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

A tangential and penalty-free finite element method for the surface Stokes problem

Alan Demlow and Michael Neilan Department of Mathematics, Texas A&M University, College Station, TX, 77843 [email protected] Department of Mathematics, University of Pittsburgh, Pittsburgh, PA 15260 [email protected]
Abstract.

Surface Stokes and Navier-Stokes equations are used to model fluid flow on surfaces. They have attracted significant recent attention in the numerical analysis literature because approximation of their solutions poses significant challenges not encountered in the Euclidean context. One challenge comes from the need to simultaneously enforce tangentiality and H1H^{1} conformity (continuity) of discrete vector fields used to approximate solutions in the velocity-pressure formulation. Existing methods in the literature all enforce one of these two constraints weakly either by penalization or by use of Lagrange multipliers. Missing so far is a robust and systematic construction of surface Stokes finite element spaces which employ nodal degrees of freedom, including MINI, Taylor-Hood, Scott-Vogelius, and other composite elements which can lead to divergence-conforming or pressure-robust discretizations. In this paper we construct surface MINI spaces whose velocity fields are tangential. They are not H1H^{1}-conforming, but do lie in H(div)H({\rm div}) and do not require penalization to achieve optimal convergence rates. We prove stability and optimal-order energy-norm convergence of the method and demonstrate optimal-order convergence of the velocity field in L2L_{2} via numerical experiments. The core advance in the paper is the construction of nodal degrees of freedom for the velocity field. This technique also may be used to construct surface counterparts to many other standard Euclidean Stokes spaces, and we accordingly present numerical experiments indicating optimal-order convergence of nonconforming tangential surface Taylor-Hood 21\mathbb{P}^{2}-\mathbb{P}^{1} elements.

Key words and phrases:
surface Stokes equation; finite element method; MINI element
2000 Mathematics Subject Classification:
65N12, 65N15, 65N30
The first author was partially supported by NSF grant DMS-2012326. The second author was partially supported by NSF grants DMS-2011733 and DMS-2309425

1. Introduction

In this paper, we consider the surface Stokes problem:

(1.1a) 𝚷divγDefγ𝒖+γp+𝒖\displaystyle-\bm{\Pi}{\rm div}_{\gamma}{\rm Def}_{\gamma}\bm{u}+\nabla_{\gamma}p+\bm{u} =𝒇on γ,\displaystyle={\bm{f}}\qquad\text{on }\gamma,
(1.1b) divγ𝒖\displaystyle{\rm div}_{\gamma}\bm{u} =0on γ.\displaystyle=0\qquad\text{on }\gamma.

Here, γ3\gamma\subset\mathbb{R}^{3} is a smooth and connected two-dimensional surface with outward unit normal 𝝂{\bm{\nu}}, 𝚷=𝐈𝝂𝝂\bm{\Pi}={\bf I}-{\bm{\nu}}\otimes{\bm{\nu}} is the orthogonal projection onto the tangent space of γ\gamma, and γ\nabla_{\gamma} and divγ{\rm div}_{\gamma} are the surface gradient and surface divergence operators, respectively. Furthermore, Defγ{\rm Def}_{\gamma} is the tangential deformation operator, and the forcing function 𝒇{\bm{f}} is assumed to be tangential to the surface to ensure well-posedness. Further assumptions and notation are given in Section 2; cf. [20] for derivation of the surface Stokes problem and related models and further discussion of their properties. The system of equations (1.1) is subject to the tangential velocity constraint 𝒖𝝂=0\bm{u}\cdot{\bm{\nu}}=0. To address degeneracies related to Killing fields, i.e., non-trivial tangential vector fields in the kernel of Defγ{\rm Def}_{\gamma}, we include a zeroth-order mass term in the momentum equations (1.1a) (cf. Remark 4.1).

We consider surface finite element methods (SFEMs), a natural methodology mimicking the variational formulation and built upon classical Galerkin principles. In this approach the domain γ\gamma is approximated by a polyhedral (or higher-order) surface Γh\Gamma_{h} whose faces constitute the finite element mesh. Similar to the Euclidean setting, SFEMs for the surface Stokes problem based on the standard velocity-pressure formulation must use compatible discrete spaces. Specifically, a discrete inf-sup condition must be satisfied. Given that SFEMs utilize the same framework as their Euclidean counterparts, employing mappings via affine or polynomial diffeomorphisms, one may anticipate that numerous classical inf-sup stable Stokes pairs can be adapted to their surface analogues, readily enabling the construction of stable SFEMs for (1.1).

However, the tangential velocity constraint poses a significant hurdle to constructing stable and convergent SFEMs. As Γh\Gamma_{h} is merely Lipschitz continuous, its outward unit normal is discontinuous at mesh edges and vertices. As a result, the tangential projection of continuous, piecewise smooth functions does not lead to 𝑯1\bm{H}^{1}-conforming functions. Moreover, there do not exist canonical, degree-of-freedom-preserving pullbacks for tangential 𝑯1\bm{H}^{1} vector fields, in particular, the Piola transform preserves tangentiality and in-plane normal continuity, but not in-plane tangential continuity. Finally, a continuous, tangential, and piecewise smooth vector field on Γh\Gamma_{h} must necessarily vanish on mesh corners except in exceptional cases where all incident triangles are coplanar. Indeed, at a mesh corner there are at least three faces emanating from a common vertex, whose outward unit normal vectors span 3\mathbb{R}^{3}. Therefore tangentiality of a continuous vector field with respect to each of the three planes implies that it vanishes at the vertex. Thus any piecewise polynomial space simultaneously satisfying both tangentiality and continuity exhibits a locking-type phenomenon with poor approximation properties.

There is a substantial recent literature on numerical approximation of the surface Stokes and related problems such as the surface vector Laplace equation. Most of these circumvent the difficulties described above in one of three ways: by relaxing the pointwise tangential constraint, by relaxing 𝑯1\bm{H}^{1}-conformity of the finite element space, or by using a different formulation of the surface Stokes problem. For the former, one can weakly impose the tangential constraint via penalization or Lagrange multipliers [16, 17, 25, 26, 18, 21, 4]. In principle, this allows one to use inf-sup stable Euclidean Stokes pairs to solve the analogous surface problem. However, this methodology requires superfluous degrees of freedom, as the velocity space is approximated by arbitrary vectors in 3\mathbb{R}^{3} rather than tangential vectors. In addition an unnatural high-order geometric approximation of the unit normal of the true surface is needed to obtain optimal-order approximations. Therefore for problems in which full information of the exact surface is unknown (e.g., free-boundary problem), these penalization schemes lead to SFEMs with sub-optimal convergence properties. However, it was shown recently in [19] that the tangential component of the solution converges optimally for a standard isoparametric geometry approximation in most cases assuming a correct choice of penalty parameters. The only exception is the case where tangential L2L_{2} errors are considered along with affine (polyhedral) surface approximations. Alternatively, one may relax 𝑯1\bm{H}^{1}-conformity and use finite element trial and test functions that are not continuous on the discrete surface Γh\Gamma_{h}. In this direction, SFEMs utilizing tangentially- and 𝑯(div)\bm{H}({\rm div})-conforming finite element spaces such as Raviart-Thomas and Brezzi-Douglas-Marini combined with discontinuous Galerkin techniques are proposed and analyzed in [3, 23]; cf. [10] for similar methods for Euclidean Stokes equations. Here, additional consistency, symmetry, and stability terms are added to the method. These terms add some complexity to the implementation, especially for higher-order surface approximations, but are standard in the context of discontinuous Galerkin methods. Optimal-order convergence is observed experimentally for a standard SFEM formulation that does not require higher-order approximations of any geometric information. Discretizations of stream function formulations of the surface Stokes equations have also appeared in the literature [24, 28, 5, 4]. However, as with methods weakly enforcing tangentiality, they require higher-order approximation to the surface normal and in addition require computation of curvature information which can in and of itself be a challenging problem. These methods are also restricted to simply connected surfaces. As a final note, trace SFEMs, in which discretizations of surface PDE are formulated with respect to a background 3D mesh and a corresponding 3D finite element space, are especially important in the context of dynamic surface fluid computations. Trace formulations are well-developed for H1H^{1} conforming/tangentially nonconforming methods and stream function formulations, but have not yet appeared for 𝑯(div)\bm{H}({\rm div})-conforming methods.

In this paper, we design a SFEM for the surface Stokes problem (1.1) using a strongly tangential finite element space that is based on a conforming, inf-sup stable Euclidean pair. The method is based on the standard variational formulation for the Stokes problem and does not require additional consistency terms or extrinsic penalization. As far as we are aware, this is the first SFEM for the surface Stokes problem with these properties. The key issue that we address is the assignment of degrees of freedom (DOFs) of tangential vector fields at Lagrange nodes, in particular, at vertices of the surface triangulation.

To expand on this last point and to describe our proposed approach, consider a vertex/DOF, call aa, of the triangulation of the discrete geometry approximation Γh\Gamma_{h}, and let 𝒯a\mathcal{T}_{a} denote the set of faces in the triangulation that have aa as a vertex. We wish to interpret and define the values of tangential vector fields forming our finite element space at this vertex in a way that ensures the resulting discrete spaces have desirable approximation and weak-continuity properties. As the mesh elements in 𝒯h\mathcal{T}_{h} generally lie in different planes, it is immediate that such vector fields are generally multi-valued at aa.

Let 𝒑=𝒑γ\bm{p}=\bm{p}_{\gamma} denote the closest point projection onto γ\gamma, and note that, because γ\gamma is smooth, continuous and tangential vector fields are well-defined and single-valued at 𝒑(a)\bm{p}(a). Thus, as the Piola transform preserves tangentiality, a natural assignment is to construct finite element functions 𝒗\bm{v} with the property 𝒗|K(a)=𝒫𝒑1\wideparen𝒗|𝒑(K)K𝒯a\bm{v}|_{K}(a)=\mathcal{P}_{\bm{p}^{-1}}\wideparen{\bm{v}}\big{|}_{\bm{p}(K)}\ \forall K\in\mathcal{T}_{a} for some vector field \wideparen𝒗\wideparen{\bm{v}} tangent at 𝒑(a)\bm{p}(a), where 𝒫𝒑1\mathcal{P}_{\bm{p}^{-1}} is the Piola transform of the inverse mapping 𝒑1:γΓh\bm{p}^{-1}:\gamma\to\Gamma_{h}; see Figure 1. Imposing this condition on Lagrange finite element DOFs likely leads to the sought out approximation and weak-continuity properties, and thus, conceptually may lead to convergent SFEMs for (1.1). However, the implementation of the resulting finite element method requires explicit information about the exact surface γ\gamma and its closest point projection. Therefore this construction is of little practical value.

Instead of this idealized construction, we fix an arbitrary face Ka𝒯aK_{a}\in\mathcal{T}_{a}. Given the value 𝒗|Ka(a)\bm{v}|_{K_{a}}(a) and K𝒯aK\in\mathcal{T}_{a}, we then assign 𝒗|K(a)=\EuScriptP𝒑Ka1𝒗|Ka(a)\bm{v}|_{K}(a)=\EuScript{P}_{\bm{p}^{-1}_{K_{a}}}\bm{v}|_{K_{a}}(a), the Piola transform of 𝒗|Ka(a)\bm{v}|_{K_{a}}(a) with respect to the inverse of the closest point projection onto the plane containing KaK_{a}; see Figure 2. This transform is linear with a relatively simple formula (cf. Definition 2.3), and it only uses geometric information from Γh\Gamma_{h}. Moreover, we show that this construction is only an O(h2)O(h^{2}) perturbation from the idealized setting. As a result, the constructed finite element spaces possess sufficient weak continuity properties to ensure that the resulting scheme is convergent for the surface Stokes problem (1.1).

To clearly communicate the main ideas and to keep technicalities at minimum, we focus on a polyhedral approximation to γ\gamma and on the lowest-order MINI pair, which in the Euclidean setting takes the discrete velocity space to be the (vector-valued) linear Lagrange space enriched with cubic bubbles, and the discrete pressure space to be the (scalar) Lagrange space. We expect the main ideas to be applicable to other finite element pairs (e.g, Taylor–Hood, Scott-Vogelius [29, 22, 31, 2]), although the stability must be shown on a case-by-case basis. Below we present numerical experiments demonstrating the viability of our approach for 2\mathbb{P}^{2} surface approximations paired with a 21\mathbb{P}^{2}-\mathbb{P}^{1} Taylor-Hood finite element space and plan to address generalizations of our approach more fully in future works.

The rest of the paper is organized as follows. In the next section, we introduce the notation and provide some preliminary results. In Section 3, we define the surface finite element spaces based on the classical MINI pair. Here, we show that the spaces have optimal-order approximation properties and are inf-sup stable. We also establish weak continuity properties of the discrete velocity space via an H1H^{1}-conforming relative on the true surface. In Section 4, we define the finite element space and prove optimal-order estimates in the energy norm. Finally in Section 5 we provide numerical experiments which support the theoretical results.

γ\gamma𝒑(a)\bm{p}(a)\wideparen𝒗\wideparen{\bm{v}}Γh\Gamma_{h}aaKKKaK_{a}𝒫𝒑1\wideparen𝒗|𝒑(K)\mathcal{P}_{\bm{p}^{-1}}\wideparen{\bm{v}}|_{\bm{p}(K)}𝒫𝒑1\wideparen𝒗|𝒑(Ka)\mathcal{P}_{\bm{p}^{-1}}\wideparen{\bm{v}}|_{\bm{p}(K_{a})}
Figure 1. A pictorial description of an idealized assignment of nodal values on a one-dimensional surface. Here, a tangential vector \wideparen𝒗\wideparen{\bm{v}} at the point 𝒑(a)\bm{p}(a) is mapped to each element KK via the Piola transform with respect to the inverse mapping 𝒑1|K\bm{p}^{-1}|_{K}.
Γh\Gamma_{h}aaKKKaK_{a}𝒗|Ka(a)\bm{v}|_{K_{a}}(a)𝒫𝒑Ka1𝒗|Ka(a)\mathcal{P}_{\bm{p}_{K_{a}}^{-1}}\bm{v}|_{K_{a}}(a)
Figure 2. A pictorial description of our construction on a one-dimensional surface. The value of a vector field 𝒗\bm{v} at vertex aa restricted to KaK_{a} is mapped to KK via the Piola transform with respect to the inverse of the closest point projection onto the plane containing KaK_{a}.

2. Notation and Preliminaries

We assume γ3\gamma\subset\mathbb{R}^{3} is a smooth, connected, and orientable two-dimensional surface without boundary. The signed distance function of γ\gamma is denoted by dd, which satisfies d<0d<0 in the interior of γ\gamma and d>0d>0 in the exterior. We set 𝝂(x)=d(x){\bm{\nu}}(x)=\nabla d(x) to be the outward-pointing unit normal (where the gradient is understood as a column vector) and 𝐇(x)=D2d(x){\bf H}(x)=D^{2}d(x) the Weingarten map. The tangential projection operator is 𝚷=𝐈𝝂𝝂\bm{\Pi}={\bf I}-{\bm{\nu}}\otimes{\bm{\nu}}, where 𝐈{\bf I} is the 3×33\times 3 identity matrix, and the outer product of two vectors 𝒂{\bm{a}} and 𝒃{\bm{b}} satisfies (𝒂𝒃)i,j=aibj({\bm{a}}\otimes{\bm{b}})_{i,j}=a_{i}b_{j}. The smoothness of γ\gamma ensures the existence of δ>0\delta>0 sufficiently small such that the closest point projection

𝒑(x):=xd(x)𝝂(x)\bm{p}(x):=x-d(x){\bm{\nu}}(x)

is well defined in the tubular region U={x3:dist(x,γ)δ}U=\{x\in\mathbb{R}^{3}:\ {\rm dist}(x,\gamma)\leq\delta\}.

For a scalar function q:γq:\gamma\to\mathbb{R} we define its extension qe:Uq^{e}:U\to\mathbb{R} via qe=q𝒑q^{e}=q\circ\bm{p}. Likewise, for 𝒗=(v1,v2,v3):γ3\bm{v}=(v_{1},v_{2},v_{3})^{\intercal}:\gamma\to\mathbb{R}^{3} its extension 𝒗e:U3\bm{v}^{e}:U\to\mathbb{R}^{3} satisfies (𝒗e)i=vie(\bm{v}^{e})_{i}=v_{i}^{e} for i=1,2,3i=1,2,3. Define the surface gradient γq=𝚷qe\nabla_{\gamma}q=\bm{\Pi}\nabla q^{e}, and for a (column) vector field 𝒗=(v1,v2,v3):γ3\bm{v}=(v_{1},v_{2},v_{3})^{\intercal}:\gamma\to\mathbb{R}^{3}, we let 𝒗e=(v1e,v2e,v3e)\nabla\bm{v}^{e}=(\nabla v_{1}^{e},\nabla v_{2}^{e},\nabla v_{3}^{e})^{\intercal} denote the Jacobian matrix of 𝒗e\bm{v}^{e}. We then see (𝒗e𝚷)i,:=((vie)𝚷)=(𝚷vie)=(γvi)(\nabla\bm{v}^{e}\bm{\Pi})_{i,:}=((\nabla v^{e}_{i})^{\intercal}\bm{\Pi})=(\bm{\Pi}\nabla v^{e}_{i})^{\intercal}=(\nabla_{\gamma}v_{i})^{\intercal}, i.e., the ithith row of 𝒗e𝚷\nabla\bm{v}^{e}\bm{\Pi} coincides with (γvi)(\nabla_{\gamma}v_{i})^{\intercal}. The tangential surface gradient (covariant derivative) of 𝒗\bm{v} is defined by γ𝒗=𝚷𝒗e𝚷\nabla_{\gamma}\bm{v}=\bm{\Pi}\nabla\bm{v}^{e}\bm{\Pi}, and the surface divergence operator of 𝒗\bm{v} is divγ𝒗=tr(γ𝒗){\rm div}_{\gamma}\bm{v}={\rm tr}(\nabla_{\gamma}\bm{v}). The deformation of a tangential vector field is defined as the symmetric part of its surface gradient, i.e.,

Defγ𝒗=12(γ𝒗+(γ𝒗)).{\rm Def}_{\gamma}\bm{v}=\frac{1}{2}\big{(}\nabla_{\gamma}\bm{v}+(\nabla_{\gamma}\bm{v})^{\intercal}\big{)}.

For a matrix field 𝐀:3×3{\bf A}:\mathbb{R}^{3\times 3}, the divergence divγ𝐀{\rm div}_{\gamma}{\bf A} is understood to act row-wise.

Let L2(γ)L_{2}(\gamma) denote the space of square-integrable functions on γ\gamma and let L̊2(γ)\mathring{L}_{2}(\gamma) be the subspace of L2(γ)L_{2}(\gamma) consisting of L2L_{2}-functions with vanishing mean. We let Wpm(γ)W_{p}^{m}(\gamma) be the Sobolev space of order mm and exponent pp on γ\gamma with corresponding norm Wpm(γ)\|\cdot\|_{W^{m}_{p}(\gamma)}. We use the notation Hm(γ)=W2m(γ)H^{m}(\gamma)=W_{2}^{m}(\gamma) with Hm(γ)=W2m(γ)\|\cdot\|_{H^{m}(\gamma)}=\|\cdot\|_{W^{m}_{2}(\gamma)}, and the convention ||H0=L2|\cdot|_{H^{0}}=\|\cdot\|_{L_{2}}, ||Wp0=Lp|\cdot|_{W^{0}_{p}}=\|\cdot\|_{L_{p}}. Analogous vector-valued spaces are denoted in boldface (e.g., 𝑳2(γ)=(L2(γ))3\bm{L}_{2}(\gamma)=(L_{2}(\gamma))^{3} and 𝑯1(γ)=(H1(γ))3\bm{H}^{1}(\gamma)=(H^{1}(\gamma))^{3}). We let 𝑯T1(γ)\bm{H}_{T}^{1}(\gamma) be the subspace of 𝑯1(γ)\bm{H}^{1}(\gamma) whose members are tangent to γ\gamma, and set

𝑯(divγ;γ)={𝒗𝑳2(γ):divγ𝒗L2(γ)}.\bm{H}({\rm div}_{\gamma};\gamma)=\{\bm{v}\in\bm{L}_{2}(\gamma):\ {\rm div}_{\gamma}\bm{v}\in L_{2}(\gamma)\}.

