\omega}:=\Bigl{(}\|\cdot\|_{\omega}^{2}+\|\operatorname{div}\cdot\|_{\omega}^{2}\Bigr{)}^{1/2}. We also need the topological dual H1(ω)H^{-1}(\omega) of H01(ω)H^{1}_{0}(\omega) with norm 1,ω\|\cdot\|_{-1,\omega}. We generally drop index ω\omega in the notation of norms and inner products when ω=Ω\omega=\Omega. The trace of a tensor τ\tau is tr(τ):=τ:id\operatorname{tr}(\tau):=\tau:\operatorname{\mathrm{id}} with identity tensor id\operatorname{\mathrm{id}} and Frobenius product “::”; the deviatoric part of τ\tau reads dev(τ):=τtr(τ)id/3\operatorname{dev}(\tau):=\tau-\operatorname{tr}(\tau)\operatorname{\mathrm{id}}/3.

We consider a regular decomposition 𝒯\mathcal{T} of Ω\Omega into shape-regular tetrahedra and formally denote by 𝒮:={T;T𝒯}\mathcal{S}:=\{\partial T;\;T\in\mathcal{T}\} its skeleton. We need the piecewise constant function h𝒯h_{\mathcal{T}} with h𝒯|T:=diam(T)h_{\mathcal{T}}|_{T}:=\operatorname{diam}(T) for T𝒯T\in\mathcal{T}. The set of faces of T𝒯T\in\mathcal{T} is (T)\mathcal{F}(T), and the sets of all (resp. interior) faces is :=T𝒯(T)\mathcal{F}:=\cup_{T\in\mathcal{T}}\mathcal{F}(T) (resp. (Ω)\mathcal{F}(\Omega)). In the following, ,K\langle{}\cdot\hskip 1.42262pt,\cdot\rangle_{K} denotes the generic L2L_{2}-duality on K{Γ}K\in\mathcal{F}\cup\{\Gamma\}. Decomposition 𝒯\mathcal{T} gives rise to product spaces, with corresponding notation where Ω\Omega is replaced with 𝒯\mathcal{T}, e.g., Hs(𝒯;U):=ΠT𝒯Hs(T;U)H^{s}(\mathcal{T};U):=\Pi_{T\in\mathcal{T}}H^{s}(T;U). Throughout, we identify functions of product spaces with piecewise defined functions. Mesh 𝒯\mathcal{T} induces the canonical trace operator

γ:\displaystyle\gamma:\; {H1(Ω;3)H(div,𝒯;SS),vγ(v),τ𝒮:=(ε(v),τ)+(v,divτ)𝒯\displaystyle\left\{\begin{array}[]{cll}H^{1}(\Omega;\mathbb{R}^{3})&\rightarrow&H(\operatorname{div},\mathcal{T};\SS)^{*},\\ v&\mapsto&\langle{}\gamma(v)\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}:=(\varepsilon(v)\hskip 1.42262pt,\tau)+(v\hskip 1.42262pt,\operatorname{div}\tau)_{\mathcal{T}}\end{array}\right.

with support on 𝒮\mathcal{S} and trace space

H1/2(𝒮;3):=γ(H1(Ω;3)).H^{1/2}(\mathcal{S};\mathbb{R}^{3}):=\gamma(H^{1}(\Omega;\mathbb{R}^{3})).

Here, (,)𝒯(\cdot\hskip 1.42262pt,\cdot)_{\mathcal{T}} is the generic notation for L2(𝒯;U)L_{2}(\mathcal{T};U) dualities, U{,3,SS}U\in\{\mathbb{R},\mathbb{R}^{3},\SS\}. There is an inherent duality between H1/2(𝒮;3)H^{1/2}(\mathcal{S};\mathbb{R}^{3}) and H(div,𝒯;SS)H(\operatorname{div},\mathcal{T};\SS),

φ,τ𝒮:=γ(v),τ𝒮(φH1/2(𝒮;3),τH(div,𝒯;SS)),\displaystyle\langle{}\varphi\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}:=\langle{}\gamma(v)\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}\quad(\varphi\in H^{1/2}(\mathcal{S};\mathbb{R}^{3}),\tau\in H(\operatorname{div},\mathcal{T};\SS)), (3)

where vv is any element of the pre-image γ1(φ)H1(Ω;3)\gamma^{-1}(\varphi)\subset H^{1}(\Omega;\mathbb{R}^{3}), φ=γ(v)\varphi=\gamma(v).

Given the tangential and normal trace operators

γt:\displaystyle\gamma_{t}:\; {H1(Ω;3)H1(𝒯;3),vγt(v),z𝒮:=(curlv,z)(v,curlz)𝒯,\displaystyle\left\{\begin{array}[]{cll}H^{1}(\Omega;\mathbb{R}^{3})&\rightarrow&H^{1}(\mathcal{T};\mathbb{R}^{3})^{*},\\ v&\mapsto&\langle{}\gamma_{t}(v)\hskip 1.42262pt,z\rangle_{\mathcal{S}}:=(\operatorname{curl}v\hskip 1.42262pt,z)-(v\hskip 1.42262pt,\operatorname{curl}z)_{\mathcal{T}},\end{array}\right.
γn:\displaystyle\gamma_{n}:\; {H1(Ω;3)H1(𝒯;3),vγn(v),z𝒮:=(divv,z)(v,z)𝒯,\displaystyle\left\{\begin{array}[]{cll}H^{1}(\Omega;\mathbb{R}^{3})&\rightarrow&H^{1}(\mathcal{T};\mathbb{R}^{3})^{*},\\ v&\mapsto&\langle{}\gamma_{n}(v)\hskip 1.42262pt,z\rangle_{\mathcal{S}}:=(\operatorname{div}v\hskip 1.42262pt,z)-(v\hskip 1.42262pt,\nabla z)_{\mathcal{T}},\end{array}\right.

we introduce their kernels Hn1(Ω,𝒮;3):=ker(γt)H^{1}_{n}(\Omega,\mathcal{S};\mathbb{R}^{3}):=\mathrm{ker}(\gamma_{t}) and Ht1(Ω,𝒮;3):=ker(γn).H^{1}_{t}(\Omega,\mathcal{S};\mathbb{R}^{3}):=\mathrm{ker}(\gamma_{n}). These are closed subspaces of H1(Ω;3)H^{1}(\Omega;\mathbb{R}^{3}) and give rise to the trace spaces

H~0,n1/2(𝒮;3)\displaystyle\widetilde{H}^{1/2}_{0,n}(\mathcal{S};\mathbb{R}^{3}) :=γ(Hn1(Ω,𝒮;3)H01(Ω;3)),\displaystyle:=\gamma(H^{1}_{n}(\Omega,\mathcal{S};\mathbb{R}^{3})\cap H^{1}_{0}(\Omega;\mathbb{R}^{3})), (4)
H~0,t1/2(𝒮;3)\displaystyle\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}) :=γ(Ht1(Ω,𝒮;3)H01(Ω;3))\displaystyle:=\gamma(H^{1}_{t}(\Omega,\mathcal{S};\mathbb{R}^{3})\cap H^{1}_{0}(\Omega;\mathbb{R}^{3}))

which consist of normal and tangential vector functions on 𝒮\mathcal{S}. To be more specific, let us introduce some further notation. For a function vH1(T;U)v\in H^{1}(T;U) (T𝒯T\in\mathcal{T}, U{,3}U\in\{\mathbb{R},\mathbb{R}^{3}\}) its traces on the boundary T\partial T and faces F(T)F\in\mathcal{F}(T) are denoted by v|TH1/2(T;U)v|_{\partial T}\in H^{1/2}(\partial T;U) and v|FH1/2(F;U)v|_{F}\in H^{1/2}(F;U), respectively. Restrictions of functions φH1/2(𝒮;3)\varphi\in H^{1/2}(\mathcal{S};\mathbb{R}^{3}) to subsets of 𝒮\mathcal{S} are defined through their extensions, e.g., φ|Γ:=v|Γ\varphi|_{\Gamma}:=v|_{\Gamma} for vH1(Ω;3)v\in H^{1}(\Omega;\mathbb{R}^{3}) with γ(v)=φ\gamma(v)=\varphi. We denote by H~1/2(F)H1/2(F)\widetilde{H}^{1/2}(F)\subset H^{1/2}(F) the restriction onto FF of the subspace of functions vH1/2(T)v\in H^{1/2}(\partial T) that vanish on TF¯\partial T\setminus\overline{F}, F(T)F\in\mathcal{F}(T), T𝒯T\in\mathcal{T}.

Proposition 1 (H~0,n1/2(𝒮;3)\widetilde{H}^{1/2}_{0,n}(\mathcal{S};\mathbb{R}^{3})).

Any function φH1/2(𝒮;3)\varphi\in H^{1/2}(\mathcal{S};\mathbb{R}^{3}) satisfies φH~0,n1/2(𝒮;3)\varphi\in\widetilde{H}^{1/2}_{0,n}(\mathcal{S};\mathbb{R}^{3}) if and only if πtφ|F=0\pi_{t}\varphi|_{F}=0, φn|FH~1/2(F)\varphi\cdot n|_{F}\in\widetilde{H}^{1/2}(F) for all F(Ω)F\in\mathcal{F}(\Omega), and φ|Γ=0\varphi|_{\Gamma}=0.

Proof.

Given φH1/2(𝒮;3)\varphi\in H^{1/2}(\mathcal{S};\mathbb{R}^{3}), πtφ|F=0\pi_{t}\varphi|_{F}=0 for all F(Ω)F\in\mathcal{F}(\Omega) and φ|Γ=0\varphi|_{\Gamma}=0 imply φH~0,n1/2(𝒮;3)\varphi\in\widetilde{H}^{1/2}_{0,n}(\mathcal{S};\mathbb{R}^{3}). To see the other direction consider vH1(Ω;3)v\in H^{1}(\Omega;\mathbb{R}^{3}) with φ=γ(v)\varphi=\gamma(v). For a tetrahedron T𝒯T\in\mathcal{T} let xjx_{j} denote its vertices with opposite faces FjF_{j} (j=1,,4j=1,\ldots,4). The corresponding barycentric coordinates and normal vectors are λj\lambda_{j} and njn_{j}. Assuming that the tangential traces of vv on faces FjF_{j} vanish we have to show that vnj|FjH~1/2(Fj)v\cdot n_{j}|_{F_{j}}\in\widetilde{H}^{1/2}(F_{j}) (j=1,,4j=1,\ldots,4).

It is enough to show that vn1|F1H~1/2(F1)v\cdot n_{1}|_{F_{1}}\in\widetilde{H}^{1/2}(F_{1}). We complement vector n1n_{1} with two linearly independent vectors t1,1,t1,2t_{1,1},t_{1,2} that are orthogonal to n1n_{1}. There are constants cj,dj,αjc_{j},d_{j},\alpha_{j}\in\mathbb{R}, αj0\alpha_{j}\not=0 (j=2,3,4j=2,3,4) such that

n1+cjt1,1+djt1,2=αjnj×nj′′,j=2,3,4,n_{1}+c_{j}t_{1,1}+d_{j}t_{1,2}=\alpha_{j}n_{j^{\prime}}\times n_{j^{\prime\prime}},\quad j=2,3,4,

with j,j′′{2,3,4}{j}j^{\prime},j^{\prime\prime}\in\{2,3,4\}\setminus\{j\}, j<j′′j^{\prime}<j^{\prime\prime}, that is, (j,j,j′′)(j,j^{\prime},j^{\prime\prime}) is a permutation of (2,3,4)(2,3,4). In fact, since {n1,t1,2,t1,2}\{n_{1},t_{1,2},t_{1,2}\} is a basis of 3\mathbb{R}^{3}, there are numbers ξj,ηj,ρj\xi_{j},\eta_{j},\rho_{j}\in\mathbb{R} such that ξjn1+ηjt1,1+ρjt1,2=nj×nj′′\xi_{j}n_{1}+\eta_{j}t_{1,1}+\rho_{j}t_{1,2}=n_{j^{\prime}}\times n_{j^{\prime\prime}} (j=2,3,4j=2,3,4). Note that ξj0\xi_{j}\not=0 because n1(nj×nj′′)0n_{1}\cdot(n_{j^{\prime}}\times n_{j^{\prime\prime}})\not=0 by the linear independence of n1,nj,nj′′n_{1},n_{j^{\prime}},n_{j^{\prime\prime}}. It follows that cj=ηj/ξjc_{j}=\eta_{j}/\xi_{j}, dj=ρj/ξjd_{j}=\rho_{j}/\xi_{j} and αj=1/ξj\alpha_{j}=1/\xi_{j} (j=2,3,4j=2,3,4). We continue to select

g:=j=24(n1+cjt1,1+djt1,2)λjg:=\sum_{j=2}^{4}(n_{1}+c_{j}t_{1,1}+d_{j}t_{1,2})\lambda_{j}

