Partially discontinuous nodal finite elements for and
Abstract.
We investigate discretization of and in two and three space dimensions by partially discontinuous nodal finite elements, i.e., vector-valued Lagrange finite elements with discontinuity in certain directions. These spaces can be implemented as a combination of continuous and discontinuous Lagrange elements and fit in de Rham complexes. We construct well-conditioned nodal bases.
1. Introduction
The edge elements [33, 34] have achieved tremendous success in computational electromagnetism. These elements fix the notorious spurious solutions in both eigenvalue and source problems [3, 12, 41]. The key of this success is that the Nédélec element fits in a finite element de Rham complex (c.f., [41]). Even though discrete differential forms [5, 11, 25] have been a mature solution for many problems in computational electromagnetism, there are still interests in schemes with proper modifications of the vector-valued Lagrange elements for both theoretical and practical reasons. A theoretical question is what elements can solve the Maxwell equations without spurious solutions. In [41] the authors showed that the vector-valued Lagrange element on the Powell-Sabin mesh avoids spurious eigenmodes since the kernel of the operator is characterized by the Powell-Sabin element. In a modern language, this means that the vector-valued Lagrange space on this mesh fits in a smoother de Rham (Stokes) complex [10]. Nevertheless, the continuity may still lead to spurious solutions of the source problem on L-shaped domains [3, Chapter 5]. Practically, it is crucial to construct efficient bases for high order elements. This is always an art since one has to balance condition numbers, sparsity, fast evaluation, rotational symmetry, etc. See, e.g., [1, 30, 38, 42, 44, 46, 48, 49]. Nodal elements may also be preferred due to the canonical implementation [8, 14]. Finite elements with extra vertex/edge smoothness (supersmoothness) also find applications in solving the Hamilton-Jacobi-Bellman equations and may avoid averaging in semi-Lagrange methods [43].
This paper investigates another variant of the vector-valued Lagrange finite elements with the following desirable properties: 1) The elements allow discontinuity (not -conforming), such that spurious solutions arising from regularity should not appear. 2) The elements fit in complexes. 3) There exist Lagrange type nodal bases, and thus the construction (condition number, orthogonality) can be based on results of scalar problems, c.f., [31]. The idea of the construction is to allow discontinuity of certain components across the faces of the elements. We will refer to these elements as partially discontinuous nodal elements. In two space dimensions (2D), the Stenberg element [40] satisfies all these conditions. The corresponding complex was investigated in [16] (see also [22] for the cubical case). Nevertheless, the three dimensional (3D) case was left open in [16] (see Section 2.2 for a detailed review).
This paper aims to construct the 3D version of the partially discontinuous nodal finite elements. In particular, we obtain two finite element de Rham complexes: one contains an -conforming partially discontinuous nodal element, and another contains an version (note that they cannot fit in the same complex). The element was constructed in [16]. To the best of our knowledge, the complex constructed herein is new. Finite elements for with a higher order supersmoothness can be found in [16]. However, they do not admit Lagrange type bases. Complexes containing vector-valued Lagrange elements on special meshes were constructed in [21, 23], primally in the context of fluid problems. Nevertheless, the continuity does not exclude spurious solutions for the Maxwell source problems. In this paper, we obtain a new partially discontinuous nodal finite element and the corresponding complex on the Worsey-Farin split of tetrahedra.
Breaking some face degrees of freedom (e.g., treating the degrees of freedom of the tangential components as interior ones by allowing multi-values on neighboring elements) is equivalent to enriching face-based bubble functions. The idea of using matrix-valued Lagrange spaces enriched with symmetric matrix bubbles led to partially discontinuous nodal elements for the Hellinger-Reissner principle [26, 27, 28], see [15] for an extension of this idea.
The rest of this paper will be organized as follows. In Section 2, we introduce notation and review existing results. In Section 3 we present a new element and the corresponding complex. In Section 4, we construct a discrete de Rham complex which contains the nodal elements [16]. In Section 5, we construct well-conditioned bases for both 2D and 3D partially discontinuous nodal elemets. We end the paper with concluding remarks.
2. Preliminaries and existing results
2.1. Preliminaries
We assume that is a polygonal or polyhedral domain. Given a mesh, we will use to denote the set of vertices, for edges, for faces, and for the 3D simplices. We use , , , and to denote the number of vertices, edges, faces, and tetrahedra, respectively. From Euler’s formula, one has in 2D and in 3D for contractible domains.