Let Γh\Gamma_{h} be a polyhedral surface approximation of γ\gamma with triangular faces. We assume that Γh\Gamma_{h} is an O(h2)O(h^{2}) approximation in the sense that d(x)=O(h2)d(x)=O(h^{2}) for all xΓhx\in\Gamma_{h}. We further assume hh is sufficiently small to ensure ΓhU\Gamma_{h}\subset U, in particular, the closest point projection is well-defined on Γh\Gamma_{h}. We denote by 𝒯h\mathcal{T}_{h} the set of faces of Γh\Gamma_{h}, and assume this triangulation is shape-regular (i.e., the ratio of the diameters of the inscribed and circumscribed circles of each face is uniformly bounded). For simplicity and to ease the presentation, we further assume that 𝒯h\mathcal{T}_{h} is quasi-uniform, i.e., h:=maxKdiam(K)diam(K)h:=\max_{K^{\prime}}{\rm diam}(K^{\prime})\approx{\rm diam}(K) for all K𝒯hK\in\mathcal{T}_{h}. The image of the mesh elements and the resulting set on the exact surface are given, respectively, by

Kγ=𝒑(K),𝒯hγ={𝒑(K):K𝒯h}.K^{\gamma}=\bm{p}(K),\qquad\mathcal{T}_{h}^{\gamma}=\{\bm{p}(K):\ K\in\mathcal{T}_{h}\}.

We use the notation aba\lesssim b (resp., aba\gtrsim b) if there exists a constant C>0C>0 independent of the mesh parameter hh such that aCba\leq Cb (resp., aCba\geq Cb). The statement aba\approx b means aba\lesssim b and aba\gtrsim b.

Set 𝒱h\mathcal{V}_{h} to be the set of vertices in 𝒯h\mathcal{T}_{h}, and for each K𝒯hK\in\mathcal{T}_{h}, let 𝒱K\mathcal{V}_{K} denote the set of three vertices of KK. For each a𝒱ha\in\mathcal{V}_{h}, let 𝒯a𝒯h\mathcal{T}_{a}\subset\mathcal{T}_{h} denote the set of faces having aa as a vertex. For K𝒯hK\in\mathcal{T}_{h}, we define the patches

ωK=K𝒯hK¯K¯K,ωK=K𝒯hK¯ω¯KK,\omega_{K}=\mathop{\bigcup_{K^{\prime}\in\mathcal{T}_{h}}}_{\bar{K}^{\prime}\cap\bar{K}\neq\emptyset}K^{\prime},\qquad\omega_{K}^{\prime}=\mathop{\bigcup_{K^{\prime}\in\mathcal{T}_{h}}}_{\bar{K}^{\prime}\cap\overline{\omega}_{K}\neq\emptyset}K^{\prime},

so that ωKωKΓh\omega_{K}\subset\omega_{K}^{\prime}\subset\Gamma_{h}. The patches ωKγ\omega_{K^{\gamma}} and ωKγ\omega^{\prime}_{K^{\gamma}} associated with Kγ=𝒑(K)K^{\gamma}=\bm{p}(K) are defined analogously.

The (piecewise constant) outward unit normal of Γh\Gamma_{h} is denoted by 𝝂h{\bm{\nu}}_{h}, and we shall use the notation 𝝂K=𝝂h|K3{\bm{\nu}}_{K}={\bm{\nu}}_{h}|_{K}\in\mathbb{R}^{3}, its restriction to K𝒯hK\in\mathcal{T}_{h}. We assume that |𝝂𝒑𝝂h|h|{\bm{\nu}}\circ\bm{p}-{\bm{\nu}}_{h}|\lesssim h. The tangential projection with respect to Γh\Gamma_{h} is 𝚷h=𝐈𝝂h𝝂h\bm{\Pi}_{h}={\bf I}-{\bm{\nu}}_{h}\otimes{\bm{\nu}}_{h}, and we assume there exists c>0c>0 independent of hh such that 𝝂𝝂hc>0{\bm{\nu}}\cdot{\bm{\nu}}_{h}\geq c>0 on Γh\Gamma_{h}. We let μh(x)\mu_{h}(x) satisfy μhdσh(x)=dσ(𝒑(x))\mu_{h}d\sigma_{h}(x)=d\sigma(\bm{p}(x)), where dσd\sigma and dσhd\sigma_{h} are surface measures of γ\gamma and Γh\Gamma_{h}, respectively. In particular,

Γh(q𝒑)μh=γqqL1(γ).\int_{\Gamma_{h}}(q\circ\bm{p})\mu_{h}=\int_{\gamma}q\qquad\forall q\in{L_{1}(\gamma)}.

From [11, Proposition 2.5], we have

(2.1) μh(x)=𝝂(x)𝝂h(x)i=12(1d(x)κi(x))xΓh,\mu_{h}(x)={\bm{\nu}}(x)\cdot{\bm{\nu}}_{h}(x)\prod_{i=1}^{2}(1-d(x)\kappa_{i}(x))\qquad x\in\Gamma_{h},

and

(2.2) |1μh(x)|h2,|1-\mu_{h}(x)|\lesssim h^{2},

where {κ1,κ2}\{\kappa_{1},\kappa_{2}\} are the eigenvalues of 𝐇{\bf H}, whose corresponding eigenvectors are orthogonal to 𝝂{\bm{\nu}}. We set μK=μh|K\mu_{K}=\mu_{h}|_{K} to be the restriction of μh\mu_{h} to K𝒯hK\in\mathcal{T}_{h}.

Surface differential operators with respect to Γh\Gamma_{h} are denoted and defined analogously to those on γ\gamma. We also set (mm\in\mathbb{N})

𝑯hm(Γh)={𝒗𝑳2(Γh):𝒗|K𝑯m(K)K𝒯h},𝒗Hhm(K)2=K𝒯h𝒗Hm(K)2\bm{H}^{m}_{h}(\Gamma_{h})=\{\bm{v}\in\bm{L}_{2}(\Gamma_{h}):\ \bm{v}|_{K}\in\bm{H}^{m}(K)\ \forall K\in\mathcal{T}_{h}\},\qquad\|\bm{v}\|_{H^{m}_{h}(K)}^{2}=\sum_{K\in\mathcal{T}_{h}}\|\bm{v}\|_{H^{m}(K)}^{2}

to be the piecewise HmH^{m} Sobolev space and norm, respectively. Likewise, 𝑯hm(γ)\bm{H}^{m}_{h}(\gamma) is the piecewise Sobolev space with respect to 𝒯hγ\mathcal{T}_{h}^{\gamma} with corresponding norm 𝒗Hhm(γ)2=K𝒯h𝒗Hm(Kγ)2\|\bm{v}\|_{H^{m}_{h}(\gamma)}^{2}=\sum_{K\in\mathcal{T}_{h}}\|\bm{v}\|_{H^{m}(K^{\gamma})}^{2}, and Defγ,h{\rm Def}_{\gamma,h} denotes the piecewise deformation operator with respect to 𝒯hγ\mathcal{T}_{h}^{\gamma}.

We end this section by stating a well-known characterization of 𝑯(divΓh;Γh)={𝒗𝑳2(Γh):divΓh𝒗L2(Γh)}\bm{H}({\rm div}_{\Gamma_{h}};\Gamma_{h})=\{\bm{v}\in\bm{L}_{2}(\Gamma_{h}):\ {\rm div}_{\Gamma_{h}}\bm{v}\in L_{2}(\Gamma_{h})\}. For each edge ee of the mesh, denote by K1e,K2e𝒯hK_{1}^{e},K_{2}^{e}\in\mathcal{T}_{h} the two triangles in the mesh such that e=K1eK2ee={\partial}K_{1}^{e}\cap{\partial}K_{2}^{e}. Let 𝒏je\bm{n}_{j}^{e} denote the outward in-plane normal to Kje{\partial}K_{j}^{e}, and note that in general, 𝒏1e𝒏2e\bm{n}^{e}_{1}\neq-\bm{n}^{e}_{2} on ee. Then a vector field 𝒗𝑯h1(Γh)\bm{v}\in\bm{H}^{1}_{h}(\Gamma_{h}) satisfies 𝒗𝑯(divΓh;Γh)\bm{v}\in\bm{H}({\rm div}_{\Gamma_{h}};\Gamma_{h}) if and only if [23]

(2.3) 𝒗1𝒏1e|e+𝒗2𝒏2e|e=0for all edges e,\bm{v}_{1}\cdot\bm{n}_{1}^{e}|_{e}+\bm{v}_{2}\cdot\bm{n}_{2}^{e}|_{e}=0\quad\text{for all edges }e,

where 𝒗j=𝒗|Kje\bm{v}_{j}=\bm{v}|_{K_{j}^{e}}.

2.1. Extensions and lifts

For the rest of the paper, we view the closest point projection as a mapping from the discrete surface approximation to the true surface, i.e., 𝒑:Γhγ\bm{p}:\Gamma_{h}\to\gamma. Restricted to Γh\Gamma_{h} the projection is a bijection, and in particular has a well-defined inverse: 𝒑1:γΓh\bm{p}^{-1}:\gamma\to\Gamma_{h}. Recall that for a scalar or vector-valued function qq on γ\gamma, its extension (now to Γh\Gamma_{h}) is qe=q𝒑q^{e}=q\circ\bm{p}. For a scalar or vector-valued function qq defined on Γh\Gamma_{h}, we define its lift via q=q𝒑1q^{\ell}=q\circ\bm{p}^{-1}. Note that (q)e=q(q^{\ell})^{e}=q on γ\gamma and likewise, (qe)=q(q^{e})^{\ell}=q on Γh\Gamma_{h}. For qHhm(γ)(m=0,1,2)q\in H^{m}_{h}(\gamma)\ (m=0,1,2), there holds

(2.4) qHm(Kγ)qeHm(K)K𝒯h,\|q\|_{H^{m}(K^{\gamma})}\approx\|q^{e}\|_{H^{m}(K)}\qquad\forall K\in\mathcal{T}_{h},

which follows from a change of variables, the chain rule, and the smoothness assumptions of γ\gamma (cf. [13]).

2.2. Surface Piola transforms

Following [30, 9, 3] we summarize the divergence-conforming Piola transform with respect to a mapping between surfaces. Let 𝒮0\mathcal{S}_{0} and 𝒮1\mathcal{S}_{1} be two sufficiently smooth surfaces, and let Φ:𝒮0𝒮1\Phi:\mathcal{S}_{0}\to\mathcal{S}_{1} be a diffeomorphism with inverse Φ1:𝒮1𝒮0\Phi^{-1}:\mathcal{S}_{1}\to\mathcal{S}_{0}. Let dσid\sigma_{i} be the surface measure of 𝒮i\mathcal{S}_{i}, and let μ\mu formally satisfy μdσ0=dσ1\mu d\sigma_{0}=d\sigma_{1}. Then the Piola transform of a vector field 𝒗:𝒮03\bm{v}:\mathcal{S}_{0}\to\mathbb{R}^{3} with respect to Φ\Phi is given by

(𝒫Φ𝒗)Φ=μ1DΦ𝒗.(\mathcal{P}_{\Phi}\bm{v})\circ\Phi=\mu^{-1}D\Phi\bm{v}.

Likewise, for 𝒗:𝒮13\bm{v}:\mathcal{S}_{1}\to\mathbb{R}^{3} its Piola transform with respect to Φ1\Phi^{-1} is

(𝒫Φ1𝒗)Φ1=(μΦ1)DΦ1𝒗.(\mathcal{P}_{\Phi^{-1}\bm{v})\circ\Phi^{-1}=(\mu\circ\Phi^{-1})D\Phi^{-1}\bm{v}.}

Similar to the Euclidean setting, there holds

(2.5) div𝒮0𝒗=μdiv𝒮1𝒫Φ𝒗𝒗𝑯(div;𝒮0),{\rm div}_{\mathcal{S}_{0}}\bm{v}=\mu{\rm div}_{\mathcal{S}_{1}}\mathcal{P}_{\Phi}\bm{v}\qquad\forall\bm{v}\in\bm{H}({\rm div};\mathcal{S}_{0}),

in particular, 𝒫Φ:𝑯(div𝒮0;𝒮0)𝑯(div𝒮1;𝒮1)\mathcal{P}_{\Phi}:\bm{H}({\rm div}_{\mathcal{S}_{0}};\mathcal{S}_{0})\to\bm{H}({\rm div}_{\mathcal{S}_{1}};\mathcal{S}_{1}) and 𝒫Φ1:𝑯(div𝒮1;𝒮1)𝑯(div𝒮0;𝒮0)\mathcal{P}_{\Phi^{-1}}:\bm{H}({\rm div}_{\mathcal{S}_{1}};\mathcal{S}_{1})\to\bm{H}({\rm div}_{\mathcal{S}_{0}};\mathcal{S}_{0}) are bounded mappings. Moreover, as DΦD\Phi and DΦ1D\Phi^{-1} are tangent maps, the Piola transform yields tangential vector fields: if 𝝂j{\bm{\nu}}_{j} is the unit normal of the surface 𝒮j\mathcal{S}_{j}, then (𝒫Φ𝒗)𝝂1=0(\mathcal{P}_{\Phi}\bm{v})\cdot{\bm{\nu}}_{1}=0 on 𝒮1\mathcal{S}_{1} and (𝒫Φ1𝒗)𝝂0=0(\mathcal{P}_{\Phi^{-1}}\bm{v})\cdot{\bm{\nu}}_{0}=0 on 𝒮0\mathcal{S}_{0}.

In the case Φ=𝒑\Phi=\bm{p}, 𝒮0=Γh\mathcal{S}_{0}=\Gamma_{h}, and 𝒮1=γ\mathcal{S}_{1}=\gamma (so that μ=μh\mu=\mu_{h}), the Piola transform of 𝒗:Γh3\bm{v}:\Gamma_{h}\to\mathbb{R}^{3} with 𝚷h𝒗=𝒗\bm{\Pi}_{h}\bm{v}=\bm{v} is [9, 3]

(2.6) \wideparen𝒗𝒑:=𝒫𝒑𝒗=1μh[𝚷d𝐇]𝒗,\wideparen\bm{v}\circ\bm{p}:=\mathcal{P}_{\bm{p}}\bm{v}=\frac{1}{\mu_{h}}\big{[}\bm{\Pi}-d{\bf H}\big{]}\bm{v},

whereas the Piola transform of 𝒗:γ3\bm{v}:\gamma\to\mathbb{R}^{3} with respect to the inverse 𝒑1\bm{p}^{-1} is given by

(2.7) 𝒗˘:=𝒫𝒑1𝒗=μh[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1(𝒗𝒑).\breve{\bm{v}}:=\mathcal{P}_{\bm{p}^{-1}}\bm{v}=\mu_{h}\Big{[}{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\Big{]}[{\bf I}-d{\bf H}]^{-1}(\bm{v}\circ\bm{p}).

Note that \wideparen𝒗˘=𝒗\wideparen{\breve{\bm{v}}}=\bm{v} on γ\gamma and \wideparen𝒗˘=𝒗\breve{\wideparen{\bm{v}}}=\bm{v} on Γh\Gamma_{h}. Moreover, it follows from (2.5) that for all K𝒯hK\in\mathcal{T}_{h},

(2.8a) Kγ(divγ𝒗)q\displaystyle\int_{K^{\gamma}}({\rm div}_{\gamma}\bm{v})q =K(divΓh𝒗˘)qe𝒗𝑯(divγ;Kγ),qL2(Kγ).\displaystyle=\int_{K}({\rm div}_{\Gamma_{h}}\breve{\bm{v}})q^{e}\qquad\forall\bm{v}\in\bm{H}({\rm div}_{\gamma};K^{\gamma}),\ q\in{L_{2}(K^{\gamma})}.
and
(2.8b) K(divΓh𝒗)q\displaystyle\int_{K}({\rm div}_{\Gamma_{h}}\bm{v})q =Kγ(divγ\wideparen𝒗)q𝒗𝑯(divΓh;K),qL2(K).\displaystyle=\int_{K^{\gamma}}({\rm div}_{\gamma}\wideparen\bm{v})q^{\ell}\qquad\forall\bm{v}\in\bm{H}({\rm div}_{\Gamma_{h}};K),\ q\in{L_{2}(K)}.

The following lemma states the equivalence of norms of vector fields and their Piola transforms.

Lemma 2.1.

For K𝒯hK\in\mathcal{T}_{h}, let 𝐯:Kγ3\bm{v}:K^{\gamma}\to\mathbb{R}^{3} and 𝐯˘=𝒫𝐩1𝐯:K3\breve{\bm{v}}=\mathcal{P}_{\bm{p}^{-1}}\bm{v}:K\to\mathbb{R}^{3} be related by (2.7) restricted to KK. Then if 𝐯𝐇m(Kγ)\bm{v}\in\bm{H}^{m}(K^{\gamma}) for some m{0,1,2}m\in\{0,1,2\}, there holds 𝐯˘Hm(K)\breve{\bm{v}}\in H^{m}(K). Moreover,

(2.9) 𝒗Hm(Kγ)𝒗˘Hm(K).\|\bm{v}\|_{H^{m}(K^{\gamma})}\approx\|\breve{\bm{v}}\|_{H^{m}(K)}.
Proof.

The proof for the cases m=0,1m=0,1 is found in [3, Lemma 4.1]. The case m=2m=2 follows from similar arguments and is therefore omitted. ∎

We also need a similar result that relates the L2{L_{2}} norm of the deformation tensors of 𝒗\bm{v} and its Piola transform \wideparen𝒗\wideparen{\bm{v}}. The proof of the following result is given in the appendix.

Lemma 2.2.

For K𝒯hK\in\mathcal{T}_{h}, let 𝐯𝐇T1(K)\bm{v}\in\bm{H}_{T}^{1}(K) and \wideparen𝐯𝐇T1(Kγ)\wideparen\bm{v}\in\bm{H}_{T}^{1}(K^{\gamma}) be related via \wideparen𝐯=𝒫𝐩𝐯\wideparen\bm{v}=\mathcal{P}_{\bm{p}}\bm{v}. Then

(2.10) |Defγ\wideparen𝒗(DefΓh𝒗)𝒑1|h(|(Γh𝒗)𝒑1|+|𝒗𝒑1|).\displaystyle|{\rm Def}_{\gamma}\wideparen\bm{v}-({\rm Def}_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}|\lesssim h\big{(}|(\nabla_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}|+|\bm{v}\circ\bm{p}^{-1}|\big{)}.

Consequently, by a change of variables,

(2.11) Defγ\wideparen𝒗L2(Kγ)DefΓh𝒗L2(K)+h(Γh𝒗L2(K)+𝒗L2(K)).\displaystyle\|{\rm Def}_{\gamma}\wideparen\bm{v}\|_{L_{2}(K^{\gamma})}\lesssim\|{\rm Def}_{\Gamma_{h}}\bm{v}\|_{L_{2}(K)}+h\big{(}\|\nabla_{\Gamma_{h}}\bm{v}\|_{L_{2}(K)}+\|\bm{v}\|_{L_{2}(K)}\big{)}.

We now apply the above definitions of Piola transforms to mappings between planes (surface triangles), which is critical to our construction of vertex degrees of freedom for vector fields on Γh\Gamma_{h}.

Definition 2.3.

For each vertex a𝒱ha\in\mathcal{V}_{h} in the triangulation, we arbitrarily choose a single (fixed) face Ka𝒯aK_{a}\in\mathcal{T}_{a}. For K𝒯aK\in\mathcal{T}_{a}, we define aK:33\mathcal{M}_{a}^{K}:\mathbb{R}^{3}\to\mathbb{R}^{3} by

(2.12) aK𝒙=(𝝂Ka𝝂K[𝐈𝝂Ka𝝂K𝝂Ka𝝂K])𝒙,\mathcal{M}_{a}^{K}\bm{x}=\Big{(}{\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}\Big{[}{\bf I}-\frac{{\bm{\nu}}_{K_{a}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}}\Big{]}\Big{)}\bm{x},

where we recall 𝝂Ka{\bm{\nu}}_{K_{a}} and 𝝂K{\bm{\nu}}_{K} are the outward unit normals of KaK_{a} and KK, respectively. In particular, aK𝒙\mathcal{M}_{a}^{K}\bm{x} is the Piola transform of 𝒙\bm{x} with respect to the inverse of the closest point projection onto the plane containing KaK_{a} (cf. (2.7)).

Remark 2.4.

