Trimmed serendipity finite element differential forms

By Andrew Gillette and Tyler Kloefkorn

Abstract

We introduce the family of trimmed serendipity finite element differential form spaces, defined on cubical meshes in any number of dimensions, for any polynomial degree, and for any form order. The relation between the trimmed serendipity family and the (non-trimmed) serendipity family developed by Arnold and Awanou [Math. Comp. 83 (2014), pp. 1551–1570] is analogous to the relation between the trimmed and (non-trimmed) polynomial finite element differential form families on simplicial meshes from finite element exterior calculus. We provide degrees of freedom in the general setting and prove that they are unisolvent for the trimmed serendipity spaces. The sequence of trimmed serendipity spaces with a fixed polynomial order provides an explicit example of a system described by Christiansen and Gillette [ESAIM:M2AN 50 (2016), pp. 883–850], namely, a minimal compatible finite element system on squares or cubes containing order polynomial differential forms.

1. Introduction

The “Periodic table of the finite elements” Reference 8 identifies four families of polynomial differential form spaces: , , , and . The families and define finite element spaces on -simplices while and define finite element spaces on -dimensional cubes. In this paper, we present a fifth family, that is closely related to but distinct from the serendipity family Reference 5. In particular, the relationships between the families and are analogous to the relationships between and .

We first define the spaces as

where denotes the Koszul operator. The spaces nest in between serendipity spaces via the inclusions:

The exterior derivative makes into a cochain complex and the associated sequence

is exact. The spaces in the above sequence have minimal dimension for or in the following sense: the sequence is a minimal compatible finite element system on -cubes that contains for each . All the results just mentioned, as well as others identified in this paper, hold true if is put in place of and the spaces are taken over -simplices instead of -cubes. Since spaces have been called trimmed polynomial spaces, we refer to the spaces as trimmed serendipity spaces.

We describe the trimmed serendipity family of finite elements using the language of finite element exterior calculus (FEEC) Reference 6Reference 7. The FEEC framework has also been used to describe the famous elements of Nédélec Reference 21Reference 22, Raviart-Thomas Reference 23, and Brezzi-Douglas-Marini Reference 11, as well as the more recently defined elements of Arnold and Awanou Reference 4Reference 5. Here, we show how the FEEC framework can also describe the recently defined elements on squares of Arbogast and Correa Reference 2, the elements on squares and cubes of Cockburn and Fu Reference 17, and the virtual element serendipity spaces of Beirão da Veiga, Brezzi, Marini, and Russo Reference 10. A detailed comparison to these newer elements is given at the end of Section 2. Two key features of our approach that distinguish it from related papers are: (i) a generalized definition of degrees of freedom suitable for any , and ; and (ii) the extensive use of tools from exterior calculus, allowing generalization to arbitrary dimension and instant coordination with other results from FEEC.

Christiansen and Gillette Reference 14 raised the question of a minimal compatible finite element system on -cubes containing and computed the number of degrees of freedom that such a system would need to associate to the interior of an -cube, . While we do not use the harmonic extension approach of Reference 14 to construct the spaces, we do recover the expected degree of freedom counts associated to each piece of the cubical geometry. We state the dimension of for , , and in Table 1.

The spaces and are of potentially great interest to the computational electromagnetics community as they can be used in - and -conforming methods on meshes of affinely-mapped cubes. Their dimensions satisfy and as well as and , with equality only in the the case . Hence, a significant savings in degrees of freedom should be possible, compared to tensor product and even serendipity methods. At the end of Section 4, we present an example illustrating the reduction in the degrees of freedom in the context of a mixed method for Poisson’s problem.

The elements of most immediate relevance to applications are those for small values of and . We now examine some of these cases in greater detail, using a mix of exterior calculus and vector calculus notation. Formal definitions of the notation and generalized formulae using exclusively exterior calculus notation are given in Sections 24 and a description of how to convert between vector and exterior calculus notation is given in Appendix A.

The spaces

The element diagrams in Figure 1 indicate the association of degrees of freedom to portions of the square geometry for and . The degrees of freedom for are

The notation above should be interpreted as the vector proxies for the exterior derivative applied to homogenous polynomials of degree in two variables. Observe that if we exclude only the degrees of freedom associated to , we are left with the degrees of freedom for the regular serendipity space .