For a triangle , we denote its Clough-Tocher split as , which is the collection of three triangles obtained by connecting the vertices to an interior point of . For a tetrahedron in , the Worsey-Farin split of , denoted by , is obtained by adding a vertex to the interior of , connecting this vertex to the vertices of ; then adding an interior vertex to each face of , and connecting this point to the vertices of the face and the interior point (see Figure 1). Therefore, each face is divided into three triangles like the Clough-Tocher split. The set of the Worsy-Farin splits of all the tetrahedra in is referred to as the Worsey-Farin mesh and denoted by . We refer to [21, 23, 32] for more details.
We use and to denote the unit normal and tangential vectors of a simplex , respectively. In 2D, the tangential and normal directions of an edge are uniquely defined up to an orientation. For edges in 3D there are one unit tangential and two linearly independent unit normal directions, and for faces in 3D there are one unit normal and two linearly independent unit tangential directions. We will write , and , , for these cases, respectively.
In our discussions, includes piecewise smooth functions for which derivatives up to order coincide at vertices. Similarly we can define and , for the space of functions with continuity on the edges, faces, and 3D cells, repectively.
In D (), we define the gradient operator and the operator . In 3D, for , we define
In 2D, we define
We review some basic facts about the de Rham complex; further details can be found, for instance, in [5, 13]. The de Rham complex consists of differential forms and exterior derivatives. In 3D, the de Rham complex on reads
(2.1) |
with . A complex is called exact at a space if the kernel of the subsequent operator is identical to the range of the previous one. For example, we say that the complex (2.1) is exact at if . Here for a linear operator , denotes the kernel of and denotes the range. The complex (2.1) is called exact if it is exact at all the spaces.
If there are finite element spaces , , , and such that
(2.2) |
is a complex, we say that (2.2) is a finite element subcomplex of (2.1). In this case, a necessary condition for the exactness is the following dimension condition
(2.3) |
As an example, the Lagrange element space, Nédélec element space of the first kind [33], Raviart-Thomas (RT) element space [33], and discontinuous element space fit into a discrete de Rham complex. The Lagrange element space, Nédélec element space of the second kind [34], Brezzi-Douglas-Marini (BDM) element space [34], and discontinuous element space also fit into a discrete de Rham complex. For a detailed discussion of these finite element spaces, see, e.g., [9].
We use to denote the polynomial space of degree (a nonnegative integer). Denote
as the shape function spaces of the first-kind Nédélec element with polynomial degree , and
as the shape function space of the RT element with the polynomial degree . The shape function space of the second-kind Nédélec element and the BDM element with the polynomial degree are
respectively. We denote the Nédélec finite element space of the first and second kind with the polynomial degree on mesh as and , respectively.
In 2D, the de Rham complex reads
or
2.2. Existing results
In this section, we briefly review the constructions in 2D in [16] and some questions that were left open there.
In 2D, the Stenberg element can be obtained by breaking the tangential component of a vector-valued Lagrange element on each interior edge of the mesh, i.e., the tangential trace of the functions in the Stenberg element space from two sides of an edge can be different. Therefore, the continuity of the Stenberg element is in between the corresponding vector-valued Lagrange element and the discontinuous element (see Figure 2). The Stenberg element fits in a discrete de Rham complex in 2D, which begins with the Hermite element taking vertex derivatives as degrees of freedom [16].

The 2D Stenberg element can be extended to 3D (see [16, Section 3.5] and in (4.1)). However, the complex containing it was not discussed in [16].
The construction in [16] contains two complexes in 3D: one begins with the Hermite element, and another begins with a element with second order supersmoothness at the vertices (formally, , although the function is not ). The latter can be viewed as a 3D extension of the Falk-Neilan complex for the Stokes problem [19] in terms of supersmoothness. The element in the former complex does not admit nodal basis, while the element in the latter complex can be represented in terms of scalar Hermite bases. Nevertheless, it was left open in [16] to construct a sequence containing an element with Lagrange bases. We solve these problems in this paper.