By properties of the Piola transform, aK𝒙\mathcal{M}_{a}^{K}\bm{x} is tangential to KK, i.e., (aK𝒙)𝝂K=0(\mathcal{M}_{a}^{K}\bm{x})\cdot{\bm{\nu}}_{K}=0 for all x3x\in\mathbb{R}^{3}.

We next show that the “ideal” and “practical” interpretations of vectors at vertices discussed in the introduction (cf. Figures 1 and 2) do not differ by too much.

Lemma 2.5.

Fix a𝒱ha\in\mathcal{V}_{h} and let 𝐮\bm{u} lie in the tangent plane of γ\gamma at 𝐩(a)\bm{p}(a). For K𝒯aK\in\mathcal{T}_{a}, let 𝐮˘K=𝒫𝐩1𝐮|K\breve{\bm{u}}_{K}=\mathcal{P}_{\bm{p}^{-1}}\bm{u}|_{K} be the Piola transform of 𝐮\bm{u} to KK via the inverse of the closest point projection (cf. (2.7)). Then

(2.13) |𝒖˘KaK𝒖˘Ka|h2|𝒖˘Ka|h2|𝒖|.\displaystyle|\breve{\bm{u}}_{K}-\mathcal{M}_{a}^{K}\breve{\bm{u}}_{K_{a}}|\lesssim h^{2}|\breve{\bm{u}}_{K_{a}}|\leq h^{2}|\bm{u}|.
Proof.

Using 𝐇𝝂=12|𝝂|2=0{\bf H}{\bm{\nu}}=\frac{1}{2}\nabla|{\bm{\nu}}|^{2}=0, we have 𝐇𝚷=𝐇{\bf H}\bm{\Pi}={\bf H}. Note in addition that [𝐈𝝂𝝂K𝝂𝝂K]𝚷=[𝐈𝝂𝝂K𝝂𝝂K]\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]\bm{\Pi}=\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]. Therefore by (2.7), (2.6), and (2.1) we have

𝒖˘K=μK[𝐈𝝂𝝂K𝝂𝝂K][𝐈d𝐇]1𝒖=μK[𝐈𝝂𝝂K𝝂𝝂K][𝐈d𝐇]11μKa[𝚷d𝐇]𝒖˘Ka=μK[𝐈𝝂𝝂K𝝂𝝂K][𝐈d𝐇]11μKa[𝐈d𝐇]𝚷𝒖˘Ka=𝝂𝝂K𝝂𝝂Ka[𝐈𝝂𝝂K𝝂𝝂K]𝚷𝒖˘Ka=𝝂𝝂K𝝂𝝂Ka[𝐈𝝂𝝂K𝝂𝝂K]𝒖˘Ka.\displaystyle\begin{aligned} \breve{\bm{u}}_{K}&=\mu_{K}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]\left[{\bf I}-d{\bf H}\right]^{-1}\bm{u}\\ &=\mu_{K}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]\left[{\bf I}-d{\bf H}\right]^{-1}\frac{1}{\mu_{K_{a}}}\left[\bm{\Pi}-d{\bf H}\right]\breve{\bm{u}}_{K_{a}}\\ &=\mu_{K}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]\left[{\bf I}-d{\bf H}\right]^{-1}\frac{1}{\mu_{K_{a}}}\left[{\bf I}-d{\bf H}\right]\bm{\Pi}\breve{\bm{u}}_{K_{a}}\\ &=\frac{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K_{a}}}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]{\bf\Pi}\breve{\bm{u}}_{K_{a}}\\ &=\frac{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K_{a}}}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]\breve{\bm{u}}_{K_{a}}.\end{aligned}

Now |𝝂𝝂K|+|𝝂Ka𝝂K|h|{\bm{\nu}}-{\bm{\nu}}_{K}|+|{\bm{\nu}}_{K_{a}}-{\bm{\nu}}_{K}|\lesssim h, |1𝝂𝝂K|=12|𝝂𝝂K|2h2|1-{\bm{\nu}}\cdot{\bm{\nu}}_{K}|=\frac{1}{2}|{\bm{\nu}}-{\bm{\nu}}_{K}|^{2}\lesssim h^{2}, and |1𝝂Ka𝝂K|h2|1-{\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}|\lesssim h^{2}. Thus employing (2.12) and the identity 𝝂Ka𝒖˘Ka=0{\bm{\nu}}_{K_{a}}\cdot\breve{\bm{u}}_{K_{a}}=0 we have

|𝒖˘KaK𝒖˘Ka|=|(𝝂𝝂K𝝂𝝂Ka[𝐈𝝂𝝂K𝝂𝝂K]𝝂Ka𝝂K[𝐈𝝂Ka𝝂K𝝂Ka𝝂K])𝒖˘Ka||[[𝐈𝝂𝝂K][𝐈𝝂Ka𝝂K]]𝒖˘Ka|+h2|𝒖˘Ka|=|(𝝂𝝂Ka)𝝂K𝒖˘Ka|+h2|𝒖˘Ka|=|(𝝂𝝂Ka)(𝝂K𝝂Ka)𝒖˘Ka|+h2|𝒖˘Ka|h2|𝒖˘Ka|.\displaystyle\begin{aligned} |\breve{\bm{u}}_{K}-\mathcal{M}_{a}^{K}\breve{\bm{u}}_{K_{a}}|&=\left|\left(\frac{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K_{a}}}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}\cdot{\bm{\nu}}_{K}}\right]-{\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}\left[{\bf I}-\frac{{\bm{\nu}}_{K_{a}}\otimes{\bm{\nu}}_{K}}{{\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}}\right]\right)\breve{\bm{u}}_{K_{a}}\right|\\ &\lesssim\left|\big{[}\left[{\bf I}-{\bm{\nu}}\otimes{\bm{\nu}}_{K}\right]-\left[{\bf I}-{\bm{\nu}}_{K_{a}}\otimes{\bm{\nu}}_{K}\right]\big{]}\breve{\bm{u}}_{K_{a}}\right|+h^{2}|\breve{\bm{u}}_{K_{a}}|\\ &=|({\bm{\nu}}-{\bm{\nu}}_{K_{a}})\otimes{\bm{\nu}}_{K}\breve{\bm{u}}_{K_{a}}|+h^{2}|\breve{\bm{u}}_{K_{a}}|\\ &=|({\bm{\nu}}-{\bm{\nu}}_{K_{a}})\otimes({\bm{\nu}}_{K}-{\bm{\nu}}_{K_{a}})\breve{\bm{u}}_{K_{a}}|+h^{2}|\breve{\bm{u}}_{K_{a}}|\\ &\lesssim h^{2}|\breve{\bm{u}}_{K_{a}}|.\end{aligned}

Finally noting that |𝒖˘Ka|=|𝒫𝒑|Ka1𝒖||𝒖||\breve{\bm{u}}_{K_{a}}|=|\mathcal{P}_{\bm{p}|_{K_{a}}^{-1}}\bm{u}|\lesssim{|\bm{u}|} (cf. (2.7)) completes the proof. ∎

Remark 2.6.

In SFEMs it is common to use a higher-order surface approximation Γ\Gamma to γ\gamma of polynomial degree kk (here we consider k=1k=1). In that case νK\nu_{K} is no longer constant on KK, and we have |ν(a)νK(a)|hk|\nu(a)-\nu_{K}(a)|\lesssim h^{k}. The results of Lemma 2.5 easily generalize to this situation with h2kh^{2k} replacing h2h^{2} on the right hand side of (2.13).

3. Finite element spaces and inf-sup stability

By utilizing Lemma 2.5, we can construct tangential finite element spaces on the surface approximation Γh\Gamma_{h} using nodal (Lagrange) basis functions. The essential idea is to enforce continuity at nodal degrees of freedom in a weak sense through the mapping aK\mathcal{M}_{a}^{K} given in Definition 2.3. Although this procedure does not yield a globally continuous finite element space, it preserves in-plane normal continuity and exhibits weak continuity properties. These properties are generally sufficient for achieving convergence in second-order elliptic problems.

In the following discussion, we focus on the construction of the lowest-order MINI Stokes pair for simplicity [1]. However, we expect that Definition 2.3 and Lemma 2.5 provide a general framework for constructing convergent finite element schemes based on classical and conforming finite element pairs such as Taylor-Hood and Scott-Vogelius [2].

3.1. Surface MINI space and approximation properties

Let K^\hat{K} be the reference triangle with vertices (0,0),(1,0),(0,1)(0,0),(1,0),(0,1), and for K𝒯hK\in\mathcal{T}_{h}, let FK:K^KF_{K}:\hat{K}\to K be an affine diffeomorphism. The constant Jacobian matrix of FKF_{K} is denoted by DFK3×2DF_{K}\in\mathbb{R}^{3\times 2}. Note that the columns of DFKDF_{K} span the tangential space of KK. For a vector-valued function 𝒗^:K^2\hat{\bm{v}}:\hat{K}\to\mathbb{R}^{2}, its Piola transform with respect to FKF_{K} is given by

(3.1) 𝒗(x)=(𝒫FK𝒗^)(x):=1JDFK𝒗^(x^),x=FK(x^),\displaystyle\bm{v}(x)=(\mathcal{P}_{F_{K}}\hat{\bm{v}})(x):=\frac{1}{J}DF_{K}\hat{\bm{v}}(\hat{x}),\qquad x=F_{K}(\hat{x}),

and J=det(DFKDFK)J=\sqrt{\det(DF_{K}^{\intercal}DF_{K})}.

Let bKb_{K} be the standard cubic bubble function on KK, i.e., the product of the three barycentric coordinates of KK. The local MINI space defined on the reference triangle is given by 𝑽^:=\EuScript𝑷1(K^)bK^\EuScript𝑷0(K^)\hat{\bm{V}}:=\bm{\EuScript{P}}_{1}(\hat{K})\oplus b_{\hat{K}}\bm{\EuScript{P}}_{0}(\hat{K}), where \EuScriptPk(D)\EuScript{P}_{k}(D) is the space of polynomials of degree k\leq k with domain DD, and \EuScript𝑷k(D)=[\EuScriptPk(D)]2\bm{\EuScript{P}}_{k}(D)=[\EuScript{P}_{k}(D)]^{2}. We then define the surface finite element spaces on Γh\Gamma_{h} as

𝑽h\displaystyle\bm{V}_{h} ={𝒗𝑳2(Γh):K𝒯h𝒗^𝑽^,𝒗K=𝒫FK𝒗^;𝒗K(a)=aK(𝒗Ka(a))K𝒯a,a𝒱h},\displaystyle=\{\bm{v}\in\bm{L}_{2}(\Gamma_{h}):\forall K\in\mathcal{T}_{h}\,\exists\hat{\bm{v}}\in\hat{\bm{V}},\ \bm{v}_{K}=\mathcal{P}_{F_{K}}\hat{\bm{v}}\ ;\ \bm{v}_{K}(a)=\mathcal{M}_{a}^{K}(\bm{v}_{K_{a}}(a))\ \forall K\in\mathcal{T}_{a},\ \forall a\in\mathcal{V}_{h}\},
Qh\displaystyle Q_{h} ={qH1(Γh)L̊2(Γh):qK\EuScriptP1(K)K𝒯h},\displaystyle=\{q\in H^{1}(\Gamma_{h})\cap\mathring{L}_{2}(\Gamma_{h}):\ q_{K}\in\EuScript{P}_{1}(K)\ \forall K\in\mathcal{T}_{h}\},

where 𝒗K=𝒗|K\bm{v}_{K}=\bm{v}|_{K} is the restriction of 𝒗\bm{v} to K𝒯hK\in\mathcal{T}_{h}, aK\mathcal{M}_{a}^{K} is defined in Definition 2.3, and we recall 𝒱h\mathcal{V}_{h} is the set of vertices in 𝒯h\mathcal{T}_{h}.

For 𝒗𝑽h\bm{v}\in\bm{V}_{h}, we let 𝒗L\bm{v}_{L} denote the linear portion of 𝒗\bm{v}, i.e., 𝒗L\bm{v}_{L} is the unique tangential and piecewise linear vector in 𝑽h\bm{V}_{h} satisfying (𝒗L)Ka(a)=𝒗Ka(a)(\bm{v}_{L})_{K_{a}}(a)=\bm{v}_{K_{a}}(a) for all a𝒱ha\in\mathcal{V}_{h}. We then have the following identity on each K𝒯hK\in\mathcal{T}_{h}:

(3.2) 𝒗=𝒗L+60bKK(𝒗𝒗L)𝒗𝑽h,\bm{v}=\bm{v}_{L}+60b_{K}\fint_{K}(\bm{v}-\bm{v}_{L})\qquad\forall\bm{v}\in\bm{V}_{h},

where we have used the fact KbK=|K|/60\int_{K}b_{K}=|K|/60 for all K𝒯hK\in\mathcal{T}_{h}.

Because the columns of DFKDF_{K} span the tangential space of KK, we see that functions in the discrete velocity space 𝑽h\bm{V}_{h} are tangential, i.e., 𝒗𝝂h=0\bm{v}\cdot{\bm{\nu}}_{h}=0 for all 𝒗𝑽h\bm{v}\in\bm{V}_{h}. In addition, due to the normal-preserving properties of the transform aK\mathcal{M}_{a}^{K}, the space is H(div)H({\rm div})-conforming as the next result shows.

Proposition 3.1.

There holds 𝐕h𝐇(divΓh;Γh)\bm{V}_{h}\subset\bm{H}({\rm div}_{\Gamma_{h}};\Gamma_{h}).

Proof.

Due to the properties of the cubic bubble, it is sufficient to show that the linear component of 𝒗𝑽h\bm{v}\in\bm{V}_{h} satisfies the in-plane normal continuity condition (2.3) across all edges in 𝒯h\mathcal{T}_{h}.

Let a𝒱ha\in\mathcal{V}_{h} be a vertex of 𝒯h\mathcal{T}_{h}, and let K1,K2𝒯aK_{1},K_{2}\in\mathcal{T}_{a} be two elements that have aa as a vertex and share a common edge e=K1K2e={\partial}K_{1}\cap{\partial}K_{2}. Denote by 𝒏je\bm{n}^{e}_{j} the in-plane outward unit normal vector with respect to Kj{\partial}K_{j} restricted to ee.

Using the definitions of the finite element space and the operator aK\mathcal{M}_{a}^{K}, along with the Binet-Cauchy identity, there holds for any 𝒗𝑽h\bm{v}\in\bm{V}_{h},

𝒗j(a)𝒏je\displaystyle\bm{v}_{j}(a)\cdot\bm{n}_{j}^{e} =aKj(𝒗Ka(a))𝒏je\displaystyle=\mathcal{M}_{a}^{K_{j}}(\bm{v}_{K_{a}}(a))\cdot\bm{n}^{e}_{j}
=(𝝂Ka𝝂Kj)(𝒗Ka(a)𝒏je)(𝝂Ka𝒏je)(𝝂Kj𝒗Ka(a))=(𝝂Ka×𝒗Ka(a))(𝝂Kj×𝒏je),\displaystyle=({\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K_{j}})(\bm{v}_{K_{a}}(a)\cdot\bm{n}^{e}_{j})-({\bm{\nu}}_{K_{a}}\cdot\bm{n}^{e}_{j})({\bm{\nu}}_{K_{j}}\cdot\bm{v}_{K_{a}}(a))=({\bm{\nu}}_{K_{a}}\times\bm{v}_{K_{a}}(a))\cdot({\bm{\nu}}_{K_{j}}\times\bm{n}^{e}_{j}),

where 𝒗j=𝒗Kj=𝒗|Kj\bm{v}_{j}=\bm{v}_{K_{j}}=\bm{v}|_{K_{j}}. Therefore,

𝒗1(a)𝒏1e+𝒗2(a)𝒏2e\displaystyle\bm{v}_{1}(a)\cdot\bm{n}^{e}_{1}+\bm{v}_{2}(a)\cdot\bm{n}^{e}_{2} =(𝝂Ka×𝒗Ka(a))((𝝂K1×𝒏1e)+(𝝂K2×𝒏2e))=0.\displaystyle=({\bm{\nu}}_{K_{a}}\times\bm{v}_{K_{a}}(a))\cdot\big{(}({\bm{\nu}}_{K_{1}}\times\bm{n}^{e}_{1})+({\bm{\nu}}_{K_{2}}\times\bm{n}^{e}_{2})\big{)}=0.

Because 𝒗j\bm{v}_{j} is a linear polynomial on ee, we conclude that (2.3) is satisfied on all edges. This implies the desired result 𝑽h𝑯(divΓh;Γh)\bm{V}_{h}\subset\bm{H}({\rm div}_{\Gamma_{h}};\Gamma_{h}). ∎

Lemma 3.2.

For each 𝐰𝐂(γ)𝐇T1(γ)𝐇h2(γ)\bm{w}\in\bm{C}(\gamma)\cap\bm{H}^{1}_{T}(\gamma)\cap\bm{H}_{h}^{2}(\gamma), there exists 𝐈h𝐰˘𝐕h\bm{I}_{h}\breve{\bm{w}}\in\bm{V}_{h} such that

hm𝒘˘𝑰h𝒘˘Hm(K)h2𝒘Hh2(ωKγ)K𝒯h,m=0,1,\displaystyle h^{m}\|\breve{\bm{w}}-\bm{I}_{h}\breve{\bm{w}}\|_{H^{m}(K)}\lesssim h^{2}\|\bm{w}\|_{H^{2}_{h}(\omega_{K^{\gamma}})}\qquad\forall K\in\mathcal{T}_{h},\quad m=0,1,

with 𝐰˘=𝒫𝐩1𝐰\breve{\bm{w}}=\mathcal{P}_{\bm{p}^{-1}}\bm{w}.

Proof.

Given 𝒘𝑪(γ)𝑯T1(γ)𝑯h2(γ)\bm{w}\in\bm{C}(\gamma)\cap\bm{H}^{1}_{T}(\gamma)\cap\bm{H}_{h}^{2}(\gamma), we uniquely define 𝒗:=𝑰h𝒘˘𝑽h\bm{v}:=\bm{I}_{h}\breve{\bm{w}}\in\bm{V}_{h} such that each component of 𝒗\bm{v} is a piecewise linear polynomial and satisfies

(3.3) 𝒗Ka(a)\displaystyle\bm{v}_{K_{a}}(a) =𝒘˘Ka(a)\displaystyle=\breve{\bm{w}}_{K_{a}}(a)\qquad a𝒱h.\displaystyle\forall a\in\mathcal{V}_{h}.

Let 𝒘˘I\breve{\bm{w}}_{I} be the elementwise (discontinuous), linear interpolant of 𝒘˘\breve{\bm{w}} with respect to the vertices in 𝒯h\mathcal{T}_{h}, i.e., (𝒘˘I)K\EuScript𝑷1(K)(\breve{\bm{w}}_{I})_{K}\in\bm{\EuScript{P}}_{1}(K) with (𝒘˘I)K(a)=𝒘˘K(a)(\breve{\bm{w}}_{I})_{K}(a)=\breve{\bm{w}}_{K}(a) for all K𝒯hK\in\mathcal{T}_{h} and a𝒱Ka\in\mathcal{V}_{K}. By Lemma 2.5 (with 𝒖=𝒘(a)\bm{u}=\bm{w}(a)), (3.3), and the definition of 𝑽h\bm{V}_{h} we have for each vertex a𝒱ha\in\mathcal{V}_{h},

|(𝒘˘I𝒗)K(a)|=|(𝒘˘KaK𝒗Ka)(a)|=|(𝒘˘KaK𝒘˘Ka)(a)|h2|𝒘˘Ka(a)|K𝒯a.\displaystyle|(\breve{\bm{w}}_{I}-\bm{v})_{K}(a)|=|(\breve{\bm{w}}_{K}-\mathcal{M}_{a}^{K}\bm{v}_{K_{a}})(a)|=|(\breve{\bm{w}}_{K}-\mathcal{M}_{a}^{K}\breve{\bm{w}}_{K_{a}})(a)|\lesssim h^{2}|\breve{\bm{w}}_{K_{a}}(a)|\qquad\forall K\in\mathcal{T}_{a}.