The spaces

Moving to cubes, element diagrams for the spaces are shown in Figure 2. In these figures, degrees of freedom associated to vertices, edges, or faces of the cube are shown on the front face only while the number of degrees of freedom associated to the interior of the cube are indicated by . Looking only at the front face degrees of freedom in Figure 2 for , we see exactly the same sequence as shown in the top row of Figure 1, reflecting the fact that the spaces have the trace property. We also observe that and . Further, the lowest order spaces also coincide with the tensor product differential form spaces, i.e., for .

The degrees of freedom for are

As in the case, we observe that removing the degrees of freedom associated to and leaves only the degrees of freedom for .

The degrees of freedom for are

Again, excluding the degrees of freedom associated to , we are left with the degrees of freedom for .

The remainder of the paper is organized as follows. In Section 2, we review relevant background and notation from finite element exterior calculus and compare the trimmed serendipity elements to other elements in the literature. In Section 3, we prove various properties of the spaces, including a formula to compute their dimension. In Section 4, we state a set of degrees of freedom and prove they are unisolvent for . We also explain and establish minimality in the context of compatible finite element systems. Finally, we summarize the key results of our work and give an outlook on the future directions they suggest in Section 5. Appendix A provides a detailed description of the relation between exterior calculus and vector calculus notation in the context studied here.

2. Notation and relation to prior work

We use the same notation as Arnold and Awanou Reference 5 and will now review the relevant definitions to aid in comparison to prior work. Let be a multi-index and let be a subset of consisting of distinct elements with . The form monomial is the differential -form on given by

The degree of is . The space of differential -forms with polynomial coefficients of homogeneous degree is denoted . A basis for this space is the set of form monomials such that and . The exterior derivative and Koszul operator are maps

In coordinates, they are defined on form monomials by

The notation indicates that the term is omitted from the wedge product. We will make frequent use of the homotopy formula in this context Reference 6, Theorem 3.1, which is also called Cartan’s magic formula:

As shown in Reference 6, equation (3.10), it follows that

The space of polynomial differential -forms of degree at most is

The definitions of and extend linearly over . As a consequence,

Observe that if can be written as both an image of and as an image of , then .

The “trimmed” space of polynomial differential -forms of degree at most is

The relation of the and spaces to the well-known Nédélec Reference 21Reference 22, Raviart-Thomas Reference 23, and Brezzi-Douglas-Marini Reference 11 elements on simplices is described in the work of Arnold, Falk and Winther Reference 6Reference 7 and summarized in the “Periodic table of the finite elements” Reference 8.

An essential precursor to the finite element exterior calculus framework just described is the work of Hiptmair Reference 19, in which the spaces were introduced under different notation. In place of the Koszul operator, Hiptmair uses a potential mapping, , which satisfies the formula , similar to Equation 2.4 but without the factor of . The mapping is defined in terms of a point inside a star-shaped domain whereas the Koszul operator implicitly chooses the origin as a reference point. Since there is no true inverse for the operator, using in place of could provide some additional insight, however, we have found the operator to be quite natural for characterizing the structure of polynomial finite element differential form spaces.

To build the serendipity spaces on -dimensional cubes, we need some additional definitions. Let denote the complement of , i.e., . The linear degree of is defined to be

Put differently, the linear degree of counts the number of entries in equal to 1, excluding entries whose indices appear in . Note that if , then and there is no “exclusion” in the counting of linear degree. Likewise, if , then and for any . The linear degree of the sum of two or more form monomials is defined as the minimum of the linear degrees of the summands.

The subset of that has linear degree at least is denoted

A key building block for both the serendipity and trimmed serendipity spaces is

The following proposition gives a simple and useful characterization of .

Proposition 2.1 (Reference 5, Proposition 3.1).

The space is the span of for all –form monomials with and .

Note that every element in lies in the range of . Using this fact, we develop some basic results about that will be useful in our development of the spaces. In the proof of Reference 5, Theorem 3.4, it is shown that

We can exclude images by from the right side, yielding

Further, by Equation 2.4, we have that . Since , we have , and thus

The space of serendipity differential -forms of order is given by

The fact that this sum is direct is proven in Reference 5. Note that the second summand vanishes when , since while the third summand vanishes when , since by definition. Given , the degree property from Reference 5, Theorem 3.2 ensures that

where denotes the Kronecker delta function.