3. nodal elements in three space dimensions
3.1. Motivation
A natural candidate for a partially discontinuous nodal element is obtained by breaking the normal components of the vector-valued Lagrange finite element on each face (at the same time, the continuity at vertices and on edges is retained):
This element is in between the vector-valued Lagrange element and the Nédélec element in terms of continuity. Nevertheless, this element unlikely fits in a finite element de Rham complex due to the following observation. In most existing cases, the “restriction” (precise definition discussed in [17]) of a higher-dimensional finite element (complex) to a lower-dimensional cell is a lower-dimensional finite element (complex). This holds for the standard finite element differential forms [3], as well as all the examples in [16]. Therefore, to construct a complex in 3D, we want the restrictions of the involved spaces and the degrees of freedom on each face to form a 2D complex as well. Nevertheless, this is unlikely for the case of . Because the restriction of on each face is a 2D Lagrange element, which unlikely fits in a de Rham complex, as reflected in the deficiency of (at least for low order cases) as a discretization for the Stokes problem, c.f., [9, 24, 39].
Although the above observation already suggests that may not be a good candidate for an nodal element, one may still ask if , as a space in between the vector-valued Lagrange element (which produces spurious solutions) and the Nédélec element (which excludes spurious solutions), suffers from spurious solutions for eigenvalue problems. To the best of our knowledge, the emergence of spurious solutions for eigenvalue problems is not predicted by theory in most cases. Therefore we answer this question with numerical tests.
We apply the version of the Stenberg element in 2D defined by
and in 3D to solve the following Maxwell eigenvalue problem on the domain , :
(3.1) | |||
In 2D, the eigenvalues of this problems are [3], and the corresponding eigenvectors are
where are nonnegative integers, not both 0. In 3D, the eigenvalues of this problems are , and the corresponding eigenvectors are
where are nonnegative integers, not all 0.
We first apply the Stenberg finite element space to solve the problem (3.1) in 2D. The numerical results are shown in Table 1.
exact | 1 | 1 | 2 | 4 | 4 | 5 | 5 | 8 | 9 | 9 |
---|---|---|---|---|---|---|---|---|---|---|
1.0000 | 1.0000 | 1.1375 | 2.0000 | 2.5668 | 3.1367 | 4.0000 | 4.0000 | 4.1947 | 5.0000 | |
1.0000 | 1.0000 | 2.0000 | 4.0000 | 4.0000 | 5.0000 | 5.0000 | 8.0000 | 9.0001 | 9.0001 |
From the table, we see that spurious solutions do not appear. Therefore the vector-valued Lagrange elements are remedied by breaking the normal degrees of freedom on edges. Then we use to solve the problem (3.1) for . Table 2 shows the numerical eigenvalues around 3. In this case, spurious solutions occur. Therefore relaxing the normal continuity on faces (equivalently, enriching face-based bubbles) on the vector-valued Lagrange element is not enough to give a proper discretization of the Maxwell equations.
exact | 1 | 1 | 1 | 2 | 2 | 2 | 3 | 4 | 4 | 4 |
---|---|---|---|---|---|---|---|---|---|---|
numerical | 2.9924 | 2.9960 | 2.9972 | 3.0000 | 3.0000 | 3.0042 | 3.0057 | 3.0059 | 3.0069 | 3.0158 |
Now both heuristic argument and numerical experiments inspire us to seek a different construction other than . We cannot further relax the vertex and edge continuity, which is required to ensure the existence of nodal bases. Motivated by the heuristic argument on the face modes, now we plan to construct an element such that its restriction on each face is a proper velocity finite element for the Stokes problem. Recall that the - Scott-Vogelius pair on the Clough-Tocher split of triangular meshes (see Figure 3) leads to a complex and is thus stable [7, 36]. This motivates us to use the Worsey-Farin split of a tetrahedron [32, 45], which, restricted on each face of the macro tetrahedron, is a Clough-Tocher split.
The figure shows the lowest order case: the first space is piecewise cubic, the second is continuous piecewise quadratic and the third is piecewise linear.

3.2. Construction
We consider a complex for as follows:
(3.2) |
where each space is constructed on the Worsey-Farin split. The Worsey-Farin element [45] requires a mesh alignment, the subdivision point on each face of the macro tetrahedron lies on the line connecting the subdivision points of the two macro tetrahedra sharing the face. Nevertheless, such a condition is not required in the construction below.
Space .
The scalar space is a high order generalization of the Worsey-Farin macroelement from [45]. The shape function space on macroelement consists of piecewise polynomials of degree on the Worsey-Farin split, i.e.