and define w:=gv|TH1(T)w:=g\cdot v|_{T}\in H^{1}(T). Since by assumption the tangential traces of vv on T\partial T vanish, ww satisfies

w|F1=j=24λj(n1+cjt1,1+djt1,2)v|F1=j=24λjn1v|F1=n1v|F1\displaystyle w|_{F_{1}}=\sum_{j=2}^{4}\lambda_{j}(n_{1}+c_{j}t_{1,1}+d_{j}t_{1,2})\cdot v|_{F_{1}}=\sum_{j=2}^{4}\lambda_{j}n_{1}\cdot v|_{F_{1}}=n_{1}\cdot v|_{F_{1}}

and, because λj|Fj=0\lambda_{j}|_{F_{j}}=0 and v|Fj=(vnj)nj|Fjv|_{F_{j}}=(v\cdot n_{j})n_{j}|_{F_{j}},

w|Fm=j=24λj(n1+cjt1,1+djt1,2)v|Fm=j{2,3,4}{m}λjαj(vnm)(nj×nj′′)nm|Fm=0\displaystyle w|_{F_{m}}=\sum_{j=2}^{4}\lambda_{j}(n_{1}+c_{j}t_{1,1}+d_{j}t_{1,2})\cdot v|_{F_{m}}=\sum_{j\in\{2,3,4\}\setminus\{m\}}\lambda_{j}\alpha_{j}(v\cdot n_{m})(n_{j^{\prime}}\times n_{j^{\prime\prime}})\cdot n_{m}|_{F_{m}}=0

for m=2,3,4m=2,3,4 by the orthogonality of nmn_{m} and nj×nj′′n_{j^{\prime}}\times n_{j^{\prime\prime}} (m=jm=j^{\prime} or m=j′′m=j^{\prime\prime} in the sum above). We conclude that w|TH1/2(T)w|_{\partial T}\in H^{1/2}(\partial T) is a zero extension of vn1|F1v\cdot n_{1}|_{F_{1}} to the other faces of TT, that is, φn1|F1=vn1|F1H~1/2(F1)\varphi\cdot n_{1}|_{F_{1}}=v\cdot n_{1}|_{F_{1}}\in\widetilde{H}^{1/2}(F_{1}). ∎

There is no corresponding characterization of H~0,t1/2(𝒮;3)\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}) as face-wise “H~1/2\widetilde{H}^{1/2}-space”. To see this, consider a tetrahedron TT with vertices xix_{i}, faces FiF_{i} opposite to xix_{i}, normal vectors nin_{i} on FiF_{i}, and barycentric coordinates λi\lambda_{i}, i=1,,4i=1,\ldots,4. Let e1,23{0}e_{1,2}\in\mathbb{R}^{3}\setminus\{0\} be a vector that is generated by the edge shared by F1F_{1} and F2F_{2}. Function v:=λ3λ4e1,2v:=\lambda_{3}\lambda_{4}e_{1,2} satisfies vH1(T;3)v\in H^{1}(T;\mathbb{R}^{3}), vni|Fi=0v\cdot n_{i}|_{F_{i}}=0 (i=1,2i=1,2) and v|Fi=0v|_{F_{i}}=0 (i=3,4i=3,4), but its tangential component πtv|F1=v|F1\pi_{t}v|_{F_{1}}=v|_{F_{1}} does not vanish on the shared edge. Therefore, πtv|F1\pi_{t}v|_{F_{1}} cannot be continuously extended by zero onto F2F_{2} in the trace space on T\partial T, which would be an appropriate characterization of the “H~1/2(F1;3)\widetilde{H}^{1/2}(F_{1};\mathbb{R}^{3})-property” that corresponds to the scalar case of normal components in H~1/2(F)\widetilde{H}^{1/2}(F) just considered. However, the edge-normal components of tangential face-traces of functions from H~0,t1/2(𝒮;3)\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}) do have such a property. A proper definition requires more notation.

For a simplex T𝒯T\in\mathcal{T} with face F(T)F\in\mathcal{F}(T) let nFn_{\partial F} denote the exterior unit normal vector along F\partial F that is normal to nn (a vector normal to FF). For an edge ee of FF, nF,en_{F,e} is the restriction of nFn_{\partial F} to ee. We define H~tn1/2(F;3)\widetilde{H}^{1/2}_{tn}(F;\mathbb{R}^{3}) as the subspace of functions φH1/2(F;3)\varphi\in H^{1/2}(F;\mathbb{R}^{3}) such that, for every edge ee of FF shared with a different face F(T)F^{\prime}\in\mathcal{F}(T), there exists a function vH1/2(T)v\in H^{1/2}(\partial T) with v|F=φ|FnF,ev|_{F}=\varphi|_{F}\cdot n_{F,e} and v|F=0v|_{F^{\prime}}=0. In other words, face-wise tangential traces have edge-normal components that, for every edge, can be continuously extended in H1/2(T;3)H^{1/2}(\partial T;\mathbb{R}^{3}) by zero across the edge to the neighboring face (suggesting the “tntn” index notation).

Proposition 2 (H~0,t1/2(𝒮;3)\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3})).

Any function φH1/2(𝒮;3)\varphi\in H^{1/2}(\mathcal{S};\mathbb{R}^{3}) satisfies φH~0,t1/2(𝒮;3)\varphi\in\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}) if and only if φn|F=0\varphi\cdot n|_{F}=0, φ|FH~tn1/2(F;3)\varphi|_{F}\in\widetilde{H}^{1/2}_{tn}(F;\mathbb{R}^{3}) for all F(Ω)F\in\mathcal{F}(\Omega), and φ|Γ=0\varphi|_{\Gamma}=0.

Proof.

The proof is similar to the proof of Proposition 1 and we merely discuss the non-trivial implication. Given φH~0,t1/2(𝒮;3)\varphi\in\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}), vH1(Ω;3)v\in H^{1}(\Omega;\mathbb{R}^{3}) with φ=γ(v)\varphi=\gamma(v), and T𝒯T\in\mathcal{T}, it is immediate that φn|T=0\varphi\cdot n|_{\partial T}=0, and we have to show that φ|FH~tn1/2(F;3)\varphi|_{F}\in\widetilde{H}^{1/2}_{tn}(F;\mathbb{R}^{3}) for F(T)F\in\mathcal{F}(T). Let F,F(T)F,F^{\prime}\in\mathcal{F}(T) be given with FFF\not=F^{\prime} and common edge ee. For the corresponding normal vectors n,nn,n^{\prime}, there are c,αc,\alpha\in\mathbb{R} such that nF,e+cn=αnn_{F,e}+cn=\alpha n^{\prime}. In fact, n,n,nF,en,n^{\prime},n_{F,e} are linearly dependent because they are orthogonal to edge ee and nF,en0n_{F,e}\cdot n^{\prime}\not=0. It follows that w:=(nF,e+cn)v|TH1(T)w:=(n_{F,e}+cn)\cdot v|_{T}\in H^{1}(T). Furthermore, w|F=φnF,e|Fw|_{F}=\varphi\cdot n_{F,e}|_{F} and w|F=αnφ|F=0w|_{F^{\prime}}=\alpha n^{\prime}\cdot\varphi|_{F^{\prime}}=0, that is, φ|FH~tn1/2(F,3)\varphi|_{F}\in\widetilde{H}^{1/2}_{tn}(F,\mathbb{R}^{3}) as wanted. ∎

We generalize the concept of “a point is too small” in two dimensions from Tartar [24, §17] to three dimensions. In two dimensions, it says that C0(Ω{x})C_{0}^{\infty}(\Omega\setminus\{x\}) for xΩ2x\in\Omega\subset\mathbb{R}^{2} is still dense in H01(Ω)H^{1}_{0}(\Omega).

Consider a finite set of pairwise distinct, straight lines L1,,LJL_{1},\ldots,L_{J} in 3\mathbb{R}^{3}, Lj={pj+sqj;s}L_{j}=\{p_{j}+sq_{j};\;s\in\mathbb{R}\} with pj,qj3p_{j},q_{j}\in\mathbb{R}^{3}, |qj|=1|q_{j}|=1, and LjΩL_{j}\cap\Omega\not=\emptyset, j=1,,Jj=1,\ldots,J. (Here, Ω3\Omega\subset\mathbb{R}^{3} can be any bounded open set.) The set of lines L:=L1LJL:=L_{1}\cup\ldots\cup L_{J} is “too small”, in the following sense.

Proposition 3 (density).

The set C0(ΩL)C_{0}^{\infty}(\Omega\setminus L) is dense in H01(Ω)H^{1}_{0}(\Omega).

Proof.

Without loss of generality we assume that diam(Ω)1\operatorname{diam}(\Omega)\leq 1 (otherwise we scale Ω\Omega). Let fH01(Ω)f\in H^{1}_{0}(\Omega) and ϵ>0\epsilon>0 be given. We approximate ff in several steps.

1. Since C0(Ω)H01(Ω)C^{\infty}_{0}(\Omega)\subset H^{1}_{0}(\Omega) is dense we find f0C0(Ω)f_{0}\in C^{\infty}_{0}(\Omega) with ff01<ϵ/5\|f-f_{0}\|_{1}<\epsilon/5.

2. Let us assume that |f0|1|f_{0}|\leq 1 in Ω\Omega (for the other case see the end of the proof). We approximate separately the non-negative and non-positive parts f+:=max{0,f0}f_{+}:=\max\{0,f_{0}\} and f:=f+f0f_{-}:=f_{+}-f_{0} of f0f_{0}. Note that f±H01(Ω)f_{\pm}\in H^{1}_{0}(\Omega), cf. [10, Theorem 4.4(iii)], and 0f±10\leq f_{\pm}\leq 1.

3. The function loglog(1/||)\log\log(1/|\cdot|) is a standard example of an unbounded H1H^{1}-function on sufficiently small two-dimensional neighborhoods of the origin. We consider a parameter 0<δ<10<\delta<1 to be selected later, and define

χ(r):=(log(elogδ)loglog(1/r))+for0<r<1,\chi(r):=\Bigl{(}\log(-e\log\delta)-\log\log(1/r)\Bigr{)}_{+}\quad\text{for}\quad 0<r<1,

with non-negative part ()+(\cdot)_{+}. It is elementary that χ\chi is monotonically increasing in (0,1)(0,1), χ=0\chi=0 in [0,δe][0,\delta^{e}], 0<χ<10<\chi<1 in (δe,δ)(\delta^{e},\delta), χ(δ)=1\chi(\delta)=1, and χ>1\chi>1 in (δ,1)(\delta,1); ee is Euler’s number and δe\delta^{e} means δ\delta raised to the power ee. We continue to define

gj(x):=min{1,χ(dist(x,Lj))}(xΩ,j=1,,J).g_{j}(x):=\min\{1,\chi(\operatorname{dist}(x,L_{j}))\}\quad(x\in\Omega,j=1,\ldots,J).

Transformation to cylindrical coordinates along LjL_{j} reveals

gj1,Ujπδ22π/logδforUj:={xΩ;dist(x,Lj)<δ}={xΩ;gj(x)<1}\|g_{j}\|_{1,U_{j}}\leq\sqrt{\pi\delta^{2}-2\pi/\log\delta}\quad\text{for}\quad U_{j}:=\{x\in\Omega;\;\operatorname{dist}(x,L_{j})<\delta\}=\{x\in\Omega;\;g_{j}(x)<1\}

(j=1,,Jj=1,\ldots,J) where we used that diam(Ω)1\operatorname{diam}(\Omega)\leq 1. We denote U:=U1UJ={xΩ;dist(x,L)<δ}U:=U_{1}\cup\ldots\cup U_{J}=\{x\in\Omega;\;\operatorname{dist}(x,L)<\delta\} and conclude for j=1,,Jj=1,\ldots,J that

gj1,U2k=1Jgj1,Uk2J(πδ22π/logδ)+kj|Uk|0asδ0for.\displaystyle\|g_{j}\|_{1,U}^{2}\leq\sum_{k=1}^{J}\|g_{j}\|_{1,U_{k}}^{2}\leq J(\pi\delta^{2}-2\pi/\log\delta)+\sum_{k\not=j}|U_{k}|\to 0\quad\text{as}\ \delta\to 0\ \text{for}. (5)

4. We define preliminary approximations of f±f_{\pm} by g±:=min{f±,g1,,gJ}g_{\pm}:=\min\{f_{\pm},g_{1},\ldots,g_{J}\} on Ω\Omega. They vanish in a neighborhood of LΩL\cap\Omega, dist(suppg±,L)>0\operatorname{dist}(\operatorname{supp}g_{\pm},L)>0, satisfy g±H01(Ω)g_{\pm}\in H^{1}_{0}(\Omega) and

g1(x)==gJ(x)=1f±(x)=g±(x)atxΩwithdist(x,L)>δ.g_{1}(x)=\ldots=g_{J}(x)=1\geq f_{\pm}(x)=g_{\pm}(x)\quad\text{at}\ x\in\Omega\ \text{with}\ \operatorname{dist}(x,L)>\delta.