The serendipity spaces satisfy an inclusion property Reference 5, Theorem 3.4:

and a subcomplex property Reference 5, Theorem 3.3:

They also satisfy a containment property with respect to , namely,

The proof is a direct consequence of Equation 2.7, Equation 2.14, and Equation 2.15.

The spaces can be collected into a cochain complex with decreasing , which we denote by . The resulting sequence, as well as those for and , augmented by in front of the first term, are all exact. Written out, these are

All the above sequences can serve as finite element subcomplexes of the deRham complex for a domain. Such subcomplexes help guide the selection of pairs of spaces for mixed finite element methods that have guaranteed stability and convergence properties.

Comparison to prior and contemporary work.

As mentioned in the introduction, there has been a recent spate of research into conforming finite elements on meshes of -dimensional cubes. The family was defined first in the -conforming () case in Reference 4 and then for any in Reference 5. The relation of the elements to the Brezzi-Douglas-Marini Reference 11 elements on rectangles is described in Reference 5 and by the “Periodic table of the finite elements” Reference 8. The relation between the trimmed and non-trimmed serendipity families is described by Lemma 3.4 below. For , we will see that , indicating that the trimmed and non-trimmed serendipity families are truly distinct.

By converting from exterior caclulus to vector calculus notation, we can identify the relation of the spaces to finite element spaces defined in recent work by Arbogast and Correa Reference 2 and by Cockburn and Fu Reference 17. Both works examine various families of elements, and each work presents one family that is essentially the same as the trimmed serendipity elements, as explained in the following propositions. Note that both sets of authors use to indicate polynomial degree, but we have changed the notation to to match the conventions of finite element exterior calculus. We use the notation to denote the cube .

Proposition 2.2.

Define the pair of spaces as in Reference 2. Let denote the application of the operator to all the vectors in , which has the effect of rotating each vector in the field by . Then, interpreted as differential forms via the flat operator, is identical to .

Proposition 2.3.

The sequence of spaces denoted in Reference 17, Theorem 3.3, interpreted as differential forms via the flat operator, is identical to the sequence

Further, the sequence denoted in Reference 17, Theorem 3.6, is identical to the sequence

Detailed proofs of both propositions are given in Appendix A. We can now make a precise statement about the novelty of the trimmed serendipity spaces. The spaces can be recognized as differential form analogues of (i) the mixed finite element method presented in Reference 2 when applied to affinely-mapped square meshes (as opposed to general quadrilateral meshes), and (ii) the second of the four families of elements on squares and cubes presented in Reference 17. For , the trimmed serendipity spaces are entirely new to the literature, modulo the fact that the and cases reduce to the non-trimmed serendipity spaces as described in Lemma 3.4.

Further comparison can be made in regard to degrees of freedom. We define degrees of freedom for in equation Equation 4.1 and prove they are unisolvent for in Theorem 4.2. The degrees of freedom given by Arbogast and Correa Reference 2 are slightly different, in that they are indexed in part by vectors of polynomials that vanish on certain edges of , whereas our degrees of freedom are indexed by spaces of polynomial differential forms without regard to the basis used to define them. The spaces of Cockburn and Fu Reference 17 are not equipped with degrees of freedom and so no comparison is possible in this case. Finally, we mention the virtual element space , recently defined in work by Beirão da Veiga et al. Reference 10. The number of degrees of freedom for this space appears to agree with the number of degrees of freedom for in the case of a square, the main difference being a vector calculus treatment of indexing spaces in place of the differential form terminology used here. Since the virtual element method does not employ spaces of local basis functions, further comparison between the methods is a larger question for future work.

3. The spaces

We define the trimmed serendipity spaces for , by

The trimmed serendipity spaces share many analogues with the trimmed polynomial spaces, as we now establish. Throughout, we fix the top dimension to be and omit the notation , except when it is needed for clarity.

Theorem 3.1 (Inclusion property).

Let , and . Then

Proof.

The first inclusion is immediate from Equation 3.1. For the second inclusion, the inclusion property Equation 2.17 implies that . Hence we only need to show that . Decomposing by Equation 2.15, we have , , and, by Equation 2.14, , thus completing the proof.