Across the macroelements, the functions in are of continuity at vertices and edges, and of continuity across faces, that is
For the lowest order case (, c.f., [32]), if the mesh alignment condition holds, then the functions are also of continuity across the macroelements. The subsequent discussions do not rely on such smoothness across macroelements, therefore the mesh alignments are not required.
For , the degrees of freedom can be given by
-
•
function value and the first order derivatives of at each vertex ,
-
•
moments of on each edge ,
-
•
moments of the normal derivatives of on each edge ,
-
•
moments of on each face ,
where the space
consists of the bubble functions of the 2D Clough-Tocher element on the face subdivision , see the first element in Figure 3; -
•
interior degrees of freedom on ,
where
.
The dimension of the shape function space has been given in [37, Theorem 3.5] with a homological approach, that is,
(3.3) |
Note that the above degrees of freedom at vertices, edges, and faces are linearly independent since they are a part of the degrees of freedom for defining the element in [23]. Therefore the unisolvence holds and the dimension of the bubble space (number of interior degrees of freedom) is equal to the total dimension of (3.3) subtracting the number of degrees of freedom at vertices, edges, and faces. The number of face degrees of freedom, i.e., the dimension of can also be derived in a similar way based on the dimension formula of the 2D Clough-Tocher macroelement, c.f., [17, 37].
Therefore, we obtain the following dimension count of :
Remark 1.
As a tetrahedral spline, the shape function space of has been studied with the Bernstein-Bézier technique (see [32] for example). In [32], this space is defined with the continuity at the interior subdivision point. Nevertheless, the condition is intrinsic supersmoothness [20], meaning that it can be removed from the definition and the continuity would automatically imply the condition at the subdivision point. We refer to [20] and the references therein for more details on supersmoothnesses.
Space .
The shape function space on macroelement is define as
Then the space is defined as
The degrees of freedom are given by
-
•
function value of each component of at each vertex ,
-
•
moments of each component of on each edge ,
-
•
moments of the tangential components of on each face ,
where
and is considered as a 2D vector field; -
•
interior degrees of freedom.
The interior degrees of freedom consists of function values at interior Lagrange points of each and the values of the normal component at Lagrange points on each face of . Therefore, the dimension of is
Space .
From this step, the complex branches into the standard finite element de Rham complex of discrete differential forms [5] on the Worsey-Farin mesh.
More precisely, the space is the BDM element on the Worsey-Farin mesh of the domain , i.e.,
Since each macro tetrahedron is divided into micro tetrahedra, there are micro faces in the interior of and micro faces on each face of . As a result, it follows from the definition of the BDM element that the dimension count of is
Space .
The space consists of piecewise polynomials of degree on the Worsey-Farin mesh of . The dimension of is .
Lemma 1.
For , the sequence (3.2) is a complex and exact on contractible domains.
Proof.
By construction, (3.2) is a complex. It is sufficient to show that (3.2) is exact on contractible domains. Since the last two spaces and coincides with the BDM and the DG elements in the standard finite element de Rham complexes on the Worsey-Farin mesh [5], the operator is onto. This verifies the exactness at the space . The kernel of the operator is constants, which shows the exactness at the space . By definition,
where is the second kind Nédélec finite element space on the Worsey-Farin mesh of . Therefore, the kernel space of the operator is as follows
(3.4) |
where is the Lagrange element space of degree on the Worsey-Farin mesh of . The space coincides with , which shows the exactness at . Then a dimension count in the sense of (2.3) proves the exactness at .
∎
4. nodal elements in three space dimensions
In this section, we construct a discrete de Rham complex containing the partially discontinuous finite element from [16, Section 3.5]. In particular, we will design a complex for as follows
(4.1) |
Space .
For , the finite element space
coincides with the element in [16], which is also identical to each component of the velocity element in the discrete 3D Stokes complex by Neilan [35]. The space admits enhanced and continuities at vertices and edges, respectively, but only continuity across faces. Therefore no higher global continuity than is imposed.
The global dimension of on the mesh reads
Space .
The finite element space
consists of piecewise polynomials of degree less than or equal to for each component, and the degrees of freedom for can be given by
-
•
function values and the first order derivatives of at each vertex ,
-
•
moments of on each edge ,
-
•
moments of on each edge ,
where and are two linearly independent unit normal vectors of edge ;
-
•
integrals on each face ,
where is a face bubble associate with face , is the unit normal vector of , and
-
•
interior degrees of freedom in each ,
(4.2)
The dimension of is
Theorem 1.