Consequently, by standard inverse estimates,

hm𝒘˘I𝒗Hm(K)\displaystyle h^{m}\|\breve{\bm{w}}_{I}-\bm{v}\|_{H^{m}(K)} 𝒘˘I𝒗L2(K)\displaystyle\lesssim\|\breve{\bm{w}}_{I}-\bm{v}\|_{L_{2}(K)}
h𝒘˘I𝒗L(K)=hmaxa𝒱K|(𝒘˘I𝒗)K(a)|h3maxa𝒱K|𝒘˘Ka(a)|m=0,1.\displaystyle\lesssim h\|\breve{\bm{w}}_{I}-\bm{v}\|_{L_{\infty}(K)}=h\max_{a\in\mathcal{V}_{K}}|(\breve{\bm{w}}_{I}-\bm{v})_{K}(a)|\lesssim h^{3}\max_{a\in\mathcal{V}_{K}}|\breve{\bm{w}}_{K_{a}}(a)|\quad m=0,1.

Using inverse estimates once again, and applying standard interpolation results yields

maxa𝒱K|𝒘˘Ka(a)|\displaystyle\max_{a\in\mathcal{V}_{K}}|\breve{\bm{w}}_{K_{a}}(a)| 𝒘˘IL(ωK)\displaystyle\leq\|\breve{\bm{w}}_{I}\|_{L_{\infty}(\omega_{K})}
h1𝒘˘IL2(ωK)\displaystyle\lesssim h^{-1}\|\breve{\bm{w}}_{I}\|_{L_{2}(\omega_{K})}
h1(𝒘˘L2(ωK)+𝒘˘𝒘˘IL2(ωK))\displaystyle\lesssim h^{-1}(\|\breve{\bm{w}}\|_{L_{2}(\omega_{K})}+\|\breve{\bm{w}}-\breve{\bm{w}}_{I}\|_{L_{2}(\omega_{K})})
h1(𝒘˘L2(ωK)+h2|𝒘˘|Hh2(ωK)).\displaystyle\lesssim h^{-1}(\|\breve{\bm{w}}\|_{L_{2}(\omega_{K})}+h^{2}\ |\breve{\bm{w}}|_{H^{2}_{h}(\omega_{K})}).

Therefore,

hm𝒘˘I𝒗Hm(K)h2𝒘˘Hh2(ωK)m=0,1,\displaystyle h^{m}\|\breve{\bm{w}}_{I}-\bm{v}\|_{H^{m}(K)}\lesssim h^{2}\|\breve{\bm{w}}\|_{H^{2}_{h}(\omega_{K})}\qquad m=0,1,

and so by (2.9),

hm𝒘˘𝒗Hm(K)\displaystyle h^{m}\|\breve{\bm{w}}-{\bm{v}}\|_{H^{m}(K)} hm𝒘˘𝒘˘IHm(K)+hm𝒘˘I𝒗Hm(K)h2𝒘Hh2(ωKγ).\displaystyle\lesssim h^{m}\|\breve{\bm{w}}-\breve{\bm{w}}_{I}\|_{H^{m}(K)}+h^{m}\|\breve{\bm{w}}_{I}-\bm{v}\|_{H^{m}(K)}\lesssim h^{2}\|\bm{w}\|_{H^{2}_{h}(\omega_{K^{\gamma}})}.

Lemma 3.3.

There exists a constant α>0\alpha>0 independent of hh such that

(3.4) sup𝒗𝑽h\{0}Γh(divΓh𝒗)q𝒗Hh1(Γh)αqL2(Γh)qQh.\sup_{\bm{v}\in\bm{V}_{h}\backslash\{0\}}\frac{\int_{\Gamma_{h}}({\rm div}_{\Gamma_{h}}\bm{v})q}{\|\bm{v}\|_{H^{1}_{h}(\Gamma_{h})}}\geq\alpha\|q\|_{L_{2}(\Gamma_{h})}\qquad\forall q\in Q_{h}.
Proof.

Fix qQhq\in Q_{h}, and let 𝒘𝑯T1(γ)\bm{w}\in\bm{H}^{1}_{T}(\gamma) satisfy [20, 14, 11]

(3.5) divγ𝒘=(μh1q)L̊2(γ),and𝒘H1(γ)(μh1q)L2(γ)qL2(Γh),{\rm div}_{\gamma}\bm{w}=(\mu_{h}^{-1}q)^{\ell}\in\mathring{L}_{2}(\gamma),\quad\text{and}\quad\|\bm{w}\|_{H^{1}(\gamma)}\lesssim\|(\mu_{h}^{-1}q)^{\ell}\|_{L_{2}(\gamma)}\lesssim\|q\|_{L_{2}(\Gamma_{h})},

where we used (2.4) and (2.2) in the last step. Let 𝑰hSZ𝒘e\bm{I}^{SZ}_{h}\bm{w}^{e} be the Scott-Zhang interpolant of the extension 𝒘e𝑯1(Γh)\bm{w}^{e}\in\bm{H}^{1}(\Gamma_{h}) onto the space of continuous piecewise linear polynomials with respect to 𝒯h\mathcal{T}_{h} [7, 12], and set 𝒘h=𝚷(𝑰hSZ𝒘e)𝑯T1(γ)\bm{w}_{h}=\bm{\Pi}(\bm{I}^{SZ}_{h}\bm{w}^{e})^{\ell}\in\bm{H}^{1}_{T}(\gamma). From (2.4), 𝒘𝒘h=Π(𝒘𝑰hSZ𝒘e)\bm{w}-\bm{w}_{h}=\Pi(\bm{w}-\bm{I}^{SZ}_{h}\bm{w}^{e}), and approximation properties of the Scott-Zhang interpolant, there holds on each K𝒯hK\in\mathcal{T}_{h},

(3.6) 𝒘hH1(Kγ)+h1𝒘𝒘hL2(Kγ)(𝑰hSZ𝒘e)H1(Kγ)+h1𝒘(𝑰hSZ𝒘e)L2(Kγ)𝑰hSZ𝒘eH1(K)+h1𝒘e𝑰hSZ𝒘eL2(K)𝒘eH1(ωK)𝒘H1(ωKγ).\begin{split}\|\bm{w}_{h}\|_{H^{1}(K^{\gamma})}+h^{-1}\|\bm{w}-\bm{w}_{h}\|_{L_{2}(K^{\gamma})}&\lesssim\|(\bm{I}^{SZ}_{h}\bm{w}^{e})^{\ell}\|_{H^{1}(K^{\gamma})}+h^{-1}\|\bm{w}-(\bm{I}^{SZ}_{h}\bm{w}^{e})^{\ell}\|_{L_{2}(K^{\gamma})}\\ &\lesssim\|\bm{I}^{SZ}_{h}\bm{w}^{e}\|_{H^{1}(K)}+h^{-1}\|\bm{w}^{e}-\bm{I}^{SZ}_{h}\bm{w}^{e}\|_{L_{2}(K)}\\ &\lesssim\|\bm{w}^{e}\|_{H^{1}(\omega_{K})}\lesssim\|\bm{w}\|_{H^{1}(\omega_{K^{\gamma}})}.\end{split}

Noting 𝒘h𝑪(γ)𝑯T1(γ)𝑯h2(γ)\bm{w}_{h}\in\bm{C}(\gamma)\cap\bm{H}^{1}_{T}(\gamma)\cap\bm{H}_{h}^{2}(\gamma), we define 𝒗𝑽h\bm{v}\in\bm{V}_{h} such that

𝒗K=(𝑰h𝒘˘h)K+60bKK(𝒘˘𝑰h𝒘˘h)K𝒯h,\bm{v}_{K}=\big{(}{\bm{I}_{h}\breve{\bm{w}}_{h}}\big{)}_{K}+60b_{K}\fint_{K}(\breve{\bm{w}}-{\bm{I}_{h}\breve{\bm{w}}_{h}})\qquad\forall K\in\mathcal{T}_{h},

where 𝑰h𝒘˘h\bm{I}_{h}\breve{\bm{w}}_{h} is given in Lemma 3.2. We then have K𝒗=K𝒘˘\int_{K}\bm{v}=\int_{K}\breve{\bm{w}}, and by (3.6) and Lemma 3.2,

𝒗H1(K)\displaystyle\|\bm{v}\|_{H^{1}(K)} 𝑰h𝒘˘hH1(K)+𝒗𝑰h𝒘˘hH1(K)\displaystyle\leq\|\bm{I}_{h}\breve{\bm{w}}_{h}\|_{H^{1}(K)}+\|\bm{v}-{\bm{I}_{h}\breve{\bm{w}}_{h}}\|_{H^{1}(K)}
𝒘˘H1(K)+𝒘˘𝑰h𝒘˘hH1(K)+h1𝒘˘𝑰h𝒘˘hL2(K)\displaystyle\lesssim\|\breve{\bm{w}}\|_{H^{1}(K)}+\|\breve{\bm{w}}-\bm{I}_{h}\breve{\bm{w}}_{h}\|_{H^{1}(K)}+h^{-1}\|\breve{\bm{w}}-{\bm{I}_{h}\breve{\bm{w}}_{h}}\|_{L_{2}(K)}
𝒘˘H1(K)+𝒘˘𝒘˘hH1(K)+𝒘˘h𝑰h𝒘˘hH1(K)\displaystyle\leq\|\breve{\bm{w}}\|_{H^{1}(K)}+\|\breve{\bm{w}}-\breve{\bm{w}}_{h}\|_{H^{1}(K)}+\|\breve{\bm{w}}_{h}-{\bm{I}_{h}\breve{\bm{w}}_{h}}\|_{H^{1}(K)}
+h1(𝒘˘𝒘˘hL2(K)+𝒘˘h𝑰h𝒘˘hL2(K))\displaystyle~{}~{}~{}~{}~{}~{}~{}+h^{-1}(\|\breve{\bm{w}}-\breve{\bm{w}}_{h}\|_{L_{2}(K)}+\|\breve{\bm{w}}_{h}-{\bm{I}_{h}\breve{\bm{w}}_{h}}\|_{L_{2}(K)})
𝒘H1(ωKγ)+h𝒘hHh2(ωKγ).\displaystyle\lesssim\|\bm{w}\|_{H^{1}(\omega_{K^{\gamma}})}+h\|\bm{w}_{h}\|_{H^{2}_{h}(\omega_{K^{\gamma}})}.

By (2.4), a standard inverse estimate, and the H1H^{1}-stability properties of the Scott-Zhang interpolant,

h𝒘hHh2(ωKγ)h(𝑰hSZ𝒘e)Hh2(ωKγ)h𝑰hSZ𝒘eHh2(ωK)𝑰hSZ𝒘eH1(ωK)𝒘H1(ωKγ),h\|\bm{w}_{h}\|_{H^{2}_{h}(\omega_{K^{\gamma}})}\lesssim h\|(\bm{I}^{SZ}_{h}\bm{w}^{e})^{\ell}\|_{H^{2}_{h}(\omega_{K^{\gamma}})}\lesssim h\|\bm{I}^{SZ}_{h}\bm{w}^{e}\|_{H^{2}_{h}(\omega_{K})}\lesssim\|\bm{I}^{SZ}_{h}\bm{w}^{e}\|_{H^{1}(\omega_{K})}\lesssim\|\bm{w}\|_{H^{1}(\omega^{\prime}_{K^{\gamma}})},

and so by (3.5),

(3.7) 𝒗H1(K)𝒘H1(ωKγ)K𝒯h𝒗Hh1(Γh)𝒘H1(γ)qL2(Γh).\displaystyle\|\bm{v}\|_{H^{1}(K)}\lesssim\|\bm{w}\|_{H^{1}(\omega^{\prime}_{K_{\gamma}})}\ \forall K\in\mathcal{T}_{h}\quad\Longrightarrow\quad\|\bm{v}\|_{H^{1}_{h}(\Gamma_{h})}\lesssim\|\bm{w}\|_{H^{1}(\gamma)}\lesssim\|q\|_{L_{2}(\Gamma_{h})}.

Next, we recall from Proposition 3.1 that 𝒗𝑯(divΓh;Γh)\bm{v}\in\bm{H}({\rm div}_{\Gamma_{h}};\Gamma_{h}) and therefore (2.3) is satisfied. Thus, by integration by parts, the identity K𝒗=K𝒘˘\int_{K}\bm{v}=\int_{K}\breve{\bm{w}}, and applying (2.8) yields

Γh(divΓh𝒗)q\displaystyle\int_{\Gamma_{h}}({\rm div}_{\Gamma_{h}}\bm{v})q =Γh𝒗Γhq=Γh𝒘˘Γhq=Γh(divΓh𝒘˘)q=γ(divγ𝒘)q.\displaystyle=-\int_{\Gamma_{h}}\bm{v}\cdot\nabla_{\Gamma_{h}}q=-\int_{\Gamma_{h}}\breve{\bm{w}}\cdot\nabla_{\Gamma_{h}}q=\int_{\Gamma_{h}}({\rm div}_{\Gamma_{h}}\breve{\bm{w}})q=\int_{\gamma}({\rm div}_{\gamma}\bm{w})q^{\ell}.

We then use (3.5), (2.2), and (2.4) to obtain

Γh(divΓh𝒗)qqL2(Γh)2.\displaystyle\int_{\Gamma_{h}}({\rm div}_{\Gamma_{h}}\bm{v})q\gtrsim\|q\|_{L_{2}(\Gamma_{h})}^{2}.

This identity combined with (3.7) completes the proof. ∎

3.2. HT1\textit{{H}}^{\hskip 1.0pt1}_{T}-conforming approximations to discrete functions

While the finite element space 𝑽h\bm{V}_{h} is merely 𝑯(divΓh;Γh)\bm{H}({\rm div}_{\Gamma_{h}};\Gamma_{h})-conforming (cf. Proposition 3.1), the following lemma shows that functions in this space are “close” to an 𝑯1\bm{H}^{1}-conforming relative.

Lemma 3.4.

Given 𝐯𝐕h\bm{v}\in\bm{V}_{h}, denote by \wideparen𝐯=𝒫𝐩𝐯\wideparen{\bm{v}}=\mathcal{P}_{\bm{p}}\bm{v} its Piola transform via the closest point projection to γ\gamma. Then there exists \wideparen𝐯c𝐇T1(γ)\wideparen\bm{v}_{c}\in\bm{H}_{T}^{1}(\gamma) such that

(3.8) \wideparen𝒗\wideparen𝒗cL2(Kγ)+h|\wideparen𝒗\wideparen𝒗c|Hh1(Kγ)h2\wideparen𝒗L2(Kγ)Kγ𝒯hγ.\displaystyle\|\wideparen{\bm{v}}-\wideparen{\bm{v}}_{c}\|_{L_{2}(K^{\gamma})}+h|\wideparen{\bm{v}}-\wideparen\bm{v}_{c}|_{H_{h}^{1}(K^{\gamma})}\lesssim h^{2}\|\wideparen{\bm{v}}\|_{L_{2}(K^{\gamma})}\quad\forall K^{\gamma}\in\mathcal{T}_{h}^{\gamma}.
Proof.

On an element K𝒯hK\in\mathcal{T}_{h}, we first write 𝒗=𝒗L+𝜶bK\bm{v}=\bm{v}_{L}+{\bm{\alpha}}b_{K}, where 𝒗L\bm{v}_{L} is componentwise affine on KK, and 𝜶3{{\bm{\alpha}}}\in\mathbb{R}^{3} is tangent to KK (cf. (3.2)). Likewise, \wideparen𝒗=\wideparen𝒗L+\wideparen𝜶bK\wideparen\bm{v}=\wideparen\bm{v}_{L}+\wideparen{{\bm{\alpha}}}b_{K}^{\ell} is the Piola transform of 𝒗\bm{v} to γ\gamma. We next let 𝒘\bm{w} be the unique continuous piecewise linear polynomial with respect to 𝒯h\mathcal{T}_{h} satisfying 𝒘(a)=𝒗Ka(a)\bm{w}(a)=\bm{v}_{K_{a}}(a) for each vertex a𝒱ha\in\mathcal{V}_{h}. We then set

\wideparen𝒗c=𝚷d𝑯(1dκ1)(1dκ2)𝒘+\wideparen𝜶bK=μh1𝝂𝝂h(𝚷d𝑯)𝒘+\wideparen𝜶bK.\displaystyle\wideparen\bm{v}_{c}=\frac{\bm{\Pi}-d\bm{H}}{(1-d\kappa_{1})(1-d\kappa_{2})}\bm{w}^{\ell}+\wideparen{{\bm{\alpha}}}b_{K}^{\ell}=\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}+\wideparen{{\bm{\alpha}}}b_{K}^{\ell}.

Note that \wideparen𝒗c𝑯T1(γ)\wideparen\bm{v}_{c}\in\bm{H}_{T}^{1}(\gamma), and

\wideparen𝒗\wideparen𝒗c=\wideparen𝒗Lμh1𝝂𝝂h(𝚷d𝑯)𝒘.\displaystyle\wideparen\bm{v}-\wideparen\bm{v}_{c}=\wideparen\bm{v}_{L}-\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}.

Fixing K𝒯hK\in\mathcal{T}_{h}, by norm equivalence (cf. (2.9)) we prove (3.8) by establishing that

(3.25) 𝒗L(μh1𝝂𝝂h(𝚷d𝑯)𝒘L2(K)+h|𝒗L(μh1𝝂𝝂h(𝚷d𝑯)𝒘|Hh1(K)h2𝒗L2(K),\displaystyle\|\bm{v}_{L}-\mathchoice{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\displaystyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\textstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{39.86778pt}{2.72223pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{30.4194pt}{1.94444pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptscriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}\|_{L_{2}(K)}+h|\bm{v}_{L}-\mathchoice{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\displaystyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\textstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{39.86778pt}{2.72223pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{30.4194pt}{1.94444pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptscriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}|_{H_{h}^{1}(K)}\lesssim h^{2}\|\bm{v}\|_{L_{2}(K)},

where

(

μh1𝝂𝝂h(𝚷d𝑯)𝒘
=𝒫𝒑|K1μh1𝝂𝝂h(𝚷d𝑯)𝒘
\mathchoice{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\displaystyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\textstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{39.86778pt}{2.72223pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{30.4194pt}{1.94444pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptscriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}=\mathcal{P}_{\bm{p}|_{K}^{-1}}\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}
. We show (3.25) in three steps.

(i) Employing (2.7), (2.1), and 𝚷h(𝝂𝝂h𝝂𝝂h)𝚷=𝝂𝝂h𝝂𝝂h\bm{\Pi}_{h}({\bm{\nu}}\cdot{\bm{\nu}}_{h}-{\bm{\nu}}\otimes{\bm{\nu}}_{h})\bm{\Pi}={\bm{\nu}}\cdot{\bm{\nu}}_{h}-{\bm{\nu}}\otimes{\bm{\nu}}_{h}, we have that

(

μh1𝝂𝝂h(𝚷d𝑯)𝒘
=𝚷K((𝝂𝝂K)𝐈𝝂𝝂K)𝒘
.
\mathchoice{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\displaystyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\textstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{39.86778pt}{2.72223pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{30.4194pt}{1.94444pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptscriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}=\bm{\Pi}_{K}\left(({\bm{\nu}}\cdot{\bm{\nu}}_{K}){{\bf I}}-{\bm{\nu}}\otimes{\bm{\nu}}_{K}\right)\bm{w}.

Here we interchangeably write 𝝂h=𝝂K=𝝂h|K{\bm{\nu}}_{h}={\bm{\nu}}_{K}={\bm{\nu}}_{h}|_{K} and 𝚷h=𝚷K\bm{\Pi}_{h}=\bm{\Pi}_{K} in order to better distinguish dependence on the element KK. Using 𝒗L𝝂K=0\bm{v}_{L}\cdot{\bm{\nu}}_{K}=0 and 𝚷K𝒗L=𝒗L\bm{\Pi}_{K}\bm{v}_{L}=\bm{v}_{L}, we then have

(3.26) 𝒗L(μh1𝝂𝝂h(𝚷d𝑯)𝒘=𝚷K[𝒗L((𝝂𝝂K)𝐈𝝂𝝂K)𝒘]=[(1𝝂𝝂K)𝒗L]+[𝝂𝝂K𝚷K(𝒗L𝒘)]+[𝝂K(𝒘𝒗L)𝚷K(𝝂𝝂K)]=:I+II+III.\displaystyle\begin{aligned} \bm{v}_{L}-&\mathchoice{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\displaystyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\textstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{39.86778pt}{2.72223pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{30.4194pt}{1.94444pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptscriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}=\bm{\Pi}_{K}[\bm{v}_{L}-\left(({\bm{\nu}}\cdot{\bm{\nu}}_{K}){{\bf I}}-{\bm{\nu}}\otimes{\bm{\nu}}_{K}\right)\bm{w}]\\ &=[(1-{\bm{\nu}}\cdot{\bm{\nu}}_{K})\bm{v}_{L}]+[{\bm{\nu}}\cdot{\bm{\nu}}_{K}\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})]+[{\bm{\nu}}_{K}\cdot(\bm{w}-\bm{v}_{L})\bm{\Pi}_{K}({\bm{\nu}}-{\bm{\nu}}_{K})]\\ &=:I+II+III.\end{aligned}

(ii) We next bound the terms II, IIII, and IIIIII in L2L_{2}. Using |1𝝂𝝂K|=12|𝝂𝝂K|2h2|1-{\bm{\nu}}\cdot{\bm{\nu}}_{K}|=\frac{1}{2}|{\bm{\nu}}-{\bm{\nu}}_{K}|^{2}\lesssim h^{2} yields

(3.27) IL2(K)h2𝒗LL2(K).\displaystyle\|I\|_{L_{2}(K)}\lesssim h^{2}\|\bm{v}_{L}\|_{L_{2}(K)}.