Theorem 3.2 (Subcomplex property).

Let , and . Then

Proof.

Using Equation 3.2 and Equation 2.18, we have .

We now develop a direct sum decomposition of whose proof is straightforward by virtue of an analogous decomposition of .

Theorem 3.3 (Direct sum decomposition).

Let , and . Then , as defined by Equation 3.1, can also be written as the direct sum

Further, any element can be written as , where and . In particular, and .

Proof.

First expand Equation 3.1 via Equation 2.15. Since , we have

By Equation 2.14, we can replace by . Further, by Equation 2.13, applied to the term and re-ordering, we now have

The first two terms summands give , which establishes Equation 3.4 as a summation formula.

We now show that Equation 3.4 is direct. Observe that , with directness of the sum following from the observation after Equation 2.7 that 0 is the only polynomial differential form in the image of both and . Therefore, the intersection of with is . Now, elements of are of degree at most while elements of are of degree at least . Similarly, elements of are of degree at most while elements of are of degree at least . Hence both pairs are direct sums and Equation 3.4 is established.

For the last statement, again consider the direct sum . Given , we can thus write such that and . We have and , as seen from Equation 2.15.

Lemma 3.4.

Let .

(i)

,

(ii)

,

(iii)

.

Proof.

For (i), note that by Equation 2.7 and by Equation 2.14. We decompose according to Equation 3.1 and then decompose the summands according to Equation 2.15, yielding

By Equation 2.12, and so . Part (ii) is an immediate consequence of Equation 3.1, since there are no -forms on .

For (iii), we have by Equation 3.2 and by the subcomplex property Equation 2.18. For the reverse containment, decompose the spaces as

Observe that by Equation 2.5 and Equation 2.6, by Equation 2.14, and appears as a summand for . Thus, by Equation 2.15, .

Theorem 3.5 (Exactness).

Let . The sequence

is exact.

Proof.

By Lemma 3.4, part (i), we can rewrite the beginning of the sequence as

The sequence is exact at since the incoming and outgoing maps at are the same as those in Equation 2.20, which is exact. For , we will show that

is exact at directly.

Let and assume . Using Equation 3.4, write , where

Thus and . Since has maximum polynomial degree and has minimum polynomial degree , we see that .

Recall from Equation 2.22 that is exact. Thus, there exists such that (in particular, with an appropriate coefficient suffices). Since , we can write for some polynomial -form . By hypothesis, , but is injective on the range of by Equation 2.4. Therefore, . Also, since , we can write , where , by Equation 3.4. Setting , we have .

The spaces also have a trace property analogous to the spaces. Recall that the trace of a differential -form on a codimension 1 hyperplane is the pullback of the form via the inclusion map . Let be a form monomial as in Equation 2.1 and let be the hyperplane defined by for some fixed and constant . Then

Theorem 3.6 (Trace property).

Let , and let be a hyperplane of obtained by fixing one coordinate. Then

Proof.

We use the result of Reference 5, Theorem 3.5 and techniques from its proof to derive the result. For a fixed constant , set . Using Equation 3.4, we need to show that the traces of , , and lie in

By Reference 6, Section 3.6, and by Reference 5, Theorem 3.5,

It remains to show that . Let be a –form monomial with and . By Proposition 2.1, it suffices to show that . Without loss of generality, we will write

where . Now, as in the proof of Reference 5, Theorem 3.5, we break into cases according to whether or not . If , we define by

If , then and we are done. So we presume , whence . If , and . If , and . Either way, . Thus, , by the characterization of from Proposition 2.1 and by the direct sum decomposition Equation 3.4, respectively.

Keeping as above, we now address the case . Define by

where , and . The sign of depends on the parity of the number of permutations required to reorder into . If , then . So we presume .

We will show that , , which by Equation 2.4 implies that . Note that , since , and by definition of . Thus, . By Proposition 2.1, and thus .

We split into cases one last time based on the inequality . If , then and we have . If , we have and . Since either preserves the linear degree of a form monomial or decreases it by one, we have . Thus . Again by Proposition 2.1, .