The degrees of freedom for are unisolvent and is conforming.
Proof.
On a tetrahedron, the number of the above degree of freedom is
This coincides with the dimension of . Then it suffices to show that if all the degrees of freedom vanish on a function , then on .
Define as a 2D vector field on . The first step is to show that on each face of . Actually, from the degrees of freedom, we conclude that
-
(1)
for each vertex , it holds that where is the -th tangent vector on ;
-
(2)
for each edge , it holds that
-
(3)
in the interior of , it holds that and
The above functionals in (1) – (3) are the degrees of freedom of the shape function space , see [4].
To show , we first have and the vanishing degrees of freedom imply that on . Then it follows that for some . Without loss of generality, we assume that vanishes at one of the vertices. Now it follows from (1) – (3) that the following degrees of freedom vanish on ,
-
(1)
for each vertex , it holds that ;
-
(2)
for each edge , it holds that ;
-
(3)
in the interior of , it holds that
(4.3)
This shows that on all the edges of , which further implies that the tangential derivative of along each edge vanishes. Since vanishes at one vertex of , both and vanish on . Consequently, is of the form for some polynomial , which together with (4.3) yields . This implies , and further .
Space .
The element was defined in [16]:
where are two linearly independent unit normal vectors of an edge .
For , the degrees of freedom can be given by
-
•
values of at each vertex ,
-
•
moments of two normal components of on each edge ,
-
•
moments of the normal component of on each face ,
-
•
integrals over each tetrahedron ,
The dimension of is
Theorem 2.
The sequence (4.1) is a complex which is exact on contractible domains.
Proof.
To show that (4.1) is a complex, we only show that . This is a consequence of the identity
where is the unit normal vector of . Therefore the degrees of freedom of imply the continuity of the normal component of , meaning that .
As concluded in [16] is onto. Moreover, since is of continuity at vertices and of continuity on edges, the preimage of has to be at vertices and on edges.
Finally, a dimension count in the sense of (2.3) implies the exactness. ∎

5. Partially orthogonal nodal bases
To obtain practical bases for high order finite elements, one should take several issues into consideration, including the condition number, sparsity, symmetry and fast assembling. In particular, certain orthogonality is usually incorporated in the bases. Bases for high order Lagrange elements are relatively mature, see, e.g., [31]. In this section, we construct bases for partially discontinuous nodal and elements in both 2D and 3D based on explicit formulas of multi-variate orthogonal polynomials on simplices. We aim at orthogonality in the sense. The reason to use the orthogonality has been addressed in [29], see also [2, 47].
We choose proper Jacobi polynomials such that the bases associated to each simplex are mutually orthogonal on the reference element. As in the case of the elements, the condition numbers of the global mass and stiffness matrices still grow with the polynomial degree if no proper preconditioner is adopted .
To fix the notation, we use to denote the Jacobi polynomial on with weight , i.e.
where . By a simple transform , we obtain Jacobi polynomials on the reference interval :
where
We cite the explicit form of Jacobi polynomials on the reference simplex of dimension (c.f. [18], Section 5.3):
Theorem 3.
The following polynomials form a group of mutually orthonormal bases on the -dimensional reference simplex with weight :
(5.1) |
where and are multi-indexes of the polynomial degree and the weights, respectively, and . Here is a scaling factor [46]. When is scaled to be orthonormal, we have
In 2D we obtain a group of mutually orthogonal bases as follows
(5.2) |
From (5.2), we can obtain two groups of univariate polynomials which are orthogonal on a triangle. Let , we obtain the first group:
where is considered as the variable. For the second, we let and obtain
where and are considered as variables. Note that on a triangle, we get yet another formula with variables and :
(5.3) |
When , (5.3) is the scaled Legendre polynomials used in [49]. Due to the symmetry of variables, we will use (5.3) to build up edge based bases for triangular elements.
The 3D version of (5.1) reads
(5.4) |
We can similarly consider its restriction to an edge or a face, which will be used in the construction of edge and face bases for the tetrahedral elements.
5.1. Triangular element
-
(1)
Vertex-based basis functions at :
where is the barycentric coordinates at and is a 2D vector with the ith component 1 and the others 0.