Thus f±g±1=f±g±1,U.\|f_{\pm}-g_{\pm}\|_{1}=\|f_{\pm}-g_{\pm}\|_{1,U}. By the triangle inequality and the coarse bounds 0g±f±+g1++gJ0\leq g_{\pm}\leq f_{\pm}+g_{1}+\ldots+g_{J}, |g±||f±|+|g1|++|gJ||\nabla g_{\pm}|\leq|\nabla f_{\pm}|+|\nabla g_{1}|+\ldots+|\nabla g_{J}| (almost everywhere in UU) we find with (5) that

12f±g±1,U2\displaystyle\frac{1}{2}\|f_{\pm}-g_{\pm}\|_{1,U}^{2} f±1,U2+f±+g1++gJU2+|f±|+|g1|+|gJ|U2\displaystyle\leq\|f_{\pm}\|_{1,U}^{2}+\|f_{\pm}+g_{1}+\ldots+g_{J}\|_{U}^{2}+\||\nabla f_{\pm}|+|\nabla g_{1}|+\ldots|\nabla g_{J}|\|_{U}^{2}
f±1,U2+(J+1)(f±1,U2+g11,U2++gJ1,U2)\displaystyle\leq\|f_{\pm}\|_{1,U}^{2}+(J+1)\bigl{(}\|f_{\pm}\|_{1,U}^{2}+\|g_{1}\|_{1,U}^{2}+\ldots+\|g_{J}\|_{1,U}^{2}\bigr{)}
=(J+2)f±1,U2+(J+1)(g11,U2++gJ1,U2)0(δ0)\displaystyle=(J+2)\|f_{\pm}\|_{1,U}^{2}+(J+1)\Bigl{(}\|g_{1}\|_{1,U}^{2}+\ldots+\|g_{J}\|_{1,U}^{2}\Bigr{)}\to 0\quad(\delta\to 0)

because f±H1(Ω)f_{\pm}\in H^{1}(\Omega) and |U|0|U|\to 0 as δ0\delta\to 0. Therefore, f±g±1<ϵ/5\|f_{\pm}-g_{\pm}\|_{1}<\epsilon/5 for some fixed, sufficiently small δ>0\delta>0. By standard mollification of g±g_{\pm} we find g~±C0(ΩL)\widetilde{g}_{\pm}\in C^{\infty}_{0}(\Omega\setminus L) with g±g~±1ϵ/5\|g_{\pm}-\widetilde{g}_{\pm}\|_{1}\leq\epsilon/5.

A combination of steps 1–4 shows that g~:=g~+g~\widetilde{g}:=\widetilde{g}_{+}-\widetilde{g}_{-} satisfies g~C0(ΩL)\widetilde{g}\in C^{\infty}_{0}(\Omega\setminus L) and

fg~1ff01+f+g+1+fg1+g+g~+1+gg~1ϵ.\|f-\widetilde{g}\|_{1}\leq\|f-f_{0}\|_{1}+\|f_{+}-g_{+}\|_{1}+\|f_{-}-g_{-}\|_{1}+\|g_{+}-\widetilde{g}_{+}\|_{1}+\|g_{-}-\widetilde{g}_{-}\|_{1}\leq\epsilon.

This finishes the proof for the case f01\|f_{0}\|_{\infty}\leq 1. Otherwise we repeat steps 2 and 4 where f0f_{0} is replaced with f~0:=f0/c0\widetilde{f}_{0}:=f_{0}/c_{0} for c0:=f0<c_{0}:=\|f_{0}\|_{\infty}<\infty, and where bound ϵ/5\epsilon/5 is reduced to ϵ/(5c0)\epsilon/(5c_{0}). In particular, f~0=f+f\widetilde{f}_{0}=f_{+}-f_{-} and g~:=c0(g~+g~)\widetilde{g}:=c_{0}(\widetilde{g}_{+}-\widetilde{g}_{-}) satisfies g~C0(ΩL)\widetilde{g}\in C^{\infty}_{0}(\Omega\setminus L) and

fg~1ff01+c0(f+g+1+fg1+g+g~+1+gg~1)ϵ.\|f-\widetilde{g}\|_{1}\leq\|f-f_{0}\|_{1}+c_{0}\bigl{(}\|f_{+}-g_{+}\|_{1}+\|f_{-}-g_{-}\|_{1}+\|g_{+}-\widetilde{g}_{+}\|_{1}+\|g_{-}-\widetilde{g}_{-}\|_{1}\Bigr{)}\leq\epsilon.

Proposition 4 (inclusion).

The inclusion

H~0,n1/2(𝒮;3)H~0,t1/2(𝒮;3)\displaystyle\widetilde{H}^{1/2}_{0,n}(\mathcal{S};\mathbb{R}^{3})\oplus\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}) H01/2(𝒮;3)\displaystyle\subset H^{1/2}_{0}(\mathcal{S};\mathbb{R}^{3})

is dense. It is strict if there is a face F(Ω)F\in\mathcal{F}(\Omega) that does not touch the boundary. Moreover, every τH(div,𝒯;SS)\tau\in H(\operatorname{div},\mathcal{T};\SS) satisfies

τH(div,Ω;SS)ρ,τ𝒮=0ρH~0,n1/2(𝒮;3)H~0,t1/2(𝒮;3).\tau\in H(\operatorname{div},\Omega;\SS)\quad\Leftrightarrow\quad\langle{}\rho\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}=0\quad\forall\rho\in\widetilde{H}^{1/2}_{0,n}(\mathcal{S};\mathbb{R}^{3})\oplus\widetilde{H}^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}).
Proof.

The proof is analogous to the proof of [7, Proposition 2, (3)] by replacing density relation [24, Lemma 17.3] with Proposition 3. ∎

2.2 Normal-normal continuous stress space and dual trace operator

Duality (3) and normal trace space (4) provide us with the definition of a space of pointwise symmetric stresses with continuous normal-normal components across 𝒮\mathcal{S},

Hnn(div,𝒯;SS):={τH(div,𝒯;SS);φ,τ𝒮=0φH~0,n1/2(𝒮;3)}.H_{nn}(\operatorname{div},\mathcal{T};\SS):=\{\tau\in H(\operatorname{div},\mathcal{T};\SS);\;\langle{}\varphi\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}=0\ \forall\varphi\in\widetilde{H}^{1/2}_{0,n}(\mathcal{S};\mathbb{R}^{3})\}.

It is a closed subspace of H(div,𝒯;SS)H(\operatorname{div},\mathcal{T};\SS). The duality with this constrained space reduces the canonical trace operator γ\gamma to

γ1,𝑛𝑛:\displaystyle{\gamma_{1,\mathit{nn}}}:\; {H1(Ω;3)Hnn(div,𝒯;SS),vγ1,𝑛𝑛(v),τ𝒮:=(ε(v),τ)+(v,divτ)𝒯.\displaystyle\left\{\begin{array}[]{cll}H^{1}(\Omega;\mathbb{R}^{3})&\rightarrow&H_{nn}(\operatorname{div},\mathcal{T};\SS)^{*},\\ v&\mapsto&\langle{}{\gamma_{1,\mathit{nn}}}(v)\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}:=(\varepsilon(v)\hskip 1.42262pt,\tau)+(v\hskip 1.42262pt,\operatorname{div}\tau)_{\mathcal{T}}.\end{array}\right.

It delivers tangential traces on interior faces and canonical traces on faces FΓF\subset\Gamma. The resulting trace spaces (the latter one including homogeneous Dirichlet data) are

Ht1/2(𝒮;3)\displaystyle H^{1/2}_{t}(\mathcal{S};\mathbb{R}^{3}) :=γ1,𝑛𝑛(H1(Ω;3))andH0,t1/2(𝒮;3):=γ1,𝑛𝑛(H01(Ω;3)).\displaystyle:={\gamma_{1,\mathit{nn}}}\bigl{(}H^{1}(\Omega;\mathbb{R}^{3})\bigr{)}\quad\text{and}\quad H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}):={\gamma_{1,\mathit{nn}}}\bigl{(}H^{1}_{0}(\Omega;\mathbb{R}^{3})\bigr{)}.

It remains to specify norms. For an element τ\tau of the product space H(div,𝒯;SS)H(\operatorname{div},\mathcal{T};\SS) we use the product norm τdiv,𝒯:=(T𝒯τdiv,T2)1/2\|\tau\|_{\operatorname{div},\mathcal{T}}:=\Bigl{(}\sum_{T\in\mathcal{T}}\|\tau\|_{\operatorname{div},T}^{2}\Bigr{)}^{1/2}. Trace space Ht1/2(𝒮;3)H^{1/2}_{t}(\mathcal{S};\mathbb{R}^{3}) is endowed with the canonical trace norm

ρ1/2,t,𝒮:=inf{v1;vH1(Ω;3),γ1,𝑛𝑛(v)=ρ},ρHt1/2(𝒮;3).\|\rho\|_{1/2,t,\mathcal{S}}:=\mathrm{inf}\{\|v\|_{1};\;v\in H^{1}(\Omega;\mathbb{R}^{3}),\ {\gamma_{1,\mathit{nn}}}(v)=\rho\},\quad\rho\in H^{1/2}_{t}(\mathcal{S};\mathbb{R}^{3}).

The following conformity and norm relations are essential to develop and analyze our mixed formulation of (2).

Proposition 5 (duality).

(i) Any τHnn(div,𝒯;SS)\tau\in H_{nn}(\operatorname{div},\mathcal{T};\SS) satisfies

τH(div,Ω;SS)ρ,τ𝒮=0ρH0,t1/2(𝒮;3).\tau\in H(\operatorname{div},\Omega;\SS)\quad\Leftrightarrow\quad\langle{}\rho\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}=0\quad\forall\rho\in H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}).

(ii) Any ρHt1/2(𝒮;3)\rho\in H^{1/2}_{t}(\mathcal{S};\mathbb{R}^{3}) satisfies

ρ1/2,t,𝒮=sup0τHnn(div,𝒯;SS),(tr(τ),1)=0ρ,τ𝒮τdiv,𝒯.\|\rho\|_{1/2,t,\mathcal{S}}=\sup_{0\not=\tau\in H_{nn}(\operatorname{div},\mathcal{T};\SS),\;(\operatorname{tr}(\tau)\hskip 1.42262pt,1)=0}\frac{\langle{}\rho\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}}{\|\tau\|_{\operatorname{div},\mathcal{T}}}.
Proof.

In two dimensions, statement (i) is [7, Lemma 5]. Its proof is based on a decomposition of (a dense subspace of) H1/2(𝒮;2)H^{1/2}(\mathcal{S};\mathbb{R}^{2}) into normal and tangential trace spaces. Proposition 4 does provide such a decomposition in three dimensions and allows to apply the same technique as in [7]. For an abstract setting of this result see [14, Lemma 27]. Also statement (ii) is proved analogously to the two-dimensional case [7, Proposition 6], and for an abstract form (without the trace constraint for test functions), see [14, Lemma 26]. ∎

3 Mixed formulation

An application of trace operator γ1,nn\gamma_{1,nn} and the conformity characterization given by Proposition 5(i) lead to the following mixed formulation of (2). Find σHnn(div,𝒯;SS)\sigma\in H_{nn}(\operatorname{div},\mathcal{T};\SS), uL2(Ω;3)u\in L_{2}(\Omega;\mathbb{R}^{3}), and ηH0,t1/2(𝒮;3)\eta\in H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}) such that

(𝔸σ,δσ)+(u,divδσ)𝒯η,δσ𝒮\displaystyle(\mathbb{A}\sigma\hskip 1.42262pt,{\delta\!\sigma})+(u\hskip 1.42262pt,\operatorname{div}{\delta\!\sigma})_{\mathcal{T}}-\langle{}\eta\hskip 1.42262pt,{\delta\!\sigma}\rangle_{\mathcal{S}} =0,\displaystyle=0, (6a)
(divσ,δu)𝒯δη,σ𝒮\displaystyle(\operatorname{div}\sigma\hskip 1.42262pt,{\delta\!u})_{\mathcal{T}}-\langle{}{\delta\!\eta}\hskip 1.42262pt,\sigma\rangle_{\mathcal{S}} =(f,δu)\displaystyle=-(f\hskip 1.42262pt,{\delta\!u}) (6b)

for any δσHnn(div,𝒯;SS){\delta\!\sigma}\in H_{nn}(\operatorname{div},\mathcal{T};\SS), δuL2(Ω;3){\delta\!u}\in L_{2}(\Omega;\mathbb{R}^{3}), and δηH0,t1/2(𝒮;3){\delta\!\eta}\in H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}).

We show its well-posedness.

Theorem 6 (well-posedness of mixed formulation).