Next we use (2.12) and recall that 𝒗Ka(a)𝝂Ka=0\bm{v}_{K_{a}}(a)\cdot{\bm{\nu}}_{K_{a}}=0 to compute that for each vertex aKa\in K

(3.28) |𝚷K(𝒗L𝒘)(a)|=|𝚷K(aK𝐈)𝒗Ka(a)|=|𝚷K[((𝝂Ka𝝂K)𝐈𝝂Ka𝝂K)𝐈]𝒗Ka(a)|=|(𝝂Ka𝝂K1)𝚷K𝒗Ka(a)(𝝂K𝝂Ka)𝒗Ka(a)𝚷K(𝝂Ka𝝂K)|h2|𝒗Ka(a)|h2|𝒗K(a)|.\displaystyle\begin{aligned} |\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})(a)|&=|\bm{\Pi}_{K}(\mathcal{M}_{a}^{K}-{\bf I})\bm{v}_{K_{a}}(a)|\\ &=|\bm{\Pi}_{K}[\left(({\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}){{\bf I}}-{\bm{\nu}}_{K_{a}}\otimes{\bm{\nu}}_{K}\right)-{\bf I}]\bm{v}_{K_{a}}(a)|\\ &=|({\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}-1)\bm{\Pi}_{K}\bm{v}_{K_{a}}(a)-({\bm{\nu}}_{K}-{\bm{\nu}}_{K_{a}})\cdot\bm{v}_{K_{a}}(a)\bm{\Pi}_{K}({\bm{\nu}}_{K_{a}}-{\bm{\nu}}_{K})|\\ &\lesssim h^{2}|\bm{v}_{K_{a}}(a)|\lesssim{h^{2}}|\bm{v}_{K}(a)|.\end{aligned}

In the last step we have employed (2.6) and (2.7) to obtain |𝒗Ka(a)|=|1𝝂K𝝂Ka𝚷Ka𝒗K(a)||𝒗K(a)||\bm{v}_{K_{a}}(a)|=|\frac{1}{{\bm{\nu}}_{K}\cdot{\bm{\nu}}_{K_{a}}}\bm{\Pi}_{K_{a}}\bm{v}_{K}(a)|\lesssim|\bm{v}_{K}(a)|. We then use the fact that 𝚷K(𝒗L𝒘)\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w}) and 𝒗L\bm{v}_{L} are affine, along with inverse inequalities, to obtain

(3.29) IIL2(K)𝚷K(𝒗L𝒘)L2(K)hmaxa𝒱K|𝚷K(𝒗L𝒘)(a)|h3𝒗LL(K)h2𝒗LL2(K).\displaystyle\begin{aligned} \|II\|_{L_{2}(K)}&\lesssim\|\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})\|_{L_{2}(K)}\lesssim h\max_{a\in{\mathcal{V}_{K}}}|\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})(a)|\\ &\lesssim h^{3}\|\bm{v}_{L}\|_{L_{\infty}(K)}\lesssim h^{2}\|\bm{v}_{L}\|_{L_{2}(K)}.\end{aligned}

In order to bound IIIIII, we first proceed similarly as (3.28) to obtain

|(𝒗L𝒘)(a)|=|(𝝂Ka𝝂K1)𝒗Ka(a)(𝝂K𝝂Ka)𝒗Ka(a)𝝂Ka|h|𝒗K(a)|.\displaystyle|(\bm{v}_{L}-\bm{w})(a)|=|({\bm{\nu}}_{K_{a}}\cdot{\bm{\nu}}_{K}-1)\bm{v}_{K_{a}}(a)-({\bm{\nu}}_{K}-{\bm{\nu}}_{K_{a}})\cdot\bm{v}_{K_{a}}(a){\bm{\nu}}_{K_{a}}|\lesssim h|\bm{v}_{K}(a)|.

Using |𝚷K(𝝂𝝂K)|h|\bm{\Pi}_{K}({\bm{\nu}}-{\bm{\nu}}_{K})|\lesssim h, we thus have similar to above that

(3.30) IIIL2(K)h𝒗L𝒘L2(K)h2𝒗LL2(K).\displaystyle\|III\|_{L_{2}(K)}\lesssim h\|\bm{v}_{L}-\bm{w}\|_{L_{2}(K)}\lesssim h^{2}\|\bm{v}_{L}\|_{L_{2}(K)}.

Recalling that 𝒗=𝒗L+𝜶bK\bm{v}=\bm{v}_{L}+{\bm{\alpha}}b_{K} and bK(a)=0b_{K}(a)=0 at vertices aa, we again use inverse inequalities to obtain

(3.31) 𝒗LL2(K)h𝒗LL(K)=hmaxa𝒱K|𝒗L(a)|h𝒗L(K)𝒗L2(K).\displaystyle\|\bm{v}_{L}\|_{L_{2}(K)}\lesssim h\|\bm{v}_{L}\|_{L_{\infty}(K)}=h\max_{a\in\mathcal{V}_{K}}|\bm{v}_{L}(a)|\leq h\|\bm{v}\|_{L_{\infty}(K)}{\lesssim}\|\bm{v}\|_{L_{2}(K)}.

Collecting the inequalities (3.27), (3.29)–(3.31) yields \wideparen𝒗\wideparen𝒗cL2(Kγ)h2\wideparen𝒗L2(Kγ).\|\wideparen\bm{v}-\wideparen{\bm{v}}^{c}\|_{L_{2}(K^{\gamma})}\lesssim h^{2}\|\wideparen\bm{v}\|_{L_{2}(K^{\gamma})}.

(iii) In order to bound Γh(𝒗L

(

μh1𝝂𝝂h(𝚷d𝑯)𝒘
)
L2(K)
\|{\nabla_{\Gamma_{h}}}(\bm{v}_{L}-\mathchoice{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\displaystyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{56.0487pt}{3.8889pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\textstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{39.86778pt}{2.72223pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits}{\mathop{\vbox{\halign{#\cr\kern 0.80002pt$\hss\leavevmode\resizebox{30.4194pt}{1.94444pt}{\rotatebox[origin={c}]{90.0}{(}}\hss$\crcr\nointerlineskip\cr$\hss\scriptscriptstyle\mu_{h}^{-1}{\bm{\nu}}\cdot{\bm{\nu}}_{h}(\bm{\Pi}-d\bm{H})\bm{w}^{\ell}\hss$\crcr}}}\limits})\|_{L_{2}(K)}
, we recall (3.26) and consider first ΓhI\nabla_{\Gamma_{h}}I. First note that |(1𝝂𝝂K)|=|𝑯𝝂K|=|𝑯(𝝂K𝝂)|h|\nabla(1-{\bm{\nu}}\cdot{\bm{\nu}}_{K})|=|\bm{H}{\bm{\nu}}_{K}|=|\bm{H}({\bm{\nu}}_{K}-{\bm{\nu}})|\lesssim h. Thus using an inverse inequality and |1𝝂𝝂K|h2|1-{\bm{\nu}}\cdot{\bm{\nu}}_{K}|\lesssim h^{2}, we obtain

ΓhIL2(K)1𝝂𝝂KL(K)Γh𝒗LL2(K)+(1𝝂𝝂K)L(K)𝒗LL2(K)h𝒗LL2(K).\displaystyle\|\nabla_{\Gamma_{h}}I\|_{L_{2}(K)}\lesssim\|1-{\bm{\nu}}\cdot{\bm{\nu}}_{K}\|_{L_{\infty}(K)}\|\nabla_{\Gamma_{h}}\bm{v}_{L}\|_{L_{2}(K)}+\|\nabla(1-{\bm{\nu}}\cdot{\bm{\nu}}_{K})\|_{L_{\infty}(K)}\|\bm{v}_{L}\|_{L_{2}(K)}\lesssim h\|\bm{v}_{L}\|_{L_{2}(K)}.

Employing inverse inequalities, |(𝝂𝝂K)|h|\nabla({\bm{\nu}}\cdot{\bm{\nu}}_{K})|\lesssim h, and (3.29) also yields

ΓhIIL2(K)(𝝂𝝂K)L(K)𝚷K(𝒗L𝒘)L2(K)+𝝂𝝂KL(K)Γh[𝚷K(𝒗L𝒘)]L2(K)h𝚷K(𝒗L𝒘)L2(K)+h1𝚷K(𝒗L𝒘)L2(K)h𝒗LL2(K).\displaystyle\begin{aligned} \|\nabla_{\Gamma_{h}}II\|_{L_{2}(K)}&\leq\|\nabla({\bm{\nu}}\cdot{\bm{\nu}}_{K})\|_{L_{\infty}(K)}\|\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})\|_{L_{2}(K)}+\|{\bm{\nu}}\cdot{\bm{\nu}}_{K}\|_{L_{\infty}(K)}\|\nabla_{\Gamma_{h}}[\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})]\|_{L_{2}(K)}\\ &\lesssim h\|\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})\|_{L_{2}(K)}+h^{-1}\|\bm{\Pi}_{K}(\bm{v}_{L}-\bm{w})\|_{L_{2}(K)}\\ &\lesssim h\|\bm{v}_{L}\|_{L_{2}(K)}.\end{aligned}

We finally compute using inverse inequalities and (3.30) that

ΓhIIIL2(K)[𝚷K(𝝂𝝂K)]L(K)𝒗L𝒘L2(K)+𝚷K(𝝂𝝂K)L(K)Γh(𝒗L𝒘)L2(K)(1+hh1)𝒗L𝒘L2(K)h𝒗LL2(K).\displaystyle\begin{aligned} \|\nabla_{\Gamma_{h}}III\|_{L_{2}(K)}&\lesssim\|\nabla[\bm{\Pi}_{K}({\bm{\nu}}-{\bm{\nu}}_{K})]\|_{L_{\infty}(K)}\|\bm{v}_{L}-\bm{w}\|_{L_{2}(K)}\\ &~{}~{}~{}~{}~{}~{}~{}+\|\bm{\Pi}_{K}({\bm{\nu}}-{\bm{\nu}}_{K})\|_{L_{\infty}(K)}\|\nabla_{\Gamma_{h}}(\bm{v}_{L}-\bm{w})\|_{L_{2}(K)}\\ &\lesssim(1+hh^{-1})\|\bm{v}_{L}-\bm{w}\|_{L_{2}(K)}\lesssim h\|\bm{v}_{L}\|_{L_{2}(K)}.\end{aligned}

Collecting the above inequalities and employing (3.31) yields

Γh(I+II+III)L2(K)h𝒗LL2(K)h𝒗L2(K),\|\nabla_{\Gamma_{h}}(I+II+III)\|_{L_{2}(K)}\lesssim h\|\bm{v}_{L}\|_{L_{2}(K)}\lesssim h\|\bm{v}\|_{L_{2}(K)},

which completes the proof. ∎

3.3. Discrete Korn-type inequalities

From Lemma 3.4, we immediately obtain a discrete Korn-type inequality on the exact surface γ\gamma.

Lemma 3.5.

Given 𝐯𝐕h\bm{v}\in\bm{V}_{h}, there holds

(3.32) \wideparen𝒗Hh1(γ)\wideparen𝒗L2(γ)+Defγ,h\wideparen𝒗L2(γ),\displaystyle\|\wideparen\bm{v}\|_{H_{h}^{1}(\gamma)}\lesssim\|\wideparen\bm{v}\|_{L_{2}(\gamma)}+\|{\rm Def}_{\gamma,h}\wideparen\bm{v}\|_{L_{2}(\gamma)},

where \wideparen𝐯=𝒫𝐩𝐯\wideparen\bm{v}=\mathcal{P}_{\bm{p}}\bm{v}.

Proof.

Given 𝒗𝑽h\bm{v}\in\bm{V}_{h}, let \wideparen𝒗c𝑯T1(γ)\wideparen\bm{v}_{c}\in\bm{H}_{T}^{1}(\gamma) satisfy (3.8). A continuous Korn inequality holds for \wideparen𝒗c\wideparen\bm{v}_{c}, so using (3.8) we have

(3.33) \wideparen𝒗Hh1(γ)\wideparen𝒗cH1(γ)+\wideparen𝒗c\wideparen𝒗Hh1(γ)\wideparen𝒗cL2(γ)+Defγ\wideparen𝒗cL2(γ)+\wideparen𝒗c\wideparen𝒗Hh1(γ)\wideparen𝒗L2(γ)+Defγ,h\wideparen𝒗L2(γ)+\wideparen𝒗c\wideparen𝒗Hh1(γ)(1+h)\wideparen𝒗L2(γ)+Defγ,h\wideparen𝒗L2(γ).\begin{split}\|\wideparen\bm{v}\|_{H_{h}^{1}(\gamma)}&\lesssim\|\wideparen\bm{v}_{c}\|_{H^{1}(\gamma)}+\|\wideparen\bm{v}_{c}-\wideparen\bm{v}\|_{H_{h}^{1}(\gamma)}\\ &\lesssim\|\wideparen\bm{v}_{c}\|_{L_{2}(\gamma)}+\|{\rm Def}_{\gamma}\wideparen\bm{v}_{c}\|_{L_{2}(\gamma)}+\|\wideparen\bm{v}_{c}-\wideparen\bm{v}\|_{H_{h}^{1}(\gamma)}\\ &\lesssim\|\wideparen\bm{v}\|_{L_{2}(\gamma)}+\|{\rm Def}_{\gamma,h}\wideparen\bm{v}\|_{L_{2}(\gamma)}+\|\wideparen\bm{v}_{c}-\wideparen\bm{v}\|_{H_{h}^{1}(\gamma)}\\ &\lesssim(1+h)\|\wideparen\bm{v}\|_{L_{2}(\gamma)}+\|{\rm Def}_{\gamma,h}\wideparen\bm{v}\|_{L_{2}(\gamma)}.\end{split}

From this result, we obtain a discrete Korn inequality for 𝑽h\bm{V}_{h} on Γh\Gamma_{h}.

Lemma 3.6.

There holds

(3.34) 𝒗Hh1(Γh)𝒗L2(Γh)+DefΓh𝒗L2(Γh)𝒗𝑽h.\displaystyle\|\bm{v}\|_{H^{1}_{h}(\Gamma_{h})}\lesssim\|\bm{v}\|_{L_{2}(\Gamma_{h})}+\|{\rm Def}_{\Gamma_{h}}\bm{v}\|_{L_{2}(\Gamma_{h})}\qquad\forall\bm{v}\in\bm{V}_{h}.
Proof.

We apply Lemma 3.5, (2.11), and an inverse estimate:

𝒗Hh1(Γh)\displaystyle\|\bm{v}\|_{H^{1}_{h}(\Gamma_{h})} \wideparen𝒗Hh1(γ)\displaystyle\lesssim\|\wideparen\bm{v}\|_{H^{1}_{h}(\gamma)}
\wideparen𝒗L2(γ)+Defγ,h\wideparen𝒗L2(γ)\displaystyle\lesssim\|\wideparen\bm{v}\|_{L_{2}(\gamma)}+\|{\rm Def}_{\gamma,h}\wideparen\bm{v}\|_{L_{2}(\gamma)}
𝒗L2(Γh)+DefΓh𝒗L2(Γh)+h𝒗Hh1(Γh)\displaystyle\lesssim\|\bm{v}\|_{L_{2}(\Gamma_{h})}+\|{{\rm Def}_{\Gamma_{h}}}\bm{v}\|_{L_{2}(\Gamma_{h})}+h\|\bm{v}\|_{H^{1}_{h}(\Gamma_{h})}
𝒗L2(Γh)+DefΓh𝒗L2(Γh).\displaystyle\lesssim\|\bm{v}\|_{L_{2}(\Gamma_{h})}+\|{{\rm Def}_{\Gamma_{h}}}\bm{v}\|_{L_{2}(\Gamma_{h})}.

4. Finite element method and convergence analysis

For piecewise smooth 𝒘,𝒗\bm{w},\bm{v} with 𝒗𝑯(divγ;γ)\bm{v}\in\bm{H}({\rm div}_{\gamma};\gamma) and qL2(γ)q\in L_{2}(\gamma), we define the bilinear forms

aγ(𝒘,𝒗)\displaystyle a_{\gamma}(\bm{w},\bm{v}) =γDefγ,h𝒘:Defγ,h𝒗+γ𝒘𝒗,\displaystyle=\int_{\gamma}{\rm Def}_{\gamma,h}\bm{w}:{\rm Def}_{\gamma,h}\bm{v}+\int_{\gamma}\bm{w}\cdot\bm{v},
bγ(𝒗,q)\displaystyle b_{\gamma}(\bm{v},q) =γ(divγ𝒗)q.\displaystyle=-\int_{\gamma}({\rm div}_{\gamma}\bm{v})q.

The variational formulation for the Stokes problem (1.1) seeks (𝒖,p)𝑯T1(γ)×L̊2(γ)(\bm{u},p)\in\bm{H}_{T}^{1}(\gamma)\times\mathring{L}_{2}(\gamma) satisfying

(4.1) aγ(𝒖,𝒗)+bγ(𝒗,p)\displaystyle a_{\gamma}(\bm{u},\bm{v})+b_{\gamma}(\bm{v},p) =γ𝒇𝒗\displaystyle=\int_{\gamma}{\bm{f}}\cdot\bm{v}\qquad 𝒗𝑯T1(γ),\displaystyle\forall\bm{v}\in\bm{H}_{T}^{1}(\gamma),
bγ(𝒖,q)\displaystyle b_{\gamma}(\bm{u},q) =0\displaystyle=0\qquad qL̊2(γ).\displaystyle\forall q\in\mathring{L}_{2}(\gamma).
Remark 4.1.

In order to ensure the well-posedness of (4.1) and avoid technical complications associated with Killing fields, we include the zeroth-order mass term in the momentum equations, as mentioned earlier in the introduction. A method for incorporating Killing fields into surface finite element methods for the Stokes problem is presented in [3], and the main ideas presented there are applicable to the proposed discretization below.

We define the analogous bilinear forms with respect to the discrete surface Γh\Gamma_{h}:

aΓh(𝒘,𝒗)\displaystyle a_{\Gamma_{h}}(\bm{w},\bm{v}) =ΓhDefΓh𝒘:DefΓh𝒗+Γh𝒘𝒗,\displaystyle=\int_{\Gamma_{h}}{\rm Def}_{\Gamma_{h}}\bm{w}:{\rm Def}_{\Gamma_{h}}\bm{v}+\int_{{\Gamma_{h}}}\bm{w}\cdot\bm{v},
bΓh(𝒗,q)\displaystyle b_{\Gamma_{h}}(\bm{v},q) =Γh(divΓh𝒗)q,\displaystyle=-\int_{\Gamma_{h}}({\rm div}_{\Gamma_{h}}\bm{v})q,

where the differential operator DefΓh{\rm Def}_{\Gamma_{h}} is understood to act piecewise with respect to 𝒯h\mathcal{T}_{h}. Then the finite element method seeks (𝒖h,ph)𝑽h×Qh(\bm{u}_{h},p_{h})\in\bm{V}_{h}\times Q_{h} such that

(4.2) aΓh(𝒖h,𝒗)+bΓh(𝒗,ph)\displaystyle a_{\Gamma_{h}}(\bm{u}_{h},\bm{v})+b_{\Gamma_{h}}(\bm{v},p_{h}) =Γh𝒇h𝒗\displaystyle=\int_{{\Gamma_{h}}}{\bm{f}}_{h}\cdot\bm{v}\qquad 𝒗𝑽h,\displaystyle\forall\bm{v}\in\bm{V}_{h},
bΓh(𝒖h,q)\displaystyle b_{\Gamma_{h}}(\bm{u}_{h},q) =0\displaystyle=0\qquad qQh,\displaystyle\forall q\in Q_{h},

where 𝒇h{\bm{f}}_{h} is some approximation of 𝒇{\bm{f}} that is defined on Γh\Gamma_{h}.