-
(2)
Edge-based tangential bases on :
(5.5) where is the tangential vector of the edge and are the two barycentric coordinates of the two vertices of .
-
(3)
Edge-based normal bases on :
where and are the two elements sharing the edge and is the unit normal vector of .
-
(4)
Interior bases in (here, is a triangular mesh of a 2D domain):
where
and are the three barycentric coordinates of the three vertices of .
Remark 3.
The factor appearing in (5.5) is such that the bases decay to zero on the other two edges of the triangle. Then the index in the Jacobi polynomials is such that the bases are mutually orthogonal. A similar construction will also be used below for tetrahedral elements.
5.2. Tetrahedral element on the Worsey-Farin split
-
(1)
Vertex-based basis functions at :
where is the barycentric coordinates at and is a 3D vector with the ith component 1 and the others 0.
-
(2)
Edge-based basis functions on :
Here
where and are the barycentric coordinates of the two vertices of .
-
(3)
Face-based normal bases on , which is shared by two macroelements and :
-
•
at the face subdivision point,
where is the restriction of the hat function on one macroelement (a piecewise linear function on each macroelement);
-
•
on a face-interior edge of ,
where
with the barycentric coordinates and of the two vertices of ;
-
•
in the interior of a micro face of ,
with barycentric coordinates , , of each micro face of .
-
•
-
(4)
Face-based tangential bases on :
-
•
at the face subdivision point,
where is the hat function at the face subdivision point and are the two linearly independent unit tangent vectors of the face ;
-
•
on a face-interior edge ,
-
•
in the interior of each micro face of ,
-
•
-
(5)
Interior bases:
-
•
at the subdivision point,
where is the hat function at the interior subdivision point;
-
•
on an interior edge ,
-
•
in the interior of an interior micro face ,
where
with , and the barycentric coordinates associate with the interior face ;
-
•
in the interior of the micro tetrahedra ,
where
(5.6) and are the barycentric coordinates associate with the micro tetrahedron .
-
•
5.3. Tetrahedral elements
-
(1)
Vertex-based basis functions at :
-
(2)
Edge-based tangential bases on :
where is the tangent vector of , takes as an edge.
-
(3)
Edge-based normal bases on :
where are the two normal vectors of .
-
(4)
Face-based tangential bases on :
where are the two tangent vectors of , and are the two elements sharing . Moreover,
where , , and are the barycentric coordinates associated with ;
-
(5)
Face-based normal bases on :
where is the normal vector of .
- (6)
5.4. Condition number
We test the condition number of mass and quasi-stiffness matrices. The condition number of the mass matrix is calculated by
where , are the maximum and minimum eigenvalues of the matrix , respectively. For the quasi-stiffness matrix , only nonzero eigenvalues will be considered. We also consider the normalized mass and quasi-stiffness matrices and , whose nonzero diagonal entries are scaled to 1. The condition numbers of the four matrices are listed in Table 3. The condition numbers of the normalized mass and quasi-stiffness matrices are and when the reaches 9 and the condition number of is almost constant.
2 | 3.0345e+02 | 1.0121e+02 | 7.8806e+01 | 2.9527e+01 |
---|---|---|---|---|
3 | 9.1025e+03 | 2.2944e+02 | 3.2677e+02 | 5.8554e+00 |
4 | 7.7549e+05 | 1.9106e+03 | 9.4503e+03 | 1.9563e+01 |
5 | 6.9908e+06 | 2.7049e+03 | 1.3638e+05 | 1.3648e+01 |
6 | 1.1922e+08 | 5.3600e+03 | 1.6570e+06 | 1.4868e+01 |
7 | 1.9673e+09 | 7.4625e+03 | 2.0611e+07 | 1.6779e+01 |
9 | 3.2159e+10 | 1.2387e+04 | 2.5840e+08 | 1.9111e+01 |
6. Conclusions
This paper constructed partially discontinuous nodal finite elements for and by allowing discontinuity in the tangential or normal directions of vector-valued Lagrange elements, respectively. These elements can be implemented as a combination of standard continuous and discontinuous elements. The bases appear to be well-conditioned in the numerical tests, which shows the computational potential of these finite elements. Other bases, local complete sequences [38] and - preconditioning can be investigated in future research.
References
- [1] M. Ainsworth and J. Coyle, Hierarchic -edge element families for Maxwell’s equations on hybrid quadrilateral/triangular meshes, Computer Methods in Applied Mechanics and Engineering, 190 (2001), pp. 6709–6733.