Given fL2(Ω;3)f\in L_{2}(\Omega;\mathbb{R}^{3}), problem (6) is well posed with a unique solution (σ,u,η)(\sigma,u,\eta). The couple (u,σ)H1(Ω)×H(div,Ω;SS)(u,\sigma)\in H^{1}(\Omega)\times H(\operatorname{div},\Omega;\SS) solves (2), η=γ1,𝑛𝑛(u)\eta={\gamma_{1,\mathit{nn}}}(u), and

σdiv,𝒯+u+η1/2,t,𝒮\displaystyle\|\sigma\|_{\operatorname{div},\mathcal{T}}+\|u\|+\|\eta\|_{1/2,t,\mathcal{S}} f\displaystyle\lesssim\|f\|

holds uniformly with respect to λ(0,)\lambda\in(0,\infty) and 𝒯\mathcal{T}.

Proof.

The proof uses the standard ingredients of mixed formulations, and is analogous to the two-dimensional case studied in [7, Theorem 10]. We recall the main steps and give some more details on the uniform stability.

We need the representation

(𝔸τ,δτ)=12μ(devτ,devδτ)+13(3λ+2μ)(tr(τ),tr(δτ))τ,δτL2(Ω;SS),\displaystyle(\mathbb{A}\tau\hskip 1.42262pt,{\delta\!\tau})=\frac{1}{2\mu}(\operatorname{dev}\tau\hskip 1.42262pt,\operatorname{dev}{\delta\!\tau})+\frac{1}{3(3\lambda+2\mu)}(\operatorname{tr}(\tau)\hskip 1.42262pt,\operatorname{tr}({\delta\!\tau}))\quad\forall\tau,{\delta\!\tau}\in L_{2}(\Omega;\SS), (7)

see, e.g., [5, (9.1.8)]. The selection of δσ=id{\delta\!\sigma}=\operatorname{\mathrm{id}} in (6a) and relation η,id𝒮=η,nΓ=0\langle{}\eta\hskip 1.42262pt,\operatorname{\mathrm{id}}\rangle_{\mathcal{S}}=\langle{}\eta\hskip 1.42262pt,n\rangle_{\Gamma}=0 for ηH0,t1/2(𝒮;3)\eta\in H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}) show that (tr(σ),1)=0(\operatorname{tr}(\sigma)\hskip 1.42262pt,1)=0. We conclude that formulation (6) is equivalent to the same system where Hnn(div,𝒯;SS)H_{nn}(\operatorname{div},\mathcal{T};\SS) is replaced with

H~nn(div,𝒯;SS):={τHnn(div,𝒯;SS);(tr(τ),1)=0}.\widetilde{H}_{nn}(\operatorname{div},\mathcal{T};\SS):=\{\tau\in H_{nn}(\operatorname{div},\mathcal{T};\SS);\;(\operatorname{tr}(\tau)\hskip 1.42262pt,1)=0\}.

We continue to analyze this formulation.

The bilinear and linear forms of (6) are uniformly bounded by the selection of norms. Inf-sup properties

sup0τH(div,Ω;SS)(divτ,δu)τdivδuδuL2(Ω;3)\displaystyle\sup_{0\not=\tau\in H(\operatorname{div},\Omega;\SS)}\frac{(\operatorname{div}\tau\hskip 1.42262pt,{\delta\!u})}{\|\tau\|_{\operatorname{div}}}\gtrsim\|{\delta\!u}\|\quad\forall{\delta\!u}\in L_{2}(\Omega;\mathbb{R}^{3})

and

sup0τH~nn(div,𝒯;SS)δη,τ𝒮τdiv,𝒯δη1/2,t,𝒮δηH0,t1/2(𝒮;3)\displaystyle\sup_{0\not=\tau\in\widetilde{H}_{nn}(\operatorname{div},\mathcal{T};\SS)}\frac{\langle{}{\delta\!\eta}\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}}{\|\tau\|_{\operatorname{div},\mathcal{T}}}\geq\|{\delta\!\eta}\|_{1/2,t,\mathcal{S}}\quad\forall{\delta\!\eta}\in H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3})

hold by surjectivity of div:H(div,Ω;SS)L2(Ω;3)\operatorname{div}:\;H(\operatorname{div},\Omega;\SS)\to L_{2}(\Omega;\mathbb{R}^{3}) and Proposition 5(ii), respectively. The conformity relation of Proposition 5(i) and [6, Theorem 3.3] then show the required combined inf-sup relation

sup0τH~nn(div,𝒯;SS)(divτ,δu)𝒯δη,τ𝒮τdiv,𝒯δu+δη1/2,t,𝒮\displaystyle\sup_{0\not=\tau\in\widetilde{H}_{nn}(\operatorname{div},\mathcal{T};\SS)}\frac{(\operatorname{div}\tau\hskip 1.42262pt,{\delta\!u})_{\mathcal{T}}-\langle{}{\delta\!\eta}\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}}{\|\tau\|_{\operatorname{div},\mathcal{T}}}\gtrsim\|{\delta\!u}\|+\|{\delta\!\eta}\|_{1/2,t,\mathcal{S}}\quad

for any δuL2(Ω;3){\delta\!u}\in L_{2}(\Omega;\mathbb{R}^{3}) and δηH0,t1/2(𝒮;3){\delta\!\eta}\in H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3}).

For the uniform well-posedness of (6) it remains to verify the uniform coercivity of bilinear form (𝔸,)(\mathbb{A}\cdot\hskip 1.42262pt,\cdot) on the kernel

K0:=\displaystyle K_{0}:= {τH~nn(div,𝒯;SS);(divτ,δu)𝒯=0δuL2(Ω;3),δη,τ𝒮=0δηH0,t1/2(𝒮;3)}\displaystyle\{\tau\in\widetilde{H}_{nn}(\operatorname{div},\mathcal{T};\SS);\;(\operatorname{div}\tau\hskip 1.42262pt,{\delta\!u})_{\mathcal{T}}=0\ \forall{\delta\!u}\in L_{2}(\Omega;\mathbb{R}^{3}),\;\langle{}{\delta\!\eta}\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}=0\ \forall{\delta\!\eta}\in H^{1/2}_{0,t}(\mathcal{S};\mathbb{R}^{3})\}
=\displaystyle= {τH(div,Ω;SS);divτ=0,(tr(τ),1)=0}.\displaystyle\{\tau\in H(\operatorname{div},\Omega;\SS);\;\operatorname{div}\tau=0,\ (\operatorname{tr}(\tau)\hskip 1.42262pt,1)=0\}.

The latter identity is due to Proposition 5(i). Critical is the bound

tr(τ)C(devτ+divτ1)=CdevττK0\|\operatorname{tr}(\tau)\|\leq C\bigl{(}\|\operatorname{dev}\tau\|+\|\operatorname{div}\tau\|_{-1}\bigr{)}=C\|\operatorname{dev}\tau\|\quad\forall\tau\in K_{0}

for a constant C>0C>0, see [8, Theorem 1]. The uniform coercivity of (𝔸,)(\mathbb{A}\cdot\hskip 1.42262pt,\cdot) on K0K_{0} then follows from (7). The fact that uH1(Ω;3)u\in H^{1}(\Omega;\mathbb{R}^{3}) and σH(div,Ω;SS)\sigma\in H(\operatorname{div},\Omega;\SS) solve (2) is immediate, and η=γ1,𝑛𝑛(u)\eta={\gamma_{1,\mathit{nn}}}(u) holds by (6a) and the definition of γ1,𝑛𝑛{\gamma_{1,\mathit{nn}}}. ∎

4 Finite element approximation

Let us introduce some polynomial related notation. For K𝒯K\in\mathcal{F}\cup\mathcal{T} and U{,3,SS}U\in\{\mathbb{R},\mathbb{R}^{3},\SS\} we denote by Pk(K;U)P^{k}(K;U) the space of polynomials of degree kk on KK with values in UU. The piecewise polynomial spaces are Pk(𝒯;U)P^{k}(\mathcal{T};U). We drop UU when U=U=\mathbb{R}.

4.1 An H(div;SS)H(\operatorname{div};\SS) element

Let T3T\subset\mathbb{R}^{3} be a tetrahedron with vertices xix_{i}, barycentric coordinates λi\lambda_{i}, faces FiF_{i} opposite to xix_{i}, and unit exterior normal vectors nin_{i} on FiF_{i} (we also employ the generic notation nn), and denote n~i:=λispan{ni}\widetilde{n}_{i}:=\nabla\lambda_{i}\in\mathrm{span}\{n_{i}\}, i=1,,4i=1,\ldots,4. We also need unit vectors ei,je_{i,j} (in a certain direction) generated by the edges F¯iF¯j\overline{F}_{i}\cap\overline{F}_{j}, ij=1,,4i\not=j=1,\ldots,4. All these simplex-related objects are numbered modulus 44, e.g., x5=x1x_{5}=x_{1}. We abbreviate ab:=(ab+ba)/2SSa\odot b:=(ab^{\top}+ba^{\top})/2\in\SS for vectors a,b3a,b\in\mathbb{R}^{3}.

The construction of our H(div,T;SS)H(\operatorname{div},T;\SS)-element is based upon the constant tensors

Ti:=ei+1,i+2ei+1,i+3niei+1,i+2niei+1,i+3(i=1,,4),T5:=e1,2e3,4,T6:=e1,3e2,4T_{i}:=\frac{e_{i+1,i+2}\odot e_{i+1,i+3}}{n_{i}\cdot e_{i+1,i+2}\;n_{i}\cdot e_{i+1,i+3}}\quad(i=1,\ldots,4),\quad T_{5}:=e_{1,2}\odot e_{3,4},\quad T_{6}:=e_{1,3}\odot e_{2,4}

which from a basis of SS\SS (as seen below in Lemma 7), and the polynomial tensor spaces

Pnn0(T;SS)\displaystyle P^{0}_{nn}(T;\SS) :=span{T1,,T4},P00(T;SS):=span{T5,T6},\displaystyle:=\mathrm{span}\bigl{\{}T_{1},\ldots,T_{4}\bigr{\}},\quad P^{0}_{0}(T;\SS):=\mathrm{span}\bigl{\{}T_{5},T_{6}\bigr{\}},
P2,k(T,SS)\displaystyle P^{2,k}(T,\SS) :={τP2(T;SS);nτn|FPk(F),F(T)},k{0,1},\displaystyle:=\{\tau\in P^{2}(T;\SS);\;n\cdot\tau n|_{F}\in P^{k}(F),\ F\in\mathcal{F}(T)\},\quad k\in\{0,1\},
P02(T;SS)\displaystyle P^{2}_{0}(T;\SS) :={τP2(T;SS);nτn|T=0}.\displaystyle:=\{\tau\in P^{2}(T;\SS);\;n\cdot\tau n|_{\partial T}=0\}.

Remark. The construction of TiSST_{i}\in\SS is such that Tini+1=0T_{i}n_{i+1}=0, i=1,,4i=1,\ldots,4, that is, the image of mapping Ti:33T_{i}:\;\mathbb{R}^{3}\to\mathbb{R}^{3} is the plane through the origin that is parallel to face Fi+1F_{i+1}. Furthermore, njTinj=0n_{j}\cdot T_{i}n_{j}=0 for j=1,,4j=1,\ldots,4, i=5,6i=5,6, and we will see that njTinj=δijn_{j}\cdot T_{i}n_{j}=\delta_{ij} for i,j=1,,4i,j=1,\ldots,4.

Our H(div,T;SS)H(\operatorname{div},T;\SS) finite element on the tetrahedron TT is the vector space

𝕏(T):=(P2(T)Pnn0(T;SS)P2,1(T,SS))P00(T;SS)\mathbb{X}(T):=\bigl{(}P^{2}(T)P^{0}_{nn}(T;\SS)\cap P^{2,1}(T,\SS)\bigr{)}\oplus P^{0}_{0}(T;\SS) (8)

and the following degrees of freedom,

|F|nτn,qF\displaystyle|F|\langle{}n\cdot\tau n\hskip 1.42262pt,q\rangle_{F}\quad (qP1(F),F(T)),\displaystyle(q\in P^{1}(F),\ F\in\mathcal{F}(T)), (9a)
(τ,c)T\displaystyle(\tau\hskip 1.42262pt,c)_{T} (cP0(T;SS)),\displaystyle(c\in P^{0}(T;\SS)), (9b)
(divτ,v)T\displaystyle(\operatorname{div}\tau\hskip 1.42262pt,v)_{T} (vP1(T;3)).\displaystyle(v\in P^{1}(T;\mathbb{R}^{3})). (9c)

The proof that (TT,𝕏(T)\mathbb{X}(T),(9)) is a finite element requires some preparation.

Lemma 7.

We have P0(T;SS)=Pnn0(T;SS)P00(T;SS).P^{0}(T;\SS)=P^{0}_{nn}(T;\SS)\oplus P^{0}_{0}(T;\SS).

Proof.