By the inf-sup condition (3.4), the discrete Korn-like inequality (3.34), and standard theory of saddle-point problems, there exists a unique solution (4.2). To derive error estimates, we restrict (4.2) to the discretely divergence–free subspace 𝑿h:={𝒗𝑽h:Γh(divΓh𝒗)q=0qQh}{\bm{X}}_{h}:=\{\bm{v}\in\bm{V}_{h}:\ \int_{\Gamma_{h}}({\rm div}_{\Gamma_{h}}\bm{v})q=0\ \forall q\in Q_{h}\}. Then 𝒖h𝑿h\bm{u}_{h}\in{\bm{X}}_{h} is uniquely determined by the problem

(4.3) aΓh(𝒖h,𝒗)=Γh𝒇h𝒗𝒗𝑿h.a_{\Gamma_{h}}(\bm{u}_{h},\bm{v})=\int_{\Gamma_{h}}{\bm{f}}_{h}\cdot\bm{v}\quad\forall\bm{v}\in{\bm{X}}_{h}.

Now set \wideparen𝒖h=𝒫𝒑𝒖h\wideparen{\bm{u}}_{h}=\mathcal{P}_{\bm{p}}\bm{u}_{h}, \wideparen𝒗=𝒫𝒑𝒗\wideparen{\bm{v}}=\mathcal{P}_{\bm{p}}\bm{v}, and note that

Γh𝒇h𝒗=Γh𝒇h𝒫𝒑1\wideparen𝒗=γ𝑭h\wideparen𝒗,\displaystyle\int_{\Gamma_{h}}{\bm{f}}_{h}\cdot\bm{v}=\int_{\Gamma_{h}}{\bm{f}}_{h}\cdot\mathcal{P}_{\bm{p}^{-1}}\wideparen{\bm{v}}=\int_{\gamma}{\bm{F}}_{h}\cdot\wideparen{\bm{v}},

where 𝑭h=(𝐌𝒇h){\bm{F}}_{h}=({\bf M}^{\intercal}{\bm{f}}_{h})^{\ell}, and

(4.4) 𝐌=[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1{\bf M}=\Big{[}{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\Big{]}[{\bf I}-d{\bf H}]^{-1}

is the matrix arising in the definition of 𝒫𝒑1\mathcal{P}_{\bm{p}^{-1}}. Therefore (4.3) is equivalent to the statement

(4.5) aγ(\wideparen𝒖h,\wideparen𝒗)=γ𝑭h\wideparen𝒗+Gh(𝒖h,𝒗)𝒗𝑿h,a_{\gamma}(\wideparen{\bm{u}}_{h},\wideparen{\bm{v}})=\int_{\gamma}{{\bm{F}}_{h}\cdot\wideparen{\bm{v}}}+G_{h}(\bm{u}_{h},\bm{v})\quad\forall{\bm{v}}\in{{\bm{X}}}_{h},

where the bilinear form Gh:𝑯h1(Γh)×𝑯h1(Γh)G_{h}:{\bm{H}^{1}_{h}(\Gamma_{h})\times\bm{H}^{1}_{h}(\Gamma_{h})}\to\mathbb{R} given by

Gh(𝒘,𝒗)=aγ(\wideparen𝒘,\wideparen𝒗)aΓh(𝒘,𝒗)G_{h}(\bm{w},\bm{v})=a_{\gamma}(\wideparen{\bm{w}},\wideparen{\bm{v}})-a_{\Gamma_{h}}(\bm{w},\bm{v})

encodes geometric error.

Lemma 4.2.

There holds

(4.6) |Gh(𝒘,𝒗)|h\wideparen𝒘Hh1(γ)\wideparen𝒗Hh1(γ)|G_{h}(\bm{w},\bm{v})|\lesssim h\|\wideparen\bm{w}\|_{H^{1}_{h}(\gamma)}\|\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}

for all tangential 𝐰,𝐯𝐇h1(Γh)\bm{w},\bm{v}\in\bm{H}^{1}_{h}(\Gamma_{h}).

Proof.

We write

γDefγ,h\wideparen𝒘:Defγ,h\wideparen𝒗ΓhDefΓh𝒘:DefΓh𝒗\displaystyle\int_{\gamma}{\rm Def}_{\gamma,h}\wideparen{\bm{w}}:{\rm Def}_{\gamma,h}\wideparen{\bm{v}}-\int_{\Gamma_{h}}{\rm Def}_{\Gamma_{h}}\bm{w}:{\rm Def}_{\Gamma_{h}}\bm{v}
=γDefγ,h\wideparen𝒘:Defγ,h\wideparen𝒗γ(μh1DefΓh𝒘)𝒑1:(DefΓh𝒗)𝒑1\displaystyle\qquad=\int_{\gamma}{\rm Def}_{\gamma,h}\wideparen{\bm{w}}:{\rm Def}_{\gamma,h}\wideparen{\bm{v}}-\int_{\gamma}(\mu_{h}^{-1}{\rm Def}_{\Gamma_{h}}\bm{w})\circ\bm{p}^{-1}:({\rm Def}_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}
=γ(Defγ,h\wideparen𝒘(μh1DefΓh𝒘)𝒑1):Defγ,h\wideparen𝒗\displaystyle\qquad=\int_{\gamma}\left({\rm Def}_{\gamma,h}\wideparen{\bm{w}}-(\mu_{h}^{-1}{\rm Def}_{\Gamma_{h}}\bm{w})\circ\bm{p}^{-1}\right):{\rm Def}_{\gamma,h}\wideparen{\bm{v}}
γ(μh1DefΓh𝒘)𝒑1:((DefΓh𝒗)𝒑1Defγ,h\wideparen𝒗).\displaystyle\qquad\qquad-\int_{\gamma}(\mu_{h}^{-1}{\rm Def}_{\Gamma_{h}}\bm{w})\circ\bm{p}^{-1}:\left(({\rm Def}_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}-{\rm Def}_{\gamma,h}\wideparen{\bm{v}}\right).

Applying (2.10), (2.2), and Lemma 2.1, we obtain

(4.7) |γDefγ,h\wideparen𝒘:Defγ,h\wideparen𝒗ΓhDefΓh𝒘:DefΓh𝒗|\displaystyle\left|\int_{\gamma}{\rm Def}_{\gamma,h}\wideparen{\bm{w}}:{\rm Def}_{\gamma,h}\wideparen{\bm{v}}-\int_{\Gamma_{h}}{\rm Def}_{\Gamma_{h}}\bm{w}:{\rm Def}_{\Gamma_{h}}\bm{v}\right| h\wideparen𝒘Hh1(γ)\wideparen𝒗Hh1(γ).\displaystyle\lesssim h\|\wideparen{\bm{w}}\|_{H^{1}_{h}(\gamma)}\|\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}.

Next, we use the formula of the Piola transform involving 𝐌{\bf M} to obtain

γ\wideparen𝒘\wideparen𝒗Γh𝒘𝒗\displaystyle\int_{\gamma}\wideparen{\bm{w}}\cdot\wideparen{\bm{v}}-\int_{\Gamma_{h}}\bm{w}\cdot\bm{v} =γ\wideparen𝒘\wideparen𝒗γ(μh𝒑1)\wideparen𝒘𝐌𝐌\wideparen𝒗\displaystyle=\int_{\gamma}\wideparen{\bm{w}}\cdot\wideparen{\bm{v}}-\int_{\gamma}(\mu_{h}\circ\bm{p}^{-1})\wideparen{\bm{w}}^{\intercal}{\bf M}^{\intercal}{\bf M}\wideparen{\bm{v}}
=γ\wideparen𝒘[𝚷(μh𝒑1)𝐌𝐌]\wideparen𝒗.\displaystyle=\int_{\gamma}\wideparen{\bm{w}}^{\intercal}[\bm{\Pi}-(\mu_{h}\circ\bm{p}^{-1}){\bf M}^{\intercal}{\bf M}]\wideparen{\bm{v}}.

A short computation using (2.2) yields |𝚷(μh𝒑1)𝐌𝐌||(𝝂𝝂h𝝂𝝂h)(𝝂𝝂h𝝂𝝂h)|+h2h2|\bm{\Pi}-(\mu_{h}\circ\bm{p}^{-1}){\bf M}^{\intercal}{\bf M}|\lesssim|({\bm{\nu}}-\frac{{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}})\otimes({\bm{\nu}}-\frac{{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}})|+h^{2}\lesssim h^{2}. Thus

(4.8) |γ\wideparen𝒘\wideparen𝒗Γh𝒘𝒗|h2\wideparen𝒘L2(γ)\wideparen𝒗L2(γ).\left|\int_{\gamma}\wideparen{\bm{w}}\cdot\wideparen{\bm{v}}-\int_{\Gamma_{h}}\bm{w}\cdot\bm{v}\right|\lesssim h^{2}\|\wideparen{\bm{w}}\|_{L_{2}(\gamma)}\|\wideparen{\bm{v}}\|_{L_{2}(\gamma)}.

The result (4.6) follows from (4.7)–(4.8). ∎

The next lemma states the approximation properties of the discretely divergence–free subspace 𝑿h{\bm{X}}_{h}. The result essentially follows from the inf-sup condition (3.4) and the arguments in [6, Theorem 12.5.17]. For completeness we provide the proof.

Lemma 4.3.

Let 𝐮𝐇1(γ)\bm{u}\in\bm{H}^{1}(\gamma) satisfy divγ𝐮=0{\rm div}\,_{\gamma}\bm{u}=0. Then there holds

inf𝒗𝑿h𝒖\wideparen𝒗Hh1(γ)inf𝒗𝑽h𝒖\wideparen𝒗Hh1(γ).\inf_{\bm{v}\in{\bm{X}}_{h}}\|\bm{u}-\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}\lesssim\inf_{\bm{v}\in\bm{V}_{h}}\|\bm{u}-\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}.
Proof.

Fix 𝒗𝑽h\bm{v}\in\bm{V}_{h}. The inf-sup condition (3.4) implies there exists 𝒘𝑽h\bm{w}\in\bm{V}_{h} such that bΓh(𝒘,q)=bγ(𝒖\wideparen𝒗,q)b_{\Gamma_{h}}(\bm{w},q)=b_{\gamma}(\bm{u}-\wideparen{\bm{v}},q^{\ell}) for all qQhq\in Q_{h}, and 𝒘Hh1(Γh)𝒖\wideparen𝒗Hh1(γ)\|\bm{w}\|_{H^{1}_{h}(\Gamma_{h})}\lesssim\|\bm{u}-\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}. Then 𝒘+𝒗𝑿h\bm{w}+\bm{v}\in{\bm{X}}_{h} and 𝒖(\wideparen𝒘+\wideparen𝒗)Hh1(γ)𝒖\wideparen𝒗Hh1(γ)+\wideparen𝒘Hh1(γ)𝒖\wideparen𝒗Hh1(γ)\|\bm{u}-(\wideparen{\bm{w}}+\wideparen{\bm{v}})\|_{H^{1}_{h}(\gamma)}\leq\|\bm{u}-\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}+\|\wideparen{\bm{w}}\|_{H^{1}_{h}(\gamma)}\lesssim\|\bm{u}-\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}. This implies the desired result. ∎

Theorem 4.4.

Let (𝐮h,ph)𝐕h×Qh(\bm{u}_{h},p_{h})\in\bm{V}_{h}\times Q_{h} satisfy the finite element method (4.2). Let \wideparen𝐮h=𝒫𝐩𝐮h\wideparen{\bm{u}}_{h}=\mathcal{P}_{\bm{p}}\bm{u}_{h} denote the Piola transform of 𝐮h\bm{u}_{h} with respect to the closest point projection 𝐩\bm{p}, and let ph=ph𝐩1p^{\ell}_{h}=p_{h}\circ\bm{p}^{-1}. Then there holds

(4.9a) 𝒖\wideparen𝒖hHh1(γ)\displaystyle\|\bm{u}-\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)} inf(𝒗,q)𝑽h×Qh(𝒖\wideparen𝒗Hh1(γ)+pqL2(γ))+h2𝒇L2(γ)+𝒇𝑭hL2(γ)\displaystyle\lesssim\inf_{(\bm{v},q)\in\bm{V}_{h}\times Q_{h}}\big{(}\|\bm{u}-\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}+\|p-q^{\ell}\|_{L_{2}(\gamma)}\big{)}+h^{2}\|{\bm{f}}\|_{L_{2}(\gamma)}+\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}
+h(pL2(γ)+𝒖H1(γ)+𝒇hL2(Γh)),\displaystyle\quad+h\big{(}\|p\|_{L_{2}(\gamma)}+\|\bm{u}\|_{H^{1}(\gamma)}+\|{\bm{f}}_{h}\|_{L_{2}(\Gamma_{h})}\big{)},
(4.9b) pphL2(γ)\displaystyle\|p-p_{h}^{\ell}\|_{L_{2}(\gamma)} infqQhpqL2(γ)+𝒖\wideparen𝒖hHh1(γ)+h2𝒇L2(γ)+𝒇𝑭hL2(γ)\displaystyle\lesssim\inf_{q\in Q_{h}}\|p-q^{\ell}\|_{L_{2}(\gamma)}+\|\bm{u}-\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)}+h^{2}\|{\bm{f}}\|_{L_{2}(\gamma)}+\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}
+h(pL2(γ)+𝒖H1(γ)+𝒇hL2(Γh)).\displaystyle\quad+h\big{(}\|p\|_{L_{2}(\gamma)}+\|\bm{u}\|_{H^{1}(\gamma)}+\|{\bm{f}}_{h}\|_{L_{2}(\Gamma_{h})}\big{)}.

Therefore, by Lemma 3.2, if (𝐮,p)𝐇2(γ)×H1(γ)(\bm{u},p)\in\bm{H}^{2}(\gamma)\times H^{1}(\gamma), there holds

(4.10) 𝒖\wideparen𝒖hHh1(γ)+pphL2(γ)\displaystyle\|\bm{u}-\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)}+\|p-p_{h}^{\ell}\|_{L_{2}(\gamma)} h(𝒖H2(γ)+pH1(γ)+𝒇hL2(Γh))+𝒇𝑭hL2(γ).\displaystyle\lesssim h(\|\bm{u}\|_{H^{2}(\gamma)}+\|p\|_{H^{1}(\gamma)}+\|{\bm{f}}_{h}\|_{L_{2}(\Gamma_{h})})+\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}.
Proof.

For 𝒗𝑿h\bm{v}\in{\bm{X}}_{h}, we denote by \wideparen𝒗c𝑯T1(γ)\wideparen{\bm{v}}_{c}\in\bm{H}_{T}^{1}(\gamma) the conforming relative of \wideparen𝒗=𝒫𝒑𝒗\wideparen{\bm{v}}=\mathcal{P}_{\bm{p}}\bm{v} satisfying (3.8). Using (4.1), (4.5) and (2.8), we write

aγ(𝒖\wideparen𝒖h,\wideparen𝒗)\displaystyle a_{\gamma}(\bm{u}-\wideparen{\bm{u}}_{h},\wideparen{\bm{v}}) =aγ(𝒖,\wideparen𝒗c)+aγ(𝒖,\wideparen𝒗\wideparen𝒗c)γ𝑭h\wideparen𝒗Gh(𝒖h,𝒗)\displaystyle=a_{\gamma}(\bm{u},\wideparen{\bm{v}}_{c})+a_{\gamma}(\bm{u},\wideparen{\bm{v}}-\wideparen{\bm{v}}_{c})-\int_{\gamma}{\bm{F}}_{h}\cdot\wideparen{\bm{v}}-G_{h}(\bm{u}_{h},\bm{v})
=γ𝒇\wideparen𝒗cγ𝑭h\wideparen𝒗bγ(\wideparen𝒗c,p)+aγ(𝒖,\wideparen𝒗\wideparen𝒗c)Gh(𝒖h,𝒗)\displaystyle=\int_{\gamma}{\bm{f}}\cdot\wideparen{\bm{v}}_{c}-\int_{\gamma}{\bm{F}}_{h}\cdot\wideparen{\bm{v}}-b_{\gamma}(\wideparen{\bm{v}}_{c},p)+a_{\gamma}(\bm{u},\wideparen{\bm{v}}-\wideparen{\bm{v}}_{c})-G_{h}(\bm{u}_{h},\bm{v})
=γ𝒇(\wideparen𝒗c\wideparen𝒗)γ(𝑭h𝒇)\wideparen𝒗bγ(\wideparen𝒗,pq)bγ(\wideparen𝒗c\wideparen𝒗,p)\displaystyle=\int_{\gamma}{\bm{f}}\cdot(\wideparen{\bm{v}}_{c}-\wideparen{\bm{v}})-\int_{\gamma}({\bm{F}}_{h}-{\bm{f}})\cdot\wideparen{\bm{v}}-b_{\gamma}(\wideparen{\bm{v}},p-q^{\ell})-b_{\gamma}(\wideparen{\bm{v}}_{c}-\wideparen{\bm{v}},p)
+aγ(𝒖,\wideparen𝒗\wideparen𝒗c)Gh(𝒖h,𝒗)qQh.\displaystyle\qquad+a_{\gamma}(\bm{u},\wideparen{\bm{v}}-\wideparen{\bm{v}}_{c})-G_{h}(\bm{u}_{h},\bm{v})\qquad\forall q\in Q_{h}.

Applying continuity estimates of the bilinear forms, (3.8), and (4.6) yield

aγ(𝒖\wideparen𝒖h,\wideparen𝒗)\displaystyle a_{\gamma}(\bm{u}-\wideparen{\bm{u}}_{h},\wideparen{\bm{v}})
(h2𝒇L2(γ)+𝒇𝑭hL2(γ)+pqL2(γ)+hpL2(γ)+h𝒖H1(γ)+h\wideparen𝒖hHh1(γ))\wideparen𝒗Hh1(γ).\displaystyle~{}~{}\lesssim\big{(}h^{2}\|{\bm{f}}\|_{L_{2}(\gamma)}+\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}+\|p-q^{\ell}\|_{L_{2}(\gamma)}+h\|p\|_{L_{2}(\gamma)}+h\|\bm{u}\|_{H^{1}(\gamma)}+h\|\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)}\big{)}\|\wideparen\bm{v}\|_{H^{1}_{h}(\gamma)}.

The estimate \wideparen𝒖hHh1(γ)𝒖hHh1(Γh)𝒇hL2(Γh)\|\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)}\lesssim\|\bm{u}_{h}\|_{H^{1}_{h}(\Gamma_{h})}\lesssim\|{\bm{f}}_{h}\|_{L_{2}(\Gamma_{h})}, and standard arguments then yield

𝒖\wideparen𝒖hHh1(γ)\displaystyle\|\bm{u}-\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)} inf(𝒗,q)𝑿h×Qh(𝒖\wideparen𝒗Hh1(γ)+pqL2(γ))+h2𝒇L2(γ)+𝒇𝑭hL2(γ)\displaystyle\lesssim\inf_{(\bm{v},q)\in{\bm{X}}_{h}\times Q_{h}}\big{(}\|\bm{u}-\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}+\|p-q^{\ell}\|_{L_{2}(\gamma)}\big{)}+h^{2}\|{\bm{f}}\|_{L_{2}(\gamma)}+\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}
+h(pL2(γ)+𝒖H1(γ)+𝒇hL2(Γh)).\displaystyle\qquad+h(\|p\|_{L_{2}(\gamma)}+\|\bm{u}\|_{H^{1}(\gamma)}+\|{\bm{f}}_{h}\|_{L_{2}(\Gamma_{h})}).

The estimate (4.9a) then follows by applying Lemma 4.3.