- [2] M. Ainsworth and S. Jiang, Preconditioning the mass matrix for high order finite element approximation on tetrahedra, SIAM Journal on Scientific Computing, 43 (2021), pp. A384–A414.
- [3] D. N. Arnold, Finite element exterior calculus, SIAM, 2018.
- [4] D. N. Arnold, R. S. Falk, and R. Winther, Differential complexes and stability of finite element methods II: The elasticity complex, Compatible spatial discretizations, (2006), pp. 47–67.
- [5] , Finite element exterior calculus, homological techniques, and applications, Acta numerica, 15 (2006), p. 1.
- [6] D. N. Arnold and K. Hu, Complexes from complexes, Foundations of Computational Mathematics, (2021), pp. 1–36.
- [7] D. N. Arnold and J. Qin, Quadratic velocity/linear pressure Stokes elements, Advances in computer methods for partial differential equations, 7 (1992), pp. 28–34.
- [8] S. Badia and R. Codina, A nodal-based finite element approximation of the Maxwell problem suitable for singular solutions, SIAM Journal on Numerical Analysis, 50 (2012), pp. 398–417.
- [9] D. Boffi, F. Brezzi, and M. Fortin, Mixed finite elements for electromagnetic problems, in Mixed Finite Element Methods and Applications, Springer, 2013, pp. 625–662.
- [10] D. Boffi, J. Guzman, and M. Neilan, Convergence of Lagrange finite elements for the Maxwell eigenvalue problem in 2D, arXiv preprint arXiv:2003.08381, (2020).
- [11] A. Bossavit, Whitney forms: A class of finite elements for three-dimensional computations in electromagnetism, IEE Proceedings A-Physical Science, Measurement and Instrumentation, Management and Education-Reviews, 135 (1988), pp. 493–500.
- [12] , Solving Maxwell equations in a closed cavity, and the question of ’spurious modes’, IEEE Transactions on magnetics, 26 (1990), pp. 702–705.
- [13] R. Bott and L. W. Tu, Differential forms in algebraic topology, vol. 82, Springer Science & Business Media, 2013.
- [14] W. E. Boyse, D. R. Lynch, K. D. Paulsen, and G. N. Minerbo, Nodal-based finite-element modeling of Maxwell’s equations, IEEE Transactions on Antennas and Propagation, 40 (1992), pp. 642–651.
- [15] L. Chen and X. Huang, Geometric decompositions of div-conforming finite element tensors, arXiv preprint arXiv:2112.14351, (2021).
- [16] S. H. Christiansen, J. Hu, and K. Hu, Nodal finite element de Rham complexes, Numerische Mathematik, 139 (2018), pp. 411–446.
- [17] S. H. Christiansen and K. Hu, Generalized finite element systems for smooth differential forms and stokes’ problem, Numerische Mathematik, 140 (2018), pp. 327–371.
- [18] C. F. Dunkl and Y. Xu, Orthogonal polynomials of several variables, no. 155, Cambridge University Press, 2014.
- [19] R. S. Falk and M. Neilan, Stokes complexes and the construction of stable finite elements with pointwise mass conservation, SIAM Journal on Numerical Analysis, 51 (2013), pp. 1308–1326.
- [20] M. S. Floater and K. Hu, A characterization of supersmoothness of multivariate splines, Advances in Computational Mathematics, 46 (2020), pp. 1–15.
- [21] G. Fu, J. Guzmán, and M. Neilan, Exact smooth piecewise polynomial sequences on Alfeld splits, Mathematics of Computation, 89 (2020), pp. 1059–1091.
- [22] A. Gillette, K. Hu, and S. Zhang, Nonstandard finite element de rham complexes on cubical meshes, BIT Numerical Mathematics, 60 (2020), pp. 373–409.
- [23] J. Guzman, A. Lischke, and M. Neilan, Exact sequences on Worsey-Farin Splits, arXiv preprint arXiv:2008.05431, (2020).
- [24] J. Guzmán and R. Scott, The Scott-Vogelius finite elements revisited, arXiv preprint arXiv:1705.00020, (2017).
- [25] R. Hiptmair, Canonical construction of finite elements, Mathematics of Computation of the American Mathematical Society, 68 (1999), pp. 1325–1346.