We start by proving that {T1,,T6}\{T_{1},\ldots,T_{6}\} is a basis of P0(T;SS)P^{0}(T;\SS). Since dimP0(T;SS)=6\dim P^{0}(T;\SS)=6 and TiP0(T;SS)T_{i}\in P^{0}(T;\SS), i=1,,6i=1,\ldots,6, it suffices to show that the tensors TiT_{i} are linearly independent.

For every i{1,,4}i\in\{1,\ldots,4\}, at least one of the two vectors ei+1,i+2e_{i+1,i+2} and ei+1,i+3e_{i+1,i+3} is tangential to the faces FjF_{j} for j{i+1,i+2,i+3}j\in\{i+1,i+2,i+3\}. Furthermore, both vectors are not tangential to FiF_{i}. Consequently

(niei+1,i+2)(niei+2,i+3)0and(njei+1,i+2)(njei+1,i+3)=0forji.(n_{i}\cdot e_{i+1,i+2})(n_{i}\cdot e_{i+2,i+3})\not=0\quad\text{and}\quad(n_{j}\cdot e_{i+1,i+2})(n_{j}\cdot e_{i+1,i+3})=0\quad\text{for}\ j\not=i.

Therefore, tensors TiT_{i} are well defined and satisfy

njTinj=δijfori,j=1,,4.n_{j}\cdot T_{i}n_{j}=\delta_{ij}\quad\text{for}\ i,j=1,\ldots,4.

On the other hand, for each face FiF_{i}, one of the vectors e1,2,e3,4e_{1,2},e_{3,4} and one of the vectors e1,3,e2,4e_{1,3},e_{2,4} is tangential to FiF_{i}. This implies that

njTinj=0fori=5,6,j=1,,4.n_{j}\cdot T_{i}n_{j}=0\quad\text{for}\ i=5,6,\ j=1,\ldots,4.

It follows that T1,,T6T_{1},\ldots,T_{6} are linearly independent, and thus form a basis of P0(T;SS)P^{0}(T;\SS), if T5T_{5} and T6T_{6} are linearly independent. The latter is true because

n3T5n2=n3e1,2e3,4n20butn3T6n2=n3e1,3e2,4n2=0.n_{3}\cdot T_{5}n_{2}=n_{3}\cdot e_{1,2}\;e_{3,4}\cdot n_{2}\not=0\quad\text{but}\quad n_{3}\cdot T_{6}n_{2}=n_{3}\cdot e_{1,3}\;e_{2,4}\cdot n_{2}=0.

The splitting of P0(T;SS)P^{0}(T;\SS) then holds by definition of the two subspaces. ∎

We use a transformation of T𝒯T\in\mathcal{T} to a reference tetrahedron T^{\widehat{T}} to show that (9) are degrees of freedom of 𝕏(T)\mathbb{X}(T). We consider the tetrahedron T^{\widehat{T}} with vertices x^1=(0,0,0){\widehat{x}}_{1}=(0,0,0), x^2=(1,0,0){\widehat{x}}_{2}=(1,0,0), x^3=(0,1,0){\widehat{x}}_{3}=(0,1,0), x^4=(0,0,1){\widehat{x}}_{4}=(0,0,1), and let GT:T^TG_{T}:\;{\widehat{T}}\to T denote the affine diffeomorphism

(x^y^z^)aT+BT(x^y^z^)\begin{pmatrix}{\widehat{x}}\\ {\widehat{y}}\\ {\widehat{z}}\end{pmatrix}\mapsto a_{T}+B_{T}\begin{pmatrix}{\widehat{x}}\\ {\widehat{y}}\\ {\widehat{z}}\end{pmatrix}

with aT3a_{T}\in\mathbb{R}^{3}, BT3×3B_{T}\in\mathbb{R}^{3\times 3}, and set JT:=det(BT)J_{T}:=\det(B_{T}). We use the generic notation n^{\widehat{n}} for exterior normal vectors on faces F^(T^){\widehat{F}}\in\mathcal{F}({\widehat{T}}). Tensors τ^,σ^:T^SS{\widehat{\tau}},\,{\widehat{\sigma}}:\;{\widehat{T}}\to\SS and vector functions v^:T^3{\widehat{v}}:\;{\widehat{T}}\to\mathbb{R}^{3} are transformed onto elements T𝒯T\in\mathcal{T} as 𝒫𝒦T(τ^)=τ\mathcal{PK}_{{T}}({\widehat{\tau}})=\tau, 𝒫𝒦T(σ^)=σ\mathcal{PK}^{\perp}_{{T}}({\widehat{\sigma}})=\sigma and 𝒫T(v^)=v\mathcal{P}_{T}({\widehat{v}})=v, where

JT2τGT:=BTτ^BT,σGT:=JTBTσ^BT1,andvGT:=JTBTv^onT^.J_{T}^{2}\tau\circ G_{T}:=B_{T}{\widehat{\tau}}B_{T}^{\top},\quad\sigma\circ G_{T}:=J_{T}B_{T}^{-\top}{\widehat{\sigma}}B_{T}^{-1},\quad\text{and}\quad v\circ G_{T}:=J_{T}B_{T}^{-\top}{\widehat{v}}\quad\text{on}\ {\widehat{T}}.

These are re-scaled Piola–Kirchhoff and Piola transformations. They conserve the properties of space 𝕏(T)\mathbb{X}(T) and degrees of freedom (9), as seen next.

Lemma 8.

Let T𝒯T\in\mathcal{T} and τL2(T;SS)\tau\in L_{2}(T;\SS) be given with τ=𝒫𝒦T(τ^)\tau=\mathcal{PK}_{{T}}({\widehat{\tau}}).

(i) Property τ𝕏(T)\tau\in\mathbb{X}(T) holds if and only if τ^𝕏(T^){\widehat{\tau}}\in\mathbb{X}({\widehat{T}}).

(ii) For F(T)F\in\mathcal{F}(T), qP2(F)q\in P^{2}(F), cP0(T;SS)c\in P^{0}(T;\SS), and vP1(T;3)v\in P^{1}(T;\mathbb{R}^{3}) let F^(T^){\widehat{F}}\in\mathcal{F}({\widehat{T}}), q^P2(F^)\widehat{q}\in P^{2}(\widehat{F}), c^P0(T^;SS)\widehat{c}\in P^{0}({\widehat{T}};\SS), and v^P1(T^;3)\widehat{v}\in P^{1}({\widehat{T}};\mathbb{R}^{3}) be the transformed objects with F=GT(F^)F=G_{T}({\widehat{F}}), qGT=q^q\circ G_{T}=\widehat{q}, c=𝒫𝒦T(c^)c=\mathcal{PK}^{\perp}_{{T}}(\widehat{c}), and v=𝒫T(v^)v=\mathcal{P}_{T}({\widehat{v}}). Then,

|F|nτn,qF=|F^|n^τ^n^,q^F^,(τ,c)T=(τ^,c^)T^,(divτ,v)T=(div^τ^,v^)T^.\displaystyle|F|\langle{}n\cdot\tau n\hskip 1.42262pt,q\rangle_{F}=|{\widehat{F}}|\langle{}{\widehat{n}}\cdot{\widehat{\tau}}{\widehat{n}}\hskip 1.42262pt,\widehat{q}\rangle_{{\widehat{F}}},\quad(\tau\hskip 1.42262pt,c)_{T}=({\widehat{\tau}}\hskip 1.42262pt,\widehat{c})_{\widehat{T}},\quad(\operatorname{div}\tau\hskip 1.42262pt,v)_{T}=(\widehat{\operatorname{div}}{\widehat{\tau}}\hskip 1.42262pt,{\widehat{v}})_{\widehat{T}}.
Proof.

A proof of statement (ii) is analogous to that of the two-dimensional case on triangles, see [7, Lemma 13]. Statement (i) is consequence of relations (ii). In fact, the Piola–Kirchhoff transformation 𝒫𝒦T\mathcal{PK}_{{T}} and its inverse conserve the space of symmetric polynomial tensor functions of a fixed degree. The conservation of the moments of the normal-normal traces implies that 𝒫𝒦T\mathcal{PK}_{{T}} is a bijective mapping between Pnn0(T^;SS)P^{0}_{nn}({\widehat{T}};\SS) and Pnn0(T;SS)P^{0}_{nn}(T;\SS), and between P00(T^;SS)P^{0}_{0}({\widehat{T}};\SS) and P00(T;SS)P^{0}_{0}(T;\SS). It also means that the normal-normal traces of τ\tau on faces F(T)F\in\mathcal{F}(T) are linear polynomials if and only if this applies to the normal-normal traces of τ^{\widehat{\tau}} on F^(T^){\widehat{F}}\in\mathcal{F}({\widehat{T}}). This finishes the proof. ∎

Proposition 9 (degrees of freedom).

For a tetrahedron T3T\subset\mathbb{R}^{3} relations P0(T;SS)𝕏(T)P^{0}(T;\SS)\subset\mathbb{X}(T), dim𝕏(T)=30\dim\mathbb{X}(T)=30 hold, and (9) are degrees of freedom of 𝕏(T)\mathbb{X}(T) for a finite element in the sense of Ciarlet.

Proof.

The first relation follows from of Lemma 7 and the fact that Pnn0(T;SS)P^{0}_{nn}(T;\SS) and P00(T;SS)P^{0}_{0}(T;\SS) are subsets of 𝕏(T)\mathbb{X}(T) by construction.

Dimension. By definition of Pnn0(T;SS)P^{0}_{nn}(T;\SS),

P2(T)Pnn0(T;SS)=T1P2(T)T4P2(T)P^{2}(T)P^{0}_{nn}(T;\SS)=T_{1}P^{2}(T)\oplus\cdots\oplus T_{4}P^{2}(T)

has dimension 4040. Since nj(TiP2(T))nj=δijP2(T)n_{j}\cdot(T_{i}P^{2}(T))n_{j}=\delta_{ij}P^{2}(T) for i,j=1,,4i,j=1,\ldots,4, the constraint (TiP2(T))P2,1(T,SS)\bigl{(}T_{i}P^{2}(T)\bigr{)}\cap P^{2,1}(T,\SS) requires to select the subspaces

Pi2,1(T)\displaystyle P^{2,1}_{i}(T) :={vP2(T);v|FiP1(Fi)}=P1(T)span{λiλi+1,λiλi+2,λiλi+3}\displaystyle:=\{v\in P^{2}(T);\;v|_{F_{i}}\in P^{1}(F_{i})\}=P^{1}(T)\oplus\mathrm{span}\{\lambda_{i}\lambda_{i+1},\lambda_{i}\lambda_{i+2},\lambda_{i}\lambda_{i+3}\}

for i=1,,4i=1,\ldots,4, each of dimension 77. Summing the dimensions gives 44 times 77 plus the dimension 22 of P00(T;SS)P^{0}_{0}(T;\SS) in (8).

Degrees of freedom. By Lemma 8, (9) are degrees of freedom of 𝕏(T)\mathbb{X}(T) for a general tetrahedron T𝒯T\in\mathcal{T} if and only if they are degrees of freedom for the reference tetrahedron T=T^T={\widehat{T}}. In the remainder of this proof we therefore assume without loss of generality that T=T^T={\widehat{T}}. We need this assumption only in the final step to verify the invertibility of a 12×1212\times 12-matrix. The number of functionals in (9) is 3030: 12 in (9a), 6 in (9b), and 12 in (9c). It is therefore enough to show the linear independence of (9) on 𝕏(T)\mathbb{X}(T).

Linear independence. Let τ𝕏(T)\tau\in\mathbb{X}(T) be given with vanishing functionals (9). We have to show that τ=0\tau=0. The vanishing of (9a) leads to

τi=14(TiP2(T)P02(T,SS))P00(T;SS).\tau\in\oplus_{i=1}^{4}\bigr{(}T_{i}P^{2}(T)\cap P^{2}_{0}(T,\SS)\bigr{)}\oplus P^{0}_{0}(T;\SS).

Since

TiP2(T)P02(T,SS)=Tispan{ϕP2(T);ϕ|Fi=0}=Tispan{λi,λiλi+1,λiλi+2,λiλi+3}\displaystyle T_{i}P^{2}(T)\cap P^{2}_{0}(T,\SS)=T_{i}\,\mathrm{span}\{\phi\in P^{2}(T);\;\phi|_{F_{i}}=0\}=T_{i}\,\mathrm{span}\{\lambda_{i},\lambda_{i}\lambda_{i+1},\lambda_{i}\lambda_{i+2},\lambda_{i}\lambda_{i+3}\}

for i=1,,4i=1,\ldots,4 we deduce the representation

τ=i=14λiϕiTi+ϕ5T5+ϕ6T6withϕiP1(T) for i=1,,4, and ϕ5,ϕ6.\displaystyle\tau=\sum_{i=1}^{4}\lambda_{i}\phi_{i}T_{i}+\phi_{5}T_{5}+\phi_{6}T_{6}\quad\text{with}\quad\phi_{i}\in P^{1}(T)\text{ for }i=1,\ldots,4,\ \text{ and }\phi_{5},\phi_{6}\in\mathbb{R}.