For the pressure error, we similarly apply (2.8), (4.5), (3.8), and (4.5)–(4.6) to obtain for all 𝒗𝑽h\bm{v}\in\bm{V}_{h} and qQhq\in Q_{h},

bΓh(𝒗,phq)\displaystyle b_{\Gamma_{h}}(\bm{v},p_{h}-q) =Γh𝒇h𝒗aΓh(𝒖h,𝒗)bγ(\wideparen𝒗,q)\displaystyle=\int_{\Gamma_{h}}{\bm{f}}_{h}\cdot\bm{v}-a_{\Gamma_{h}}(\bm{u}_{h},\bm{v})-b_{\gamma}(\wideparen{\bm{v}},q^{\ell})
=γ𝑭h\wideparen𝒗aγ(\wideparen𝒖h,\wideparen𝒗)bγ(\wideparen𝒗,q)+Gh(𝒖h,𝒗)\displaystyle=\int_{\gamma}{\bm{F}}_{h}\cdot\wideparen{\bm{v}}-a_{\gamma}(\wideparen{\bm{u}}_{h},\wideparen{\bm{v}})-b_{\gamma}(\wideparen{\bm{v}},q^{\ell})+G_{h}(\bm{u}_{h},\bm{v})
=γ𝒇(\wideparen𝒗\wideparen𝒗c)+γ(𝑭h𝒇)\wideparen𝒗+aγ(𝒖\wideparen𝒖h,\wideparen𝒗)+bγ(\wideparen𝒗,pq)+Gh(𝒖h,𝒗)\displaystyle=\int_{\gamma}{\bm{f}}\cdot(\wideparen{\bm{v}}-\wideparen{\bm{v}}_{c})+\int_{\gamma}({\bm{F}}_{h}-{\bm{f}})\cdot\wideparen{\bm{v}}+a_{\gamma}(\bm{u}-\wideparen{\bm{u}}_{h},\wideparen{\bm{v}})+b_{\gamma}(\wideparen{\bm{v}},p-q^{\ell})+G_{h}(\bm{u}_{h},\bm{v})
aγ(\wideparen𝒖,\wideparen𝒗\wideparen𝒗c)bγ(\wideparen𝒗\wideparen𝒗c,p)\displaystyle\qquad-a_{\gamma}(\wideparen{\bm{u}},\wideparen{\bm{v}}-\wideparen{\bm{v}}_{c})-b_{\gamma}(\wideparen{\bm{v}}-\wideparen{\bm{v}}_{c},p)
(h2𝒇L2(γ)+𝒇𝑭hL2(γ)+𝒖\wideparen𝒖hHh1(γ)+pqL2(γ)\displaystyle\lesssim\Big{(}h^{2}\|{\bm{f}}\|_{L_{2}(\gamma)}+\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}+\|\bm{u}-\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)}+\|p-q^{\ell}\|_{L_{2}(\gamma)}
+h(\wideparen𝒖hHh1(γ)+𝒖H1(γ)+pL2(γ)))\wideparen𝒗Hh1(γ).\displaystyle\qquad+h\big{(}\|\wideparen\bm{u}_{h}\|_{H^{1}_{h}(\gamma)}+\|\bm{u}\|_{H^{1}(\gamma)}+\|p\|_{L_{2}(\gamma)}\big{)}\Big{)}\|\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}.

We conclude from the inf-sup condition (3.4) and the estimates \wideparen𝒖hHh1(γ)𝒇hL2(Γh)\|\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)}\lesssim\|{\bm{f}}_{h}\|_{L_{2}(\Gamma_{h})}, \wideparen𝒗Hh1(γ)𝒗Hh1(Γh)\|\wideparen{\bm{v}}\|_{H^{1}_{h}(\gamma)}\lesssim\|\bm{v}\|_{H^{1}_{h}(\Gamma_{h})} that

pphL2(γ)\displaystyle\|p-p_{h}^{\ell}\|_{L_{2}(\gamma)} pqL2(γ)+phqL2(γ)\displaystyle\leq\|p-q^{\ell}\|_{L_{2}(\gamma)}+\|p_{h}^{\ell}-q^{\ell}\|_{L_{2}(\gamma)}
pqL2(γ)+phqL2(Γh)\displaystyle\lesssim\|p-q^{\ell}\|_{L_{2}(\gamma)}+\|p_{h}-q\|_{L_{2}(\Gamma_{h})}
h2𝒇L2(γ)+𝒇𝑭hL2(γ)+𝒖\wideparen𝒖hHh1(γ)+pqL2(γ)\displaystyle\lesssim h^{2}\|{\bm{f}}\|_{L_{2}(\gamma)}+\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}+\|\bm{u}-\wideparen{\bm{u}}_{h}\|_{H^{1}_{h}(\gamma)}+\|p-q^{\ell}\|_{L_{2}(\gamma)}
+h(𝒖H1(γ)+pL2(γ)+𝒇hL2(γ)).\displaystyle\qquad+h\big{(}\|\bm{u}\|_{H^{1}(\gamma)}+\|p\|_{L_{2}(\gamma)}+\|{\bm{f}}_{h}\|_{L_{2}(\gamma)}\big{)}.

By taking the infimum over qQhq\in Q_{h} we obtain (4.9b). ∎

Remark 4.5.

In order to obtain a final O(h)O(h) energy error bound from (4.10) we must choose 𝒇h{\bm{f}}_{h} so that 𝒇𝑭hL2(γ)h\|{\bm{f}}-{\bm{F}}_{h}\|_{L_{2}(\gamma)}\lesssim h. A short calculation shows that 𝒇h=𝒫𝒑1𝒇{\bm{f}}_{h}=\mathcal{P}_{\bm{p}^{-1}}{\bm{f}} yields 𝒇𝐅hL2(γ)h2\|{\bm{f}}-{\bf F}_{h}\|_{L_{2}(\gamma)}\lesssim h^{2}; a variety of other choices also yield optimal convergence.

Remark 4.6.

Analysis of L2L_{2} errors in the velocity is the subject of ongoing work. Numerical experiments presented below indicate that 𝒖˘𝒖hL2(Γh)h2\|\breve{\bm{u}}-\bm{u}_{h}\|_{L_{2}(\Gamma_{h})}\lesssim h^{2}, as expected. However, the conforming approximation error estimate given in Lemma 3.4 seems insufficient to obtain an O(h2)O(h^{2}) convergence rate in L2L_{2}. In addition, the O(h)O(h) geometric error estimate in Lemma 2.2 is sufficient to establish optimal O(h)O(h) convergence in the energy norm, but not an optimal O(h2)O(h^{2}) L2L_{2} convergence rate. Obtaining O(h2)O(h^{2}) geometric error estimates sufficient to achieve optimal L2L_{2} convergence is likely possible using techniques introduced in [18] but is significantly more technical than the energy case analyzed here.

5. Numerical Experiments

In this section we briefly comment on the implementation of the finite element method (4.2) and then present numerical experiments demonstrating optimal convergence rates in the energy and L2L_{2} norms for both MINI and lowest-order Taylor-Hood elements.

5.1. Implementation notes

The main additional complication in the implementation of the surface MINI method as compared to the Euclidean case arises in choosing two individual degrees of freedom at each vertex and interpreting them on each incident element. In the Euclidean case it is natural to choose individual degrees of freedom to align with the canonical Euclidean basis vectors. Thus natural global basis functions corresponding to a vertex aa are (φa,0)(\varphi_{a},0)^{\top} and (0,φa)(0,\varphi_{a})^{\top}, with φa\varphi_{a} the usual affine hat function corresponding to aa. In the surface case there is no such natural choice, so at each vertex a𝒱ha\in\mathcal{V}_{h} one must first fix the master element KaK_{a} along with two arbitrary but mutually orthogonal unit vectors 𝒗1,a,Ka\bm{v}_{1,a,K_{a}} and 𝒗2,a,Ka\bm{v}_{2,a,K_{a}} tangent to KaK_{a}. The two global degrees of freedom corresponding to aa are then 𝒖h|Ka(a)𝒗1,a,Ka\bm{u}_{h}|_{K_{a}}(a)\cdot\bm{v}_{1,a,K_{a}} and 𝒖h|Ka(a)𝒗2,a,Ka\bm{u}_{h}|_{K_{a}}(a)\cdot\bm{v}_{2,a,K_{a}}.

Once KaK_{a}, 𝒗1,a,Ka\bm{v}_{1,a,K_{a}}, and 𝒗2,a,Ka\bm{v}_{2,a,K_{a}} are fixed, the Piola transform formula (2.12) is used to interpret these quantities appropriately on each element KaK\ni a. These bookkeeping steps are naturally implemented as a precomputation in which the necessary information is encoded into a DOF handler structure. The precomputation step costs O(#𝒱h)O(\#\mathcal{V}_{h}) and does not add significantly to the overall computational cost. Once this step is completed the rest of the FEM is implemented in a standard way, but using the DOF handler to correctly compute basis functions on each element.

We now describe more precisely some elements of the precomputation step.

Refer to caption

z^2\hat{z}_{2}

z^1\hat{z}_{1}

z^3\hat{z}_{3}

Figure 3. Degrees of freedom for the reference MINI element

Consider the reference element K^\hat{K} with associated natural degrees of freedom for the MINI element (cf. Figure 3). Given a vertex z^jK^{\hat{z}_{j}}\in\hat{K}, let ϕ^1,j\hat{\bm{\phi}}_{1,j} and ϕ^2,j\hat{\bm{\phi}}_{2,j} be the basis functions corresponding to the vertex degrees of freedom in Figure 3, i.e., ϕ^1,j(z^i)=(δij0)\hat{\bm{\phi}}_{1,j}(\hat{z}_{i})=\begin{pmatrix}\delta_{ij}\\ 0\end{pmatrix} and ϕ^2,j(z^i)=(0δij)\hat{\bm{\phi}}_{2,j}(\hat{z}_{i})=\begin{pmatrix}0\\ \delta_{ij}\end{pmatrix}. We translate vertex degrees of freedom from the reference element to physical elements as follows. For each vertex a𝒱ha\in\mathcal{V}_{h}:

  1. (1)

    Specify a master element KaaK_{a}\ni a.

  2. (2)

    Choose arbitrary unit orthogonal vectors 𝒗1,a,Ka\bm{v}_{1,a,K_{a}}, 𝒗2,a,Ka\bm{v}_{2,a,K_{a}} lying in the plane containing KaK_{a}.

  3. (3)

    For each triangle K𝒯aK\in\mathcal{T}_{a}, compute 𝒗i,a,K=aK𝒗i,a,Ka\bm{v}_{i,a,K}=\mathcal{M}_{a}^{K}\bm{v}_{i,a,K_{a}}, i=1,2i=1,2.

  4. (4)

    For each K𝒯aK\in\mathcal{T}_{a}, let jK{1,2,3}j_{K}\in\{1,2,3\} be the local numbering of aa in KK.

  5. (5)

    For each K𝒯aK\in\mathcal{T}_{a}, let 𝒫FK\mathcal{P}_{F_{K}} be as in (3.1). Solve for αi,,a,K\alpha_{i,\ell,a,K}, i,{1,2}i,\ell\in\{1,2\}, such that αi,1,a,K𝒫FKϕ^1,jK(z^jK)+αi,2,a,K𝒫FKϕ^2,jK(z^jK)=𝒗i,a,K{\alpha_{i,1,a,K}}\mathcal{P}_{F_{K}}\hat{\bm{\phi}}_{1,j_{K}}(\hat{z}_{j_{K}})+{\alpha_{i,2,a,K}}\mathcal{P}_{F_{K}}\hat{\bm{\phi}}_{2,j_{K}}(\hat{z}_{j_{K}})=\bm{v}_{i,a,K}. With degrees of freedom as pictured in Figure 3, this expression reduces to the linear system

    𝒫FK𝜶i,a,K=𝒗i,a,Kwith𝜶i,a,K=(αi,1,a,Kαi,2,a,K).\mathcal{P}_{F_{K}}{\bm{\alpha}}_{i,a,K}=\bm{v}_{i,a,K}\quad\text{with}\quad{\bm{\alpha}}_{i,a,K}=\begin{pmatrix}\alpha_{i,1,a,K}\\ \alpha_{i,2,a,K}\end{pmatrix}.

    This system is solved by application of the Moore-Penrose psuedoinverse (𝒫FK𝒫FK)1𝒫FK.(\mathcal{P}_{F_{K}}^{\intercal}\mathcal{P}_{F_{K}})^{-1}\mathcal{P}_{F_{K}}^{\intercal}.

The coefficients αi,,a,K\alpha_{i,\ell,a,K} serve as a “Rosetta stone” (or DOF handler) to translate the individual reference basis functions ϕ^i,j\hat{\bm{\phi}}_{i,j} elementwise to global basis functions. On the element KK the global basis functions corresponding to the vertex aKa\in K are concretely given by φa(𝒙)𝒫FK𝜶1,a,K\varphi_{a}(\bm{x})\mathcal{P}_{F_{K}}{\bm{\alpha}}_{1,a,K} and φa(𝒙)𝒫FK𝜶2,a,K\varphi_{a}(\bm{x})\mathcal{P}_{F_{K}}{\bm{\alpha}}_{2,a,K}, with φa\varphi_{a} the standard scalar affine hat function corresponding to aa. Recall that in the Euclidean case natural global basis functions corresponding to aa are φa(𝒙)(1,0)\varphi_{a}(\bm{x})(1,0)^{\intercal} and φa(𝒙)(0,1)\varphi_{a}(\bm{x})(0,1)^{\intercal}. Thus in both the Euclidean and surface cases, the global MINI basis functions may be expressed as the product of a scalar hat function and a vector specifying direction. However, in the surface case the vectors 𝒫FK𝜶i,a,K\mathcal{P}_{F_{K}}{\bm{\alpha}}_{i,a,K} in question are piecewise constant rather than globally constant in order to reflect variation of the tangent plane from element to element. Once these expressions for global basis functions are in hand, the other aspects of the finite element code are essentially standard. Note that sparsity patterns for the resulting system matrices are also similar to the Euclidean case, and the system solve generally has similar expense.

5.2. Numerical results

We take γ\gamma to be the ellipsoid given by Ψ(x,y,z):=x21.12+y21.22+z21.32=1\Psi(x,y,z):=\frac{x^{2}}{1.1^{2}}+\frac{y^{2}}{1.2^{2}}+\frac{z^{2}}{1.3^{2}}=1. The test solution is 𝒖=𝚷(z2,x,y)\bm{u}=\bm{\Pi}(-z^{2},x,y)^{\intercal}; cf. Figure 4. Note that 𝚷=𝐈𝝂𝝂\bm{\Pi}={\bf I}-{\bm{\nu}}\otimes{\bm{\nu}} with 𝝂=Ψ|Ψ|{\bm{\nu}}=\frac{\nabla\Psi}{|\nabla\Psi|} on γ\gamma, so 𝒖\bm{u} is componentwise a rational function and not a polynomial. The pressure is p=xy3+zp=xy^{3}+z. The incompressibility condition divγ𝒖=0{\rm div}_{\gamma}\bm{u}=0 does not hold, so the Stokes system must be solved with nonzero divergence constraint. We employed a MATLAB code built on top of the iFEM library [8].

Refer to caption
Figure 4. Test solution 𝒖\bm{u}

The left plot in Figure 5 depicts the convergence history for the MINI element on a sequence of uniformly refined meshes. Optimal convergence is clearly observed in both the energy and L2L_{2} norms, in particular O(h)O(h) for the energy norm 𝒖˘𝒖hHh1(Γh)+pephL2(Γh)\|\breve{\bm{u}}-\bm{u}_{h}\|_{H_{h}^{1}(\Gamma_{h})}+\|p^{e}-p_{h}\|_{L_{2}(\Gamma_{h})} along with O(h2)O(h^{2}) for the error 𝒖˘𝒖hL2(Γh)\|\breve{\bm{u}}-\bm{u}_{h}\|_{L_{2}(\Gamma_{h})}. Recall also that the pressure is approximated by affine functions, which can in theory approximate to order h2h^{2} in L2L_{2}. Convergence is generally restricted instead to order hh because the pressure is coupled to the velocity H1H^{1} norm in the error analysis, but superconvergence of order h3/2h^{3/2} may occur on sufficiently structured meshes [15]. We observe an initial superconvergent decrease of order h3/2h^{3/2} or higher, but the expected asymptotic rate of order hh is eventually seen; cf. [27] for discussion of similar phenomena in the Euclidean context.

Refer to caption
Refer to caption
Figure 5. Convergence for the MINI element (left) and 21\mathbb{P}^{2}-\mathbb{P}^{1} Taylor-Hood element (right)

We also approximated (𝒖,p)(\bm{u},p) using a 21\mathbb{P}^{2}-\mathbb{P}^{1} Taylor-Hood method. The discrete surface Γh\Gamma_{h} was taken to be a quadratic rather than affine approximation to γ\gamma in order to obtain a geometric error commensurate with the expected order of convergence for this element. Vertex degrees of freedom were defined as above, additionally taking into account the fact that the surface normal on a piecewise quadratic surface is, in contrast to the case of an affine surface, not elementwise constant. Quadratic Taylor-Hood vector fields have degrees of freedom at edge midpoints in addition to at vertices, and these were defined in a manner completely analogous to the vertex degrees of freedom. Because the Piola transform preserves normal continuity, this construction guarantees normal continuity at three points on each (closed) edge, thus ensuring 𝑯(divΓh;Γh)\bm{H}({\rm div}_{\Gamma_{h}};\Gamma_{h})-conformity (cf. Proposition 3.1). The right plot in Figure 5 exhibits the expected O(h2)O(h^{2}) convergence in the energy norm and O(h3)O(h^{3}) convergence for the L2L_{2} error in the velocity. This confirms that our methodology has applicability beyond the MINI element; error analysis and extension to other stable Stokes element pairs employing nodal degrees of freedom will be the subject of future work.

Appendix A Proof of Lemma 2.2

Proof.

We divide the proof of Lemma 2.2 into three steps.

Step 1: For a scalar function qq defined on Γh\Gamma_{h}, we have the identity [12, (2.2.19)]

γ(q𝒑1)=([𝐈d𝐇]1[𝐈𝝂h𝝂𝝂h𝝂]Γhq)𝒑1on γ.\nabla_{\gamma}(q\circ\bm{p}^{-1})=\big{(}[{\bf I}-d{\bf H}]^{-1}[{\bf I}-\frac{{\bm{\nu}}_{h}\otimes{\bm{\nu}}}{{\bm{\nu}}_{h}\cdot{\bm{\nu}}}]\nabla_{\Gamma_{h}}q\big{)}\circ\bm{p}^{-1}\qquad\text{on }\gamma.

Consequently, for 𝒗=(v1,v2,v3)𝑯T1(K)\bm{v}=(v_{1},v_{2},v_{3})^{\intercal}\in\bm{H}_{T}^{1}(K),

((𝒗𝒑1)𝚷)i,:\displaystyle(\nabla(\bm{v}\circ\bm{p}^{-1})\bm{\Pi})_{i,:} =(γ(vi𝒑1))\displaystyle=\big{(}\nabla_{\gamma}(v_{i}\circ\bm{p}^{-1})\big{)}^{\intercal}
=(([𝐈d𝐇]1[𝐈𝝂h𝝂𝝂h𝝂]Γhvi)𝒑1)\displaystyle=\Big{(}\big{(}[{\bf I}-d{\bf H}]^{-1}[{\bf I}-\frac{{\bm{\nu}}_{h}\otimes{\bm{\nu}}}{{\bm{\nu}}_{h}\cdot{\bm{\nu}}}]\nabla_{\Gamma_{h}}v_{i}\big{)}\circ\bm{p}^{-1}\Big{)}^{\intercal}
=((Γhvi)[𝐈𝝂h𝝂𝝂h𝝂][𝐈d𝐇])𝒑1\displaystyle=\Big{(}(\nabla_{\Gamma_{h}}v_{i})^{\intercal}[{\bf I}-\frac{{\bm{\nu}}_{h}\otimes{\bm{\nu}}}{{\bm{\nu}}_{h}\cdot{\bm{\nu}}}]^{\intercal}[{\bf I}-d{\bf H}]^{-\intercal}\Big{)}\circ\bm{p}^{-1}
=((Γhvi)[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1)𝒑1\displaystyle=\Big{(}(\nabla_{\Gamma_{h}}v_{i})^{\intercal}[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}][{\bf I}-d{\bf H}]^{-1}\Big{)}\circ\bm{p}^{-1}
=((𝒗𝚷h)i,:[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1)𝒑1.\displaystyle=\Big{(}(\nabla\bm{v}\bm{\Pi}_{h})_{i,:}[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}][{\bf I}-d{\bf H}]^{-1}\Big{)}\circ\bm{p}^{-1}.