- [26] J. Hu, Finite element approximations of symmetric tensors on simplicial grids in : the higher order case, Journal of Computational Mathematics, 33 (2015), pp. 283–296.
- [27] J. Hu and S. Zhang, A family of symmetric mixed finite elements for linear elasticity on tetrahedral grids, Science China Mathematics, 58 (2015), pp. 297–307.
- [28] J. Hu and S. Zhang, Finite element approximations of symmetric tensors on simplicial grids in : the lower order case, Mathematical Models and Methods in Applied Sciences, (2016).
- [29] K. Hu and R. Winther, Well-conditioned frames high order finite element methods, Journal of Computational Mathematics, 39 (2021), pp. 333–357.
- [30] E. Jørgensen, J. L. Volakis, P. Meincke, and O. Breinbjerg, Higher order hierarchical Legendre basis functions for electromagnetic modeling, Antennas and Propagation, IEEE Transactions on, 52 (2004), pp. 2985–2995.
- [31] G. Karniadakis and S. Sherwin, Spectral/hp element methods for computational fluid dynamics, Oxford University Press, 2013.
- [32] M.-J. Lai and L. L. Schumaker, Spline functions on triangulations, no. 110, Cambridge University Press, 2007.
- [33] J. Nédélec, Mixed finite elements in , Numerische Mathematik, 35 (1980), pp. 315–341.
- [34] , A new family of mixed finite elements in , Numerische Mathematik, 50 (1986), pp. 57–81.
- [35] M. Neilan, Discrete and conforming smooth de Rham complexes in three dimensions, Mathematics of Computation, (2015).
- [36] , The Stokes complex: A review of exactly divergence–free finite element pairs for incompressible flows, in 75 Years of Mathematics of Computation: Symposium on Celebrating 75 Years of Mathematics of Computation, November 1-3, 2018, the Institute for Computational and Experimental Research in Mathematics (ICERM), Providence, Rhode Island, vol. 754, American Mathematical Soc., 2020, p. 141.
- [37] H. Schenck and T. Sorokina, Subdivision and spline spaces, Constructive Approximation, 47 (2018), pp. 237–247.
- [38] J. Schöberl and S. Zaglmayr, High order Nédélec elements with local complete sequence properties, COMPEL-The international journal for computation and mathematics in electrical and electronic engineering, 24 (2005), pp. 374–384.
- [39] L. Scott and M. Vogelius, Norm estimates for a maximal right inverse of the divergence operator in spaces of piecewise polynomials, RAIRO-Modélisation mathématique et analyse numérique, 19 (1985), pp. 111–143.
- [40] R. Stenberg, A nonstandard mixed finite element family, Numerische Mathematik, 115 (2010), pp. 131–139.
- [41] D. Sun, J. Manges, X. Yuan, and Z. Cendes, Spurious modes in finite-element methods, IEEE Antennas and Propagation Magazine, 37 (1995), pp. 12–24.
- [42] D.-K. Sun, J.-F. Lee, and Z. Cendes, Construction of nearly orthogonal Nédélec bases for rapid convergence with multilevel preconditioned solvers, SIAM Journal on Scientific Computing, 23 (2001), pp. 1053–1076.
- [43] W. Tonnon, Semi-Lagrangian Discretization of the Incompressible Euler Equation, Master’s thesis, SAM, ETH Zürich, 2021.
- [44] J. P. Webb, Hierarchal vector basis functions of arbitrary order for triangular and tetrahedral finite elements, Antennas and Propagation, IEEE Transactions on, 47 (1999), pp. 1244–1253.
- [45] A. Worsey and G. Farin, An n-dimensional Clough-Tocher interpolant, Constructive Approximation, 3 (1987), pp. 99–110.
- [46] J. Xin and W. Cai, A well-conditioned hierarchical basis for triangular H(curl)-conforming elements, Commun. Comput. Phys, 9 (2011), pp. 780–806.
- [47] , Well-conditioned orthonormal hierarchical bases on simplicial elements, Journal of scientific computing, 50 (2012), pp. 446–461.
- [48] J. Xin, W. Cai, and N. Guo, On the construction of well-conditioned hierarchical bases for H(div)-conforming simplicial elements, Communications in Computational Physics, 14 (2013), pp. 621–638.
- [49] S. Zaglmayr, High order finite element methods for electromagnetic field computation, PhD thesis, JKU, Linz, 2006.