The vanishing of degrees of freedom (9b) means

i=14(λiϕi,1)TTi+(ϕ5,1)TT5+(ϕ6,1)TT6=0,\sum_{i=1}^{4}(\lambda_{i}\phi_{i}\hskip 1.42262pt,1)_{T}T_{i}+(\phi_{5}\hskip 1.42262pt,1)_{T}T_{5}+(\phi_{6}\hskip 1.42262pt,1)_{T}T_{6}=0,

so that, by the linear independence of T1,,T6T_{1},\ldots,T_{6},

(λiϕi,1)T=0for i=1,,4,andϕ5=ϕ6=0.\displaystyle(\lambda_{i}\phi_{i}\hskip 1.42262pt,1)_{T}=0\quad\text{for }i=1,\ldots,4,\quad\text{and}\quad\phi_{5}=\phi_{6}=0.

We conclude that

τ=i=14λiϕiTiand(λjϕj,1)T=0forj=1,,4.\displaystyle\tau=\sum_{i=1}^{4}\lambda_{i}\phi_{i}T_{i}\quad\text{and}\quad(\lambda_{j}\phi_{j}\hskip 1.42262pt,1)_{T}=0\quad\text{for}\ j=1,\ldots,4. (10)

In the following we represent ϕi=j=14ϕi(xj)λj\phi_{i}=\sum_{j=1}^{4}\phi_{i}(x_{j})\lambda_{j} and calculate ϕi=j=14ϕi(xj)n~j\nabla\phi_{i}=\sum_{j=1}^{4}\phi_{i}(x_{j})\widetilde{n}_{j} for i=1,,4i=1,\ldots,4 (recall that n~j=λjspan(nj)\widetilde{n}_{j}=\nabla\lambda_{j}\in\mathrm{span}(n_{j})). Representation (10) and the vanishing of (9c) reveal that

0=divτ(xk)\displaystyle 0=\operatorname{div}\tau(x_{k}) =i=14Ti(ϕiλi+λiϕi)(xk)=i=14Ti(ϕin~i+λij=14ϕi(xj)n~j)(xk)\displaystyle=\sum_{i=1}^{4}T_{i}\bigl{(}\phi_{i}\nabla\lambda_{i}+\lambda_{i}\nabla\phi_{i}\bigr{)}(x_{k})=\sum_{i=1}^{4}T_{i}\bigl{(}\phi_{i}\widetilde{n}_{i}+\lambda_{i}\sum_{j=1}^{4}\phi_{i}(x_{j})\widetilde{n}_{j}\bigr{)}(x_{k})
=i=14Ti(ϕi(xk)n~i+δikj=14ϕi(xj)n~j)fork=1,,4.\displaystyle=\sum_{i=1}^{4}T_{i}\bigl{(}\phi_{i}(x_{k})\widetilde{n}_{i}+\delta_{ik}\sum_{j=1}^{4}\phi_{i}(x_{j})\widetilde{n}_{j}\bigr{)}\quad\text{for}\ k=1,\ldots,4. (11)

These equations form a homogeneous linear system of 1212 equations for the 1616 unknowns ϕi(xj)\phi_{i}(x_{j}), i,j=1,,4i,j=1,\ldots,4. Instead of unknowns ϕi(xi)\phi_{i}(x_{i}) we use the scaled unknowns 2ϕi(xi)2\phi_{i}(x_{i}) for i=1,,4i=1,\ldots,4 and order the equations with respect to kk and the unknowns with inner loop over jj and outer loop over ii, namely

2ϕ1(x1),ϕ1(x2),ϕ1(x3),ϕ1(x4),ϕ2(x1),2ϕ2(x2),ϕ2(x3),ϕ2(x4),,ϕ4(x3),2ϕ4(x4).2\phi_{1}(x_{1}),\phi_{1}(x_{2}),\phi_{1}(x_{3}),\phi_{1}(x_{4}),\quad\phi_{2}(x_{1}),2\phi_{2}(x_{2}),\phi_{2}(x_{3}),\phi_{2}(x_{4}),\quad\ldots,\phi_{4}(x_{3}),2\phi_{4}(x_{4}).

Denoting tij:=Tin~jt_{ij}:=T_{i}\widetilde{n}_{j}, the corresponding matrix of system (4.1) becomes

(t11t12t13t14t22000t33000t440000t1100t21t22t23t240t33000t440000t11000t220t31t32t33t3400t440000t11000t22000t33t41t42t43t44).\begin{pmatrix}\begin{array}[]{cccc|cccc|cccc|cccc}t_{11}&t_{12}&t_{13}&t_{14}&t_{22}&0&0&0&t_{33}&0&0&0&t_{44}&0&0&0\\ \hline\cr 0&t_{11}&0&0&t_{21}&t_{22}&t_{23}&t_{24}&0&t_{33}&0&0&0&t_{44}&0&0\\ \hline\cr 0&0&t_{11}&0&0&0&t_{22}&0&t_{31}&t_{32}&t_{33}&t_{34}&0&0&t_{44}&0\\ \hline\cr 0&0&0&t_{11}&0&0&0&t_{22}&0&0&0&t_{33}&t_{41}&t_{42}&t_{43}&t_{44}\end{array}\end{pmatrix}.

It remains to incorporate the relations (λiϕi,1)T=0(\lambda_{i}\phi_{i}\hskip 1.42262pt,1)_{T}=0 for i=1,,4i=1,\ldots,4 from (10). We note that (λi,λj)T/(λi,λi+1)T=δij+1(\lambda_{i}\hskip 1.42262pt,\lambda_{j})_{T}/(\lambda_{i}\hskip 1.42262pt,\lambda_{i+1})_{T}=\delta_{ij}+1, see, e.g., [25, Theorem 3.1], and recall the representation ϕi=j=14ϕi(xj)λj\phi_{i}=\sum_{j=1}^{4}\phi_{i}(x_{j})\lambda_{j}. We find that (λiϕi,1)T=j=14ϕi(xj)(λi,λj)T=0(\lambda_{i}\phi_{i}\hskip 1.42262pt,1)_{T}=\sum_{j=1}^{4}\phi_{i}(x_{j})(\lambda_{i}\hskip 1.42262pt,\lambda_{j})_{T}=0 holds if and only if

1(λi,λi+1)Tj=14ϕi(xj)(λi,λj)T=2ϕi(xi)+ϕi(xi+1)+ϕi(xi+2)+ϕi(xi+3)=0,i=1,,4.\frac{1}{(\lambda_{i}\hskip 1.42262pt,\lambda_{i+1})_{T}}\sum_{j=1}^{4}\phi_{i}(x_{j})(\lambda_{i}\hskip 1.42262pt,\lambda_{j})_{T}=2\phi_{i}(x_{i})+\phi_{i}(x_{i+1})+\phi_{i}(x_{i+2})+\phi_{i}(x_{i+3})=0,\quad i=1,\ldots,4.

We use this relation to eliminate the unknowns 2ϕi(xi)2\phi_{i}(x_{i}) for i=1,,4i=1,\ldots,4. With the notation t~ij:=tijtii\widetilde{t}_{ij}:=t_{ij}-t_{ii} the reduced 12×1212\times 12 matrix reads

(t~12t~13t~14t2200t3300t4400t1100t~21t~23t~240t3300t4400t1100t220t~31t~32t~3400t4400t1100t2200t33t~41t~42t~43).\begin{pmatrix}\begin{array}[]{ccc|ccc|ccc|ccc}\widetilde{t}_{12}&\widetilde{t}_{13}&\widetilde{t}_{14}&t_{22}&0&0&t_{33}&0&0&t_{44}&0&0\\ \hline\cr\rule{0.0pt}{11.19443pt}t_{11}&0&0&\widetilde{t}_{21}&\widetilde{t}_{23}&\widetilde{t}_{24}&0&t_{33}&0&0&t_{44}&0\\ \hline\cr\rule{0.0pt}{11.19443pt}0&t_{11}&0&0&t_{22}&0&\widetilde{t}_{31}&\widetilde{t}_{32}&\widetilde{t}_{34}&0&0&t_{44}\\ \hline\cr\rule{0.0pt}{11.19443pt}0&0&t_{11}&0&0&t_{22}&0&0&t_{33}&\widetilde{t}_{41}&\widetilde{t}_{42}&\widetilde{t}_{43}\end{array}\end{pmatrix}. (12)

The proof of the lemma is finished by noting that this matrix is invertible. Let us give the details.

Recall that we are considering the reference tetrahedron T=T^T={\widehat{T}} with the previously introduced numbering of vertices. In particular,

λ1(x,y,z)=1xyz,λ2(x,y,z)=x,λ3(x,y,z)=y,λ4=z,\displaystyle\lambda_{1}(x,y,z)=1-x-y-z,\ \lambda_{2}(x,y,z)=x,\ \lambda_{3}(x,y,z)=y,\ \lambda_{4}=z,
n~1=(1,1,1),n~2=(1,0,0),n~3=(0,1,0),n~4=(0,0,1),\displaystyle\widetilde{n}_{1}=-(1,1,1)^{\top},\ \widetilde{n}_{2}=(1,0,0)^{\top},\ \widetilde{n}_{3}=(0,1,0)^{\top},\ \widetilde{n}_{4}=(0,0,1)^{\top},

and selecting the edge vectors

e1,2=(0,1,1),e1,3=(1,0,1),e1,4=(1,1,0),e2,3=(0,0,1),\displaystyle e_{1,2}=(0,1,-1)^{\top},\ e_{1,3}=(1,0,-1)^{\top},\ e_{1,4}=(1,-1,0)^{\top},\ e_{2,3}=(0,0,1)^{\top},
e2,4=(0,1,0),e3,4=(1,0,0),ei,j=ej,i(i>j)\displaystyle e_{2,4}=(0,1,0)^{\top},\ e_{3,4}=(1,0,0)^{\top},\ e_{i,j}=e_{j,i}\ (i>j)

we calculate

T1=e2,3e2,4n1e2,3n1e2,4=(000003/203/20),\displaystyle T_{1}=\frac{e_{2,3}\odot e_{2,4}}{n_{1}\cdot e_{2,3}\;n_{1}\cdot e_{2,4}}=\begin{pmatrix}0&0&0\\ 0&0&3/2\\ 0&3/2&0\end{pmatrix}, T2=e3,4e3,1n2e3,4n2e3,4=(101/20001/200),\displaystyle T_{2}=\frac{e_{3,4}\odot e_{3,1}}{n_{2}\cdot e_{3,4}\;n_{2}\cdot e_{3,4}}=\begin{pmatrix}1&0&-1/2\\ 0&0&0\\ -1/2&0&0\end{pmatrix},
T3=e4,1e4,2n3e4,1n3e4,2=(01/201/210000),\displaystyle T_{3}=\frac{e_{4,1}\odot e_{4,2}}{n_{3}\cdot e_{4,1}\;n_{3}\cdot e_{4,2}}=\begin{pmatrix}0&-1/2&0\\ -1/2&1&0\\ 0&0&0\end{pmatrix}, T4=e1,2e1,3n4e1,2n4e1,3=(01/21/21/201/21/21/21).\displaystyle T_{4}=\frac{e_{1,2}\odot e_{1,3}}{n_{4}\cdot e_{1,2}\;n_{4}\cdot e_{1,3}}=\begin{pmatrix}0&1/2&-1/2\\ 1/2&0&-1/2\\ -1/2&-1/2&1\end{pmatrix}.