Here, with an abuse of notation, we have suppressed the superscript for the extension 𝒗e\bm{v}^{e}. Thus, we have the identity

(A.1) (𝒗𝒑1)𝚷=(𝒗𝚷h[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1)𝒑1.\displaystyle\nabla(\bm{v}\circ\bm{p}^{-1})\bm{\Pi}=\Big{(}\nabla\bm{v}\bm{\Pi}_{h}[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}][{\bf I}-d{\bf H}]^{-1}\Big{)}\circ\bm{p}^{-1}.

Since 𝒗\bm{v} is tangential, there holds 𝒗=(𝚷h𝒗)=𝚷h𝒗\nabla\bm{v}=\nabla(\bm{\Pi}_{h}\bm{v})=\bm{\Pi}_{h}\nabla\bm{v}, because 𝚷h\bm{\Pi}_{h} is constant on KK. Thus,

(A.2) (𝒗𝒑1)𝚷=(Γh𝒗[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1)𝒑1.\displaystyle\nabla(\bm{v}\circ\bm{p}^{-1})\bm{\Pi}=\Big{(}\nabla_{\Gamma_{h}}\bm{v}[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}][{\bf I}-d{\bf H}]^{-1}\Big{)}\circ\bm{p}^{-1}.

Step 2: Write \wideparen𝒗=𝒫𝒑𝒗=(𝐋𝒗)𝒑1\wideparen\bm{v}=\mathcal{P}_{\bm{p}}\bm{v}=({\bf L}\bm{v})\circ\bm{p}^{-1} with 𝐋=μh1[𝚷d𝐇]{\bf L}={\mu_{h}^{-1}}[\bm{\Pi}-d{\bf H}]. We then have by (A.2),

γ\wideparen𝒗\displaystyle\nabla_{\gamma}\wideparen{\bm{v}} =𝚷\wideparen𝒗𝚷=𝚷(𝐋𝒗𝒑1)𝚷\displaystyle=\bm{\Pi}\nabla\wideparen{\bm{v}}\bm{\Pi}=\bm{\Pi}\nabla({\bf L}\bm{v}\circ\bm{p}^{-1})\bm{\Pi}
=𝚷𝐋(𝒗𝒑1)𝚷+𝚷𝐋𝒗𝒑1𝚷\displaystyle=\bm{\Pi}{\bf L}\nabla(\bm{v}\circ\bm{p}^{-1})\bm{\Pi}+\bm{\Pi}\nabla{\bf L}\bm{v}\circ\bm{p}^{-1}\bm{\Pi}
=𝚷𝐋(Γh𝒗[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1)𝒑1+𝚷𝐋𝒗𝒑1𝚷\displaystyle=\bm{\Pi}{\bf L}\left(\nabla_{\Gamma_{h}}\bm{v}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\right][{\bf I}-d{\bf H}]^{-1}\right)\circ\bm{p}^{-1}+\bm{\Pi}\nabla{\bf L}\bm{v}\circ\bm{p}^{-1}\bm{\Pi}
=𝐋(Γh𝒗[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1)𝒑1+𝚷(𝐋𝒑1)𝒗𝒑1𝚷,\displaystyle={\bf L}\left(\nabla_{\Gamma_{h}}\bm{v}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\right][{\bf I}-d{\bf H}]^{-1}\right)\circ\bm{p}^{-1}+\bm{\Pi}\nabla({\bf L}\circ\bm{p}^{-1})\bm{v}\circ\bm{p}^{-1}\bm{\Pi},

where

(A.3) (𝐋𝒗)i,j=k=13𝐋i,kxjvki,j=1,2,3.(\nabla{\bf L}\bm{v})_{i,j}=\sum_{k=1}^{3}\frac{{\partial}{\bf L}_{i,k}}{{\partial}x_{j}}v_{k}\qquad i,j=1,2,3.

We conclude, by adding and subtracting terms, that

γ\wideparen𝒗\displaystyle\nabla_{\gamma}\wideparen{\bm{v}} =(Γh𝒗)𝒑1+[𝐋𝚷h](Γh𝒗)𝒑1[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1\displaystyle=(\nabla_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}+[{\bf L}-\bm{\Pi}_{h}](\nabla_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\right][{\bf I}-d{\bf H}]^{-1}
+(Γh𝒗)𝒑1([𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1𝚷h)+𝚷(𝐋𝒑1)𝒗𝒑1𝚷.\displaystyle\qquad+(\nabla_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}\left(\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\right][{\bf I}-d{\bf H}]^{-1}-\bm{\Pi}_{h}\right)+\bm{\Pi}\nabla({\bf L}\circ\bm{p}^{-1})\bm{v}\circ\bm{p}^{-1}\bm{\Pi}.

Using |𝝂𝝂h|h|{\bm{\nu}}-{\bm{\nu}}_{h}|\lesssim h, |d|h2|d|\lesssim h^{2}, and (2.2), we have |𝐋𝚷h|h|{\bf L}-\bm{\Pi}_{h}|\lesssim h and |[𝐈𝝂h𝝂𝝂𝝂h][𝐈d𝐇]1𝚷h|h|[{\bf I}-\frac{{\bm{\nu}}_{h}\otimes{\bm{\nu}}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}][{\bf I}-d{\bf H}]^{-1}-\bm{\Pi}_{h}|\lesssim h. Therefore there holds

(A.4) |Defγ\wideparen𝒗(DefΓh𝒗)𝒑1|h|(Γh𝒗)𝒑1|+|𝚷(𝐋𝒑1)𝒗𝒑1𝚷|.\displaystyle|{\rm Def}_{\gamma}\wideparen\bm{v}-({\rm Def}_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}|\lesssim h|(\nabla_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}|+|\bm{\Pi}\nabla({\bf L}\circ\bm{p}^{-1})\bm{v}\circ\bm{p}^{-1}\bm{\Pi}|.

Step 3: In the final step of the proof, we bound the last term in (A.4).

Let (r)=𝐋:,r{\bm{\ell}}^{(r)}={\bf L}_{:,r} denote the rrth column of 𝐋{\bf L}. Then (A.3) and a short calculation yields

𝚷(𝐋𝒑1)𝒗𝒑1𝚷=r=13(𝚷((r)𝒑1)𝚷)vr𝒑1,\bm{\Pi}\nabla({\bf L}\circ\bm{p}^{-1})\bm{v}\circ\bm{p}^{-1}\bm{\Pi}=\sum_{r=1}^{3}(\bm{\Pi}\nabla({\bm{\ell}}^{(r)}\circ\bm{p}^{-1})\bm{\Pi})v_{r}\circ\bm{p}^{-1},

and so, by (A.1),

(A.5) 𝚷(𝐋𝒑1)𝒗𝒑1𝚷=r=13𝚷((r)𝚷h[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1vr)𝒑1=𝚷(𝐋𝒗𝚷h[𝐈𝝂𝝂h𝝂𝝂h][𝐈d𝐇]1)𝒑1.\begin{split}\bm{\Pi}\nabla({\bf L}\circ\bm{p}^{-1})\bm{v}\circ\bm{p}^{-1}\bm{\Pi}&=\sum_{r=1}^{3}\bm{\Pi}\left(\nabla{\bm{\ell}}^{(r)}\bm{\Pi}_{h}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\right]\big{[}{\bf I}-d{\bf H}\big{]}^{-1}v_{r}\right)\circ\bm{p}^{-1}\\ &=\bm{\Pi}\left(\nabla{\bf L}\bm{v}\bm{\Pi}_{h}\left[{\bf I}-\frac{{\bm{\nu}}\otimes{\bm{\nu}}_{h}}{{\bm{\nu}}\cdot{\bm{\nu}}_{h}}\right]\big{[}{\bf I}-d{\bf H}\big{]}^{-1}\right)\circ\bm{p}^{-1}.\end{split}

Taking the derivative of 𝐋i,k=μh1[𝚷i,kd𝐇i,k]{\bf L}_{i,k}=\mu^{-1}_{h}[\bm{\Pi}_{i,k}-d{\bf H}_{i,k}] yields

𝐋i,kxj\displaystyle\frac{{\partial}{\bf L}_{i,k}}{{\partial}x_{j}} =μh1(𝐋i,kμhxj+𝚷i,kxjdxj𝐇i,kd𝐇i,kxj)\displaystyle=\mu_{h}^{-1}\left(-{\bf L}_{i,k}\frac{{\partial}\mu_{h}}{{\partial}x_{j}}+\frac{{\partial}\bm{\Pi}_{i,k}}{{\partial}x_{j}}-\frac{{\partial}d}{{\partial}x_{j}}{{\bf H}_{i,k}}-d\frac{{\partial}{\bf H}_{i,k}}{{\partial}x_{j}}\right)
=μh1(𝐋i,kμhxj+νi𝐇k,j+νk𝐇i,j+νj𝐇i,k+d𝐇i,kxj).\displaystyle=-\mu_{h}^{-1}\left({\bf L}_{i,k}\frac{{\partial}\mu_{h}}{{\partial}x_{j}}+\nu_{i}{\bf H}_{k,j}+\nu_{k}{\bf H}_{i,j}+\nu_{j}{\bf H}_{i,k}+d\frac{{\partial}{\bf H}_{i,k}}{{\partial}x_{j}}\ \right).

Thus by (2.2) and (A.3), there holds

(A.6) 𝐋𝒗=μh1[(𝐋𝒗)μh+𝝂(𝐇𝒗)+(𝐇𝒗)𝝂+(𝝂𝒗)𝐇+d𝐇𝒗]=[(𝐋𝒗)μh+𝝂(𝐇𝒗)+(𝐇𝒗)𝝂]+O(h|𝒗|).\begin{split}\nabla{\bf L}\bm{v}&=-\mu^{-1}_{h}\left[({\bf L}\bm{v})\otimes\nabla\mu_{h}+{\bm{\nu}}\otimes({\bf H}\bm{v})+({\bf H}\bm{v})\otimes{\bm{\nu}}+({\bm{\nu}}\cdot\bm{v}){\bf H}+d\nabla{\bf H}\bm{v}\right]\\ &=-\left[({\bf L}\bm{v})\otimes\nabla\mu_{h}+{\bm{\nu}}\otimes({\bf H}\bm{v})+({\bf H}\bm{v})\otimes{\bm{\nu}}\right]+O(h|\bm{v}|).\end{split}

Write μh=𝝂𝝂h(1dκ1)(1dκ2)=𝝂𝝂hdet(𝐈d𝐇)\mu_{h}={\bm{\nu}}\cdot{\bm{\nu}}_{h}(1-d\kappa_{1})(1-d\kappa_{2})={\bm{\nu}}\cdot{\bm{\nu}}_{h}\det({\bf I}-d{\bf H}). Because 𝝂h{\bm{\nu}}_{h} is constant on KK and 𝐇𝝂=0{\bf H}{\bm{\nu}}=0, there holds (𝝂𝝂h)xk=𝝂h𝝂xk=(𝐇𝝂h)k=(𝐇(𝝂h𝝂))k=O(h)\frac{{\partial}({\bm{\nu}}\cdot{\bm{\nu}}_{h})}{{\partial}x_{k}}={\bm{\nu}}_{h}\cdot\frac{{\partial}{\bm{\nu}}}{{\partial}x_{k}}=({\bf H}{\bm{\nu}}_{h})_{k}=({\bf H}({\bm{\nu}}_{h}-{\bm{\nu}}))_{k}=O(h). Also by Jacobi’s formula and |d|h2|d|\lesssim h^{2},

xkdet(𝐈d𝐇)\displaystyle\frac{{\partial}}{{\partial}x_{k}}\det({\bf I}-d{\bf H}) =det(𝐈d𝐇)tr((𝐈d𝐇)1xk(𝐈d𝐇))=νktr(𝐇)+O(h2).\displaystyle=\det({\bf I}-d{\bf H}){\rm tr}\left(({\bf I}-d{\bf H})^{-1}\frac{{\partial}}{{\partial}x_{k}}\big{(}{\bf I}-d{\bf H}\big{)}\right)=-\nu_{k}{\rm tr}({\bf H})+O(h^{2}).

We then conclude using |1𝝂𝝂h|h2|1-{\bm{\nu}}\cdot{\bm{\nu}}_{h}|\lesssim h^{2} that

(A.7) μh=(𝝂𝝂h)𝝂tr(𝐇)+O(h)=𝝂tr(𝐇)+O(h).\nabla\mu_{h}=-({\bm{\nu}}\cdot{\bm{\nu}}_{h}){\bm{\nu}}{\rm tr}({\bf H})+O(h)=-{\bm{\nu}}{\rm tr}({\bf H})+O(h).

Combining (A.6)–(A.7) yields

(A.8) 𝐋𝒗\displaystyle\nabla{\bf L}\bm{v} =[tr(𝐇)(𝐋𝒗)𝝂𝝂(𝐇𝒗)(𝐇𝒗)𝝂]+O(h|𝒗|).\displaystyle=\left[{\rm tr}({\bf H})({\bf L}\bm{v})\otimes{\bm{\nu}}-{\bm{\nu}}\otimes({\bf H}\bm{v})-({\bf H}\bm{v})\otimes{\bm{\nu}}\right]+O(h|\bm{v}|).

We apply (A.8) to (A.5) along with the identity 𝚷𝝂=𝚷𝝂=0\bm{\Pi}{\bm{\nu}}=\bm{\Pi}^{\intercal}{\bm{\nu}}=0 and |𝚷𝚷h|h|\bm{\Pi}-\bm{\Pi}_{h}|\lesssim h to obtain

|𝚷(𝐋𝒑1)𝒗𝒑1𝚷|h|𝒗𝒑1|.\displaystyle|\bm{\Pi}\nabla({\bf L}\circ\bm{p}^{-1})\bm{v}\circ\bm{p}^{-1}\bm{\Pi}|\lesssim h|\bm{v}\circ\bm{p}^{-1}|.

Combining this with (A.4) yields the desired estimate

|Defγ\wideparen𝒗(DefΓh𝒗)𝒑1|h(|(Γh𝒗)𝒑1|+|𝒗𝒑1|).\displaystyle|{\rm Def}_{\gamma}\wideparen\bm{v}-({\rm Def}_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}|\lesssim h\big{(}|(\nabla_{\Gamma_{h}}\bm{v})\circ\bm{p}^{-1}|+|\bm{v}\circ\bm{p}^{-1}|\big{)}.

Acknowledgements

The authors thank Orsan Kilicer for assistance with numerical computations.

References

  • [1] D. N. Arnold, F. Brezzi, and M. Fortin, A stable finite element for the Stokes equations, Calcolo, 21 (1984), pp. 337–344 (1985).
  • [2] D. Boffi, F. Brezzi, L. F. Demkowicz, R. G. Durán, R. S. Falk, and M. Fortin, Mixed finite elements, compatibility conditions, and applications, vol. 1939 of Lecture Notes in Mathematics, Springer-Verlag, Berlin; Fondazione C.I.M.E., Florence, 2008. Lectures given at the C.I.M.E. Summer School held in Cetraro, June 26–July 1, 2006, Edited by Boffi and Lucia Gastaldi.
  • [3] A. Bonito, A. Demlow, and M. Licht, A divergence-conforming finite element method for the surface Stokes equation, SIAM J. Numer. Anal., 58 (2020), pp. 2764–2798.
  • [4] P. Brandner, T. Jankuhn, S. Praetorius, A. Reusken, and A. Voigt, Finite element discretization methods for velocity-pressure and stream function formulations of surface Stokes equations, SIAM J. Sci. Comput., 44 (2022), pp. A1807–A1832.
  • [5] P. Brandner and A. Reusken, Finite element error analysis of surface Stokes equations in stream function formulation, ESAIM Math. Model. Numer. Anal., 54 (2020), pp. 2069–2097.
  • [6] S. C. Brenner and L. R. Scott, The mathematical theory of finite element methods, vol. 15 of Texts in Applied Mathematics, Springer, New York, third ed., 2008.
  • [7] F. Camacho and A. Demlow, L2L_{2} and pointwise a posteriori error estimates for FEM for elliptic PDEs on surfaces, IMA J. Numer. Anal., 35 (2015), pp. 1199–1227.
  • [8] L. Chen, iFEM: An innovative finite element method package in Matlab, tech. rep., University of California-Irvine, 2009.
  • [9] B. Cockburn and A. Demlow, Hybridizable discontinuous Galerkin and mixed finite element methods for elliptic problems on surfaces, Math. Comp., 85 (2016), pp. 2609–2638.
  • [10] B. Cockburn, G. Kanschat, and D. Schotzau, A locally conservative LDG method for the incompressible Navier-Stokes equations, Math. Comp., 74 (2005), pp. 1067–1095.
  • [11] A. Demlow, Higher-order finite element methods and pointwise error estimates for elliptic problems on surfaces, SIAM J. Numer. Anal., 47 (2009), pp. 805–827.
  • [12] A. Demlow and G. Dziuk, An adaptive finite element method for the Laplace-Beltrami operator on implicitly defined surfaces, SIAM J. Numer. Anal., 45 (2007), pp. 421–442.
  • [13] G. Dziuk, Finite elements for the Beltrami operator on arbitrary surfaces, in Partial differential equations and calculus of variations, vol. 1357 of Lecture Notes in Math., Springer, Berlin, 1988, pp. 142–155.
  • [14] G. Dziuk and C. M. Elliott, Finite element methods for surface PDEs, Acta Numer., 22 (2013), pp. 289–396.
  • [15] H. Eichel, L. Tobiska, and H. Xie, Supercloseness and superconvergence of stabilized low-order finite element discretizations of the Stokes problem, Math. Comp., 80 (2011), pp. 697–722.
  • [16] T.-P. Fries, Higher-order surface FEM for incompressible Navier-Stokes flows on manifolds, Internat. J. Numer. Methods Fluids, 88 (2018), pp. 55–78.
  • [17] S. Gross, T. Jankuhn, M. A. Olshanskii, and A. Reusken, A trace finite element method for vector-Laplacians on surfaces, SIAM J. Numer. Anal., 56 (2018), pp. 2406–2429.
  • [18] P. Hansbo, M. G. Larson, and K. Larsson, Analysis of finite element methods for vector Laplacians on surfaces, IMA J. Numer. Anal., 40 (2020), pp. 1652–1701.
  • [19] H. Hardering and S. Praetorius, Tangential errors of tensor surface finite elements, IMA J. Numer. Anal., 43 (2023), pp. 1543–1585.
  • [20] T. Jankuhn, M. A. Olshanskii, and A. Reusken, Incompressible fluid problems on embedded surfaces: modeling and variational formulations, Interfaces Free Bound., 20 (2018), pp. 353–377.
  • [21] T. Jankuhn, M. A. Olshanskii, A. Reusken, and A. Zhiliakov, Error analysis of higher order trace finite element methods for the surface Stokes equation, J. Numer. Math., 29 (2021), pp. 245–267.
  • [22] V. John, A. Linke, C. Merdon, M. Neilan, and L. G. Rebholz, On the divergence constraint in mixed finite element methods for incompressible flows, SIAM Rev., 59 (2017), pp. 492–544.
  • [23] P. L. Lederer, C. Lehrenfeld, and J. Schöberl, Divergence-free tangential finite element methods for incompressible flows on surfaces, Internat. J. Numer. Methods Engrg., 121 (2020), pp. 2503–2533.
  • [24] I. Nitschke, A. Voigt, and J. Wensch, A finite element approach to incompressible two-phase flow on manifolds, J. Fluid Mech., 708 (2012), pp. 418–438.
  • [25] M. A. Olshanskii, A. Quaini, A. Reusken, and V. Yushutin, A finite element method for the surface Stokes problem, SIAM J. Sci. Comput., 40 (2018), pp. A2492–A2518.
  • [26] M. A. Olshanskii and V. Yushutin, A penalty finite element method for a fluid system posed on embedded surface, J. Math. Fluid Mech., 21 (2019), pp. Paper No. 14, 18.
  • [27] D. R. Q. Pacheco and O. Steinbach, On the initial higher-order pressure convergence in equal-order finite element discretizations of the Stokes system, Comput. Math. Appl., 109 (2022), pp. 140–145.
  • [28] A. Reusken, Stream function formulation of surface Stokes equations, IMA J. Numer. Anal., 40 (2020), pp. 109–139.
  • [29] 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 (1985), pp. 111–143.
  • [30] P. Steinmann, On boundary potential energies in deformational and configurational mechanics, J. Mech. Phys. Solids, 56 (2008), pp. 772–800.
  • [31] C. Taylor and P. Hood, A numerical solution of the Navier-Stokes equations using the finite element technique, Internat. J. Comput. & Fluids, 1 (1973), pp. 73–100.