This gives (note that ti,i+1=(0,0,0)t_{i,i+1}=(0,0,0)^{\top})

t11=(03/23/2),t13=(003/2),t14=(03/20),\displaystyle t_{11}=\begin{pmatrix}0\\ -3/2\\ -3/2\end{pmatrix},\ t_{13}=\begin{pmatrix}0\\ 0\\ 3/2\end{pmatrix},\ t_{14}=\begin{pmatrix}0\\ 3/2\\ 0\end{pmatrix},\ t~12=t11,t~13=(03/23),t~14=(033/2),\displaystyle\widetilde{t}_{12}=-t_{11},\ \widetilde{t}_{13}=\begin{pmatrix}0\\ 3/2\\ 3\end{pmatrix},\ \widetilde{t}_{14}=\begin{pmatrix}0\\ 3\\ 3/2\end{pmatrix},
t22=(101/2),t21=(1/201/2),t24=(1/200),\displaystyle t_{22}=\begin{pmatrix}1\\ 0\\ -1/2\end{pmatrix},\ t_{21}=\begin{pmatrix}-1/2\\ 0\\ 1/2\end{pmatrix},\ t_{24}=\begin{pmatrix}-1/2\\ 0\\ 0\end{pmatrix},\ t~21=(3/201),t~23=t22,t~24=(3/201/2),\displaystyle\widetilde{t}_{21}=\begin{pmatrix}-3/2\\ 0\\ 1\end{pmatrix},\ \widetilde{t}_{23}=-t_{22},\ \widetilde{t}_{24}=\begin{pmatrix}-3/2\\ 0\\ 1/2\end{pmatrix},
t33=(1/210),t31=(1/21/20),t32=(01/20),\displaystyle t_{33}=\begin{pmatrix}-1/2\\ 1\\ 0\end{pmatrix},\ t_{31}=\begin{pmatrix}1/2\\ -1/2\\ 0\end{pmatrix},\ t_{32}=\begin{pmatrix}0\\ -1/2\\ 0\end{pmatrix},\ t~31=(13/20),t~32=(1/23/20),t~34=t33,\displaystyle\widetilde{t}_{31}=\begin{pmatrix}1\\ -3/2\\ 0\end{pmatrix},\ \widetilde{t}_{32}=\begin{pmatrix}1/2\\ -3/2\\ 0\end{pmatrix},\ \widetilde{t}_{34}=-t_{33},
t44=(1/21/21),t42=(01/21/2),t43=(1/201/2),\displaystyle t_{44}=\begin{pmatrix}-1/2\\ -1/2\\ 1\end{pmatrix},\ t_{42}=\begin{pmatrix}0\\ 1/2\\ -1/2\end{pmatrix},\ t_{43}=\begin{pmatrix}1/2\\ 0\\ -1/2\end{pmatrix},\ t~41=t44,t~42=(1/213/2),t~43=(11/23/2),\displaystyle\widetilde{t}_{41}=-t_{44},\ \widetilde{t}_{42}=\begin{pmatrix}1/2\\ 1\\ -3/2\end{pmatrix},\ \widetilde{t}_{43}=\begin{pmatrix}1\\ 1/2\\ -3/2\end{pmatrix},

and matrix (12) reads

(0001001/2001/2003/23/230001001/2003/233/21/2000001000003/213/201/2001/203/20000001001/203/20011/21/200001000001011/21/2001/203/200003/23/21001/203/2001/20000001000001001/21/21/21003/20000011/211/2003/2001/200013/23/2).\begin{pmatrix}\begin{array}[]{ccc|ccc|ccc|ccc}0&0&0&1&0&0&-1/2&0&0&-1/2&0&0\\ 3/2&3/2&3&0&0&0&1&0&0&-1/2&0&0\\ 3/2&3&3/2&-1/2&0&0&0&0&0&1&0&0\\ \hline\cr 0&0&0&-3/2&-1&-3/2&0&-1/2&0&0&-1/2&0\\ -3/2&0&0&0&0&0&0&1&0&0&-1/2&0\\ -3/2&0&0&1&1/2&1/2&0&0&0&0&1&0\\ \hline\cr 0&0&0&0&1&0&1&1/2&1/2&0&0&-1/2\\ 0&-3/2&0&0&0&0&-3/2&-3/2&-1&0&0&-1/2\\ 0&-3/2&0&0&-1/2&0&0&0&0&0&0&1\\ \hline\cr 0&0&0&0&0&1&0&0&-1/2&1/2&1/2&1\\ 0&0&-3/2&0&0&0&0&0&1&1/2&1&1/2\\ 0&0&-3/2&0&0&-1/2&0&0&0&-1&-3/2&-3/2\end{array}\end{pmatrix}.

A lengthy calculation shows that its determinant is 2834522^{-8}3^{4}5^{2}. This finishes the proof. ∎

4.2 Approximation space

Elements 𝕏(T)\mathbb{X}(T) (T𝒯T\in\mathcal{T}) from (8) generate the product space 𝕏(𝒯)\mathbb{X}(\mathcal{T}) that gives rise to the Hnn(div,𝒯;SS)H_{nn}(\operatorname{div},\mathcal{T};\SS)-conforming finite element space

Pnn2(𝒯):=𝕏(𝒯)Hnn(div,𝒯;SS)={τ𝕏(𝒯);[nτn]|F=0F(Ω)}.P^{2}_{nn}(\mathcal{T}):=\mathbb{X}(\mathcal{T})\cap H_{nn}(\operatorname{div},\mathcal{T};\SS)=\{\tau\in\mathbb{X}(\mathcal{T});\;[n\cdot\tau n]|_{F}=0\ \forall F\in\mathcal{F}(\Omega)\}.

Here, [nτn]|F[n\cdot\tau n]|_{F} denotes the normal-normal jump of τ\tau across FF. Equivalently, this conformity is achieved by the selection of unique degrees of freedom (9a) assigned to interior faces F(Ω)F\in\mathcal{F}(\Omega). By also assigning volume-related degrees of freedom (9b), (9c) we obtain an interpolation operator

𝒯nn:H(div,Ω;SS)Lq(Ω;SS)Pnn2(𝒯)(q>2).\mathcal{I}^{nn}_{\mathcal{T}}:\;H(\operatorname{div},\Omega;\SS)\cap L_{q}(\Omega;\SS)\to P^{2}_{nn}(\mathcal{T})\quad(q>2).

In the following, div𝒯\operatorname{div}_{\mathcal{T}} is the piecewise divergence operator and Π𝒯1\Pi^{1}_{\mathcal{T}} denotes the L2L_{2}-projection onto piecewise linear vector-valued polynomials.

Proposition 10 (interpolation).

Operator 𝒯nn:H(div,Ω;SS)Lq(Ω;SS)Pnn2(𝒯)\mathcal{I}^{nn}_{\mathcal{T}}:\;H(\operatorname{div},\Omega;\SS)\cap L_{q}(\Omega;\SS)\to P^{2}_{nn}(\mathcal{T}) is well defined and bounded for q>2q>2. It commutes with the piecewise divergence operator, div𝒯𝒯nn=Π𝒯1div,\operatorname{div}_{\mathcal{T}}\mathcal{I}^{nn}_{\mathcal{T}}=\Pi^{1}_{\mathcal{T}}\operatorname{div}, and has the projection property 𝒯nnτ=τ\mathcal{I}^{nn}_{\mathcal{T}}\tau=\tau for τPnn2(𝒯)\tau\in P^{2}_{nn}(\mathcal{T}). Given 0<s10<s\leq 1, 0r20\leq r\leq 2, any τHs,r(div,Ω;SS)\tau\in H^{s,r}(\operatorname{div},\Omega;\SS) satisfies

τ𝒯nnτ\displaystyle\|\tau-\mathcal{I}^{nn}_{\mathcal{T}}\tau\| h𝒯sτs+h𝒯divτ\displaystyle\lesssim\|h_{\mathcal{T}}\|_{\infty}^{s}\|\tau\|_{s}+\|h_{\mathcal{T}}\|_{\infty}\|\operatorname{div}\tau\|\quad τHs,0(div,Ω;SS),\displaystyle\forall\tau\in H^{s,0}(\operatorname{div},\Omega;\SS),
div(τ𝒯nnτ)𝒯\displaystyle\|\operatorname{div}(\tau-\mathcal{I}^{nn}_{\mathcal{T}}\tau)\|_{\mathcal{T}} h𝒯rdivτr\displaystyle\lesssim\|h_{\mathcal{T}}\|_{\infty}^{r}\|\operatorname{div}\tau\|_{r}\quad τHs,r(div,Ω;SS)\displaystyle\forall\tau\in H^{s,r}(\operatorname{div},\Omega;\SS)

with intrinsic constants independent of τ\tau and 𝒯\mathcal{T}.

Proof.

The proof is almost verbatim to the proof of the two-dimensional case [7, Proposition 17]. We only have to verify that the face degrees of freedom are bounded functionals for τLq(Ω;SS)\tau\in L_{q}(\Omega;\SS) if q>2q>2, and to recall that Hs(Ω;SS)Lq(Ω;SS)H^{s}(\Omega;\SS)\subset L_{q}(\Omega;\SS) for 0<s10<s\leq 1 and q=6/(32s)>2q=6/(3-2s)>2 by the Sobolev embedding theorem [1, Theorem 7.57]. The former fact can be proved analogously to [3, Lemma 4.7]. Essential step in the proof of that lemma is the continuity of the extension by zero from the Sobolev space W11/q,q(E)W^{1-1/q^{\prime},q^{\prime}}(E) on an edge EE of a face FF to W11/q,q(F)W^{1-1/q^{\prime},q^{\prime}}(\partial F) where 1/q+1/q=11/q^{\prime}+1/q=1. This zero extension is also continuous as a mapping from W11/q,q(F)W^{1-1/q^{\prime},q^{\prime}}(F) to W11/q,q(T)W^{1-1/q^{\prime},q^{\prime}}(\partial T) where T𝒯T\in\mathcal{T} and F(T)F\in\mathcal{F}(T), cf. [11, Corollary 1.4.4.5], and thus gives the result. ∎

The next result can be shown analogously to [7, Lemma 20].

Lemma 11.

Any τPnn2(𝒯)\tau\in P^{2}_{nn}(\mathcal{T}) with

div𝒯τ=0,γ1,𝑛𝑛(w),τ𝒮=0wP1(𝒯;3)H01(Ω;3),and(tr(τ),1)=0\operatorname{div}_{\mathcal{T}}\tau=0,\quad\langle{}{\gamma_{1,\mathit{nn}}}(w)\hskip 1.42262pt,\tau\rangle_{\mathcal{S}}=0\ \forall w\in P^{1}(\mathcal{T};\mathbb{R}^{3})\cap H^{1}_{0}(\Omega;\mathbb{R}^{3}),\quad\text{and}\quad(\operatorname{tr}(\tau)\hskip 1.42262pt,1)=0

satisfies

tr(τ)Cdevτ\|\operatorname{tr}(\tau)\|\leq C\|\operatorname{dev}\tau\|

with a constant C>0C>0 that is independent of τ\tau and 𝒯\mathcal{T}.

4.3 Finite element scheme and locking-free convergence

Let S01(𝒯;3):=P1(𝒯;3)H01(Ω;3)S^{1}_{0}(\mathcal{T};\mathbb{R}^{3}):=P^{1}(\mathcal{T};\mathbb{R}^{3})\cap H^{1}_{0}(\Omega;\mathbb{R}^{3}) be the space of continuous, piecewise linear vector-valued functions that vanish on Γ\Gamma. We select the discrete trace space S01(𝒮):=γ1,𝑛𝑛(S01(𝒯;3)).S^{1}_{0}(\mathcal{S}):={\gamma_{1,\mathit{nn}}}\bigl{(}S^{1}_{0}(\mathcal{T};\mathbb{R}^{3})\bigr{)}. Given fL2(Ω;3)f\in L_{2}(\Omega;\mathbb{R}^{3}), our mixed finite element scheme seeks σhPnn2(𝒯)\sigma_{h}\in P^{2}_{nn}(\mathcal{T}), uhP1(𝒯;3)u_{h}\in P^{1}(\mathcal{T};\mathbb{R}^{3}), and ηhS01(𝒮)\eta_{h}\in S^{1}_{0}(\mathcal{S}) such that

(𝔸σh,δσ)+(uh,divδσ)𝒯ηh,δσ𝒮\displaystyle(\mathbb{A}\sigma_{h}\hskip 1.42262pt,{\delta\!\sigma})+(u_{h}\hskip 1.42262pt,\operatorname{div}{\delta\!\sigma})_{\mathcal{T}}-\langle{}\eta_{h}\hskip 1.42262pt,{\delta\!\sigma}\rangle_{\mathcal{S}} =0,\displaystyle=0, (13a)
(divσh,δu)𝒯δη,σh𝒮\displaystyle(\operatorname{div}\sigma_{h}\hskip 1.42262pt,{\delta\!u})_{\mathcal{T}}-\langle{}{\delta\!\eta}\hskip 1.42262pt,\sigma_{h}\rangle_{\mathcal{S}} =(f,δu)\displaystyle=-(f\hskip 1.42262pt,{\delta\!u}) (13b)

for any δσPnn2(𝒯){\delta\!\sigma}\in P^{2}_{nn}(\mathcal{T}), δuP1(𝒯;3){\delta\!u}\in P^{1}(\mathcal{T};\mathbb{R}^{3}), and δηS01(𝒮){\delta\!\eta}\in S^{1}_{0}(\mathcal{S}).

This scheme is locking free and converges quasi-optimally.

Theorem 12 (locking-free convergence).

For given fL2(Ω;3)f\in L_{2}(\Omega;\mathbb{R}^{3}) let (σ,u,η)(\sigma,u,\eta) denote the solution to (6). Mixed scheme (13) has a unique solution (σh,uh,ηh)(\sigma_{h},u_{h},\eta_{h}). It satisfies

σσhdiv,𝒯+uuh+ηηh1/2,t,𝒮\displaystyle\|\sigma-\sigma_{h}\|_{\operatorname{div},\mathcal{T}}+\|u-u_{h}\|+\|\eta-\eta_{h}\|_{1/2,t,\mathcal{S}}
inf{στdiv,𝒯+uv+ηρ1/2,t,𝒮;τPnn2(𝒯),vP1(𝒯;3),ρS01(𝒮)}\displaystyle\lesssim\inf\bigl{\{}\|\sigma-\tau\|_{\operatorname{div},\mathcal{T}}+\|u-v\|+\|\eta-\rho\|_{1/2,t,\mathcal{S}};\;\tau\in P^{2}_{nn}(\mathcal{T}),v\in P^{1}(\mathcal{T};\mathbb{R}^{3}),\rho\in S^{1}_{0}(\mathcal{S})\bigr{\}}

with a hidden constant that is independent of 𝒯\mathcal{T} and λ(0,)\lambda\in(0,\infty).

Let us furthermore assume that uH1+s(Ω;3)u\in H^{1+s}(\Omega;\mathbb{R}^{3}) and fHr(Ω;3)f\in H^{r}(\Omega;\mathbb{R}^{3}) for fixed 0<s10<s\leq 1, sr2s\leq r\leq 2, and that there is a regularity shift 0<s10<{s^{\prime}}\leq 1 with

(div𝒞ε)1g1+scggL2(Ω;3)\|(\operatorname{div}\mathcal{C}\varepsilon)^{-1}g\|_{1+{s^{\prime}}}\leq c\|g\|\quad\forall g\in L_{2}(\Omega;\mathbb{R}^{3})

(inversion subject to the homogeneous Dirichlet condition) for a constant c>0c>0 that is independent of gg and λ(0,)\lambda\in(0,\infty). Then the estimates

σσhdiv,𝒯+uuh+ηηh1/2,t,𝒮\displaystyle\|\sigma-\sigma_{h}\|_{\operatorname{div},\mathcal{T}}+\|u-u_{h}\|+\|\eta-\eta_{h}\|_{1/2,t,\mathcal{S}} h𝒯s(u1+s+σs+fs),\displaystyle\lesssim\|h_{\mathcal{T}}\|_{\infty}^{s}(\|u\|_{1+s}+\|\sigma\|_{s}+\|f\|_{s}), (14)
div(σσh)𝒯\displaystyle\|\operatorname{div}(\sigma-\sigma_{h})\|_{\mathcal{T}} h𝒯rfr,\displaystyle\lesssim\|h_{\mathcal{T}}\|_{\infty}^{r}\|f\|_{r},
uuh\displaystyle\|u-u_{h}\| h𝒯s+s(u1+s+fs)\displaystyle\lesssim\|h_{\mathcal{T}}\|_{\infty}^{s+{s^{\prime}}}\bigl{(}\|u\|_{1+s}+\|f\|_{s}\bigr{)}

hold with intrinsic constants that are independent of 𝒯\mathcal{T} and λ(0,)\lambda\in(0,\infty).

Proof.

The proof is analogous to the proofs of Theorems 21 (quasi-optimal convergence) and 22 (approximation orders) in [7]. To this end we note that, for homogeneous Dirichlet data, the condition uDH1+s(Ω;3)u_{D}\in H^{1+s}(\Omega;\mathbb{R}^{3}) with s>0s>0 for a Dirichlet extension uDu_{D} in [7, Theorem 21] is obsolete and that the tensors σ0=σ0,hSS\sigma_{0}=\sigma_{0,h}\in\SS appearing in [7, Theorem 22] vanish, cf. definitions (9) and (28) there. Specifically, Lemma 11 is critical for the robust stability of the discrete scheme and, having established the quasi-optimal convergence, error estimate (14) is consequence of Proposition 10. ∎

5 Numerical experiments

We use Ω=(0,1)3\Omega=(0,1)^{3} and consider Dirichlet boundary conditions everywhere given by the manufactured displacement

u(x1,x2,x3)=(sin(3x1)cos(3x2)cos(3x3)cos(3x1)sin(3x2)cos(3x3)cos(3x1)cos(3x3)sin(3x3)),u(x_{1},x_{2},x_{3})=\begin{pmatrix}\sin(3x_{1})\cos(3x_{2})\cos(3x_{3})\\ \cos(3x_{1})\sin(3x_{2})\cos(3x_{3})\\ \cos(3x_{1})\cos(3x_{3})\sin(3x_{3})\end{pmatrix},

and select f:=divσf:=-\operatorname{div}\sigma where σ=ε(u)\sigma=\mathbb{C}\varepsilon(u). We choose Young’s modulus E=1E=1 and different values of ν\nu. The Lamé parameters are μ=E2(1+ν)=12(1+ν)\mu=\frac{E}{2(1+\nu)}=\frac{1}{2(1+\nu)} and λ=Eν(1+ν)(12ν)=ν(1+ν)(12ν)\lambda=\frac{E\nu}{(1+\nu)(1-2\nu)}=\frac{\nu}{(1+\nu)(1-2\nu)}.

We use quasi-uniform meshes and plot the individual approximation errors uuh\|u-u_{h}\| (“u”), ε(u)ε(u~h)\|\varepsilon(u)-\varepsilon(\widetilde{u}_{h})\| (“strain”), σσh\|\sigma-\sigma_{h}\| (“sigma”), and div(σσh)𝒯\|\operatorname{div}(\sigma-\sigma_{h})\|_{\mathcal{T}} (“div sigma”), normalized by the norm of the exact solution (u12+σdiv2)1/2\bigl{(}\|u\|_{1}^{2}+\|\sigma\|_{\operatorname{div}}^{2}\bigr{)}^{1/2}. Here, u~hS01(𝒯;3)\widetilde{u}_{h}\in S^{1}_{0}(\mathcal{T};\mathbb{R}^{3}) is the piecewise linear interpolation of the continuous, piecewise linear trace approximation ηh\eta_{h}. Note that ε(u)ε(u~h)\|\varepsilon(u)-\varepsilon(\widetilde{u}_{h})\| is, up to the L2L_{2} part uu~h\|u-\widetilde{u}_{h}\|, an upper bound of γ1,𝑛𝑛(u)ηh1/2,t,𝒮\|{\gamma_{1,\mathit{nn}}}(u)-\eta_{h}\|_{1/2,t,\mathcal{S}}. The right-hand side functional (f,)(f\hskip 1.42262pt,\cdot) is approximated by a 4-point Gauss formula (integrating quadratic polynomials exactly), and the errors are calculated by a 14-point Gauss formula (an overkill to be close to the exact errors), cf. [19]. The non-homogeneous Dirichlet boundary condition is approximated by assigning the corresponding vertex values to trace approximation ηh\eta_{h}.

Figure 1(a) shows the results for ν=0.3\nu=0.3, and confirms the approximation orders with respect to h:=h𝒯h:=\|h_{\mathcal{T}}\|_{\infty} proved by Theorem 12: O(h)O(h) for the stress and trace approximations, and O(h2)O(h^{2}) for the approximations of the displacement and the divergence of the stress.

The sequence of results for ν=0.3,0.45,0.49,0.4999\nu=0.3,0.45,0.49,0.4999 shown in Figure 1 confirms that our scheme is locking free. (The fact that some relative errors are initially smaller for larger values of λ\lambda does not contradict this.) Note that tr(σ)=tr(ε(u))=9(2μ+3λ)cos(3x1)cos(3x2)cos(3x3)\operatorname{tr}(\sigma)=\operatorname{tr}(\mathbb{C}\varepsilon(u))=9(2\mu+3\lambda)\cos(3x_{1})\cos(3x_{2})\cos(3x_{3}) and

devσ=6μ(0sin(3x1)sin(3x2)cos(3x3)sin(3x1)cos(3x2)sin(3x3)0cos(3x1)sin(3x2)sin(3x3)0),\operatorname{dev}\sigma=-6\mu\begin{pmatrix}0&\sin(3x_{1})\sin(3x_{2})\cos(3x_{3})&\sin(3x_{1})\cos(3x_{2})\sin(3x_{3})\\ *&0&\cos(3x_{1})\sin(3x_{2})\sin(3x_{3})\\ *&*&0\end{pmatrix},

that is, tr(σ)\|\operatorname{tr}(\sigma)\| and (tr(σ),1)(\operatorname{tr}(\sigma)\hskip 1.42262pt,1) are unbounded as ν1/2\nu\to 1/2 while devσ\operatorname{dev}\sigma is bounded.

Refer to caption
(a) ν=0.3\nu=0.3
Refer to caption
(b) ν=0.45\nu=0.45
Refer to caption
(c) ν=0.49\nu=0.49
Refer to caption
(d) ν=0.4999\nu=0.4999
Figure 1: Relative errors for the manufactured example with different values of ν\nu.

References

  • [1] Robert A. Adams. Sobolev spaces. Academic Press, New York-London, 1975. Pure and Applied Mathematics, Vol. 65.
  • [2] Scot Adams and Bernardo Cockburn. A mixed finite element method for elasticity in three dimensions. J. Sci. Comput., 25(3):515–521, 2005.
  • [3] Chérif Amrouche, Christine Bernardi, Monique Dauge, and Vivette Girault. Vector potentials in three-dimensional non-smooth domains. Math. Methods Appl. Sci., 21(9):823–864, 1998.
  • [4] Douglas N. Arnold, Gerard Awanou, and Ragnar Winther. Finite elements for symmetric tensors in three dimensions. Math. Comp., 77(263):1229–1251, 2008.
  • [5] Daniele Boffi, Franco Brezzi, and Michel Fortin. Mixed finite element methods and applications, volume 44 of Springer Series in Computational Mathematics. Springer, Heidelberg, 2013.
  • [6] Carsten Carstensen, Leszek F. Demkowicz, and Jay Gopalakrishnan. Breaking spaces and forms for the DPG method and applications including Maxwell equations. Comput. Math. Appl., 72(3):494–522, 2016.
  • [7] Carsten Carstensen and Norbert Heuer. Normal-normal continuous symmetric stresses in mixed finite element elasticity. Math. Comp. (appeared online), DOI 10.1090/mcom/4017.
  • [8] Carsten Carstensen and Norbert Heuer. A fractional-order trace-dev-div inequality. arXiv:2403.01291, 2024.
  • [9] Long Chen and Xuehai Huang. Finite elements for div- and divdiv-conforming symmetric tensors in arbitrary dimension. SIAM J. Numer. Anal., 60(4):1932–1961, 2022.
  • [10] Lawrence C. Evans and Ronald F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
  • [11] Pierre Grisvard. Elliptic Problems in Nonsmooth Domains. Pitman Publishing Inc., Boston, 1985.
  • [12] Kåre Hellan. Analysis of elastic plates in flexure by a simplified finite element method. Acta Polytech. Scand. Civ. Eng. Build. Constr. Ser. 46, 1, 1967.
  • [13] Leonard R. Herrmann. Finite-element bending analysis for plates. J. Eng. Mech. Div., 93(5):13–26, 1967.
  • [14] Norbert Heuer. Generalized mixed and primal hybrid methods with applications to plate bending. arXiv:2406.12541, 2024.
  • [15] Jun Hu. Finite element approximations of symmetric tensors on simplicial grids in n\mathbb{R}^{n}: the higher order case. J. Comput. Math., 33(3):283–296, 2015.
  • [16] Jun Hu and ShangYou Zhang. A family of symmetric mixed finite elements for linear elasticity on tetrahedral grids. Sci. China Math., 58(2):297–307, 2015.
  • [17] Jun Hu and Shangyou Zhang. Finite element approximations of symmetric tensors on simplicial grids in n\mathbb{R}^{n}: the lower order case. Math. Models Methods Appl. Sci., 26(9):1649–1669, 2016.
  • [18] Xuehai Huang, Chao Zhang, Yaqian Zhou, and Yangxing Zhu. New low-order mixed finite element methods for linear elasticity. Adv. Comput. Math., 50(2):Paper No. 17, 31, 2024.
  • [19] Jan Jaśkowiec and N. Sukumar. High-order symmetric cubature rules for tetrahedra and pyramids. Internat. J. Numer. Methods Engrg., 122(1):148–171, 2021.
  • [20] Claes Johnson. On the convergence of a mixed finite-element method for plate bending problems. Numer. Math., 21:43–62, 1973.
  • [21] Astrid Pechstein and Joachim Schöberl. Tangential-displacement and normal-normal-stress continuous mixed finite elements for elasticity. Math. Models Methods Appl. Sci., 21(8):1761–1782, 2011.
  • [22] Astrid Pechstein and Joachim Schöberl. Anisotropic mixed finite elements for elasticity. Internat. J. Numer. Methods Engrg., 90(2):196–217, 2012.
  • [23] Astrid S. Pechstein and Joachim Schöberl. An analysis of the TDNNS method using natural norms. Numer. Math., 139(1):93–120, 2018.
  • [24] Luc Tartar. An introduction to Sobolev spaces and interpolation spaces, volume 3 of Lecture Notes of the Unione Matematica Italiana. Springer, Berlin; UMI, Bologna, 2007.
  • [25] F. J. Vermolen and A. Segal. On an integration rule for products of barycentric coordinates over simplexes in n\mathbb{R}^{n}. J. Comput. Appl. Math., 330:289–294, 2018.