We now compute the dimension of from the direct sum decomposition Equation 3.4. The computation of in Reference 5 does not rely on its direct sum decomposition Equation 2.15 and, in particular, no formula for is provided. We now derive such a formula. We use and interchangeably to denote the dimension of as a vector space over .

Lemma 3.7.

Fix . For , , we have

where

Proof.

Observe that

Since is injective on the range of , we have and . Now, recall from Equation 2.20 that is exact. Thus,

By Equation 2.6 and Reference 6, Equation (3.14), we have

Define and for ease of notation. Using Equation 3.7 and the formula for given in Reference 5, we have

We can write as the telescoping sum

Using Equation 3.9 with Equation 3.8, we produce the formula in Equation 3.6.

Theorem 3.8.

Fix and . Then

Further, each summand in Equation 3.10 has a closed-form expression in terms of binomial coefficients depending only on , , and .

Proof.

Again, since is injective on the range of , we have . Using this with Equation 3.4, we can write

From Reference 6Reference 7, we have

We have the requisite expressions for and from Lemma 3.7.

We use Theorem 3.8 and Lemma 3.7 to compute the dimension of for , , and and report the results in Table 1.

4. Degrees of freedom, unisolvence, and minimality

We now state and count a set of degrees of freedom associated to . The degrees of freedom associated to a -dimensional subface of are

for any . Observe that the first summand of the indexing space is the indexing space for , reflecting the fact that . The sum is direct since . The dimension of is given (see, e.g., Reference 6) by

Applying Equation 4.2, we have that

It is shown in Reference 6, Theorem 3.3 that

and thus

Note that when , we have , so the dimension is zero. The above formula remains valid if we interpret as 0. There are -dimensional faces of so the total number of degrees of freedom in Equation 4.1 is

To prove that the degrees of freedom in Equation 4.1 are unisolvent for we will need to consider the subspace of that has vanishing trace on . For this, we will use the notation

The next result is the analogue of Reference 5, Proposition 3.7 for the family.

Lemma 4.1.

If and

then .

Proof.

Let . By the subcomplex property Equation 3.3, we have that . Recalling the definition and the fact that is injective on the range of , we have that , a non-trimmed serendipity space. Let be any –face of and recall that commutes with . Thus, , meaning .

By Stokes’ theorem, we have

Suppose so that

By Equation 4.4 and Equation 4.5, vanishes for all such and by the above equation vanishes for all such as well. Thus, by Reference 5, Proposition 3.7 with and replaced by and , respectively, we have .

By Theorem 3.3, we can write , where and . Since and is injective on the range of , we must have . Thus . Since Equation 4.4 holds, we can apply Reference 5, Proposition 3.7 with replaced by to conclude that .

We can now establish unisolvence in the classical sense, namely, that an element is uniquely determined by the values of the degrees of freedom applied to .

Theorem 4.2 (Unisolvence).

If and all the degrees of freedom in Equation 4.1 vanish, then .

Proof.

We use induction on . The base case is trivial. Let such that all the degrees of freedom in Equation 4.1 vanish. On a face of dimension , by the trace property Equation 3.5. Since all the degrees of freedom for vanish, by the inductive hypothesis. Thus, . By Lemma 4.1, .

The careful combinatorial argument carried out in Reference 5 to establish unisolvence for the spaces and their associated degrees of freedom is essential to the proof of unisolvence for the spaces just given, as it is invoked at the end of the proof of Lemma 4.1. Notably, our proof did not require the dimension of to equal the associated number of degrees of freedom as a hypothesis. We examine this point further in the discussion of future directions at the end of the paper and in Appendix B.

We now turn to the topic of the minimality of the spaces. For this, we will employ the theory of finite element systems, developed and applied by Christiansen and collaborators in Reference 12Reference 13Reference 14Reference 15Reference 16. We will not redefine the full framework here as we are interested only in a very specific context, similar to the examples studied in Reference 14.

We have shown that the spaces have the subcomplex and trace properties in Theorems 3.2 and 3.6, respectively. These properties ensure that the collection of spaces constitute a finite element system, for any fixed . Since the associated augmented co-chain complex for this sequence was shown to be exact in Theorem 3.5, the system is said to be locally exact. Whenever unisolvence holds in the sense established in Theorem 4.2 and the number of degrees of freedom equals the dimension of the associated trimmed serendipity spaces in the sequence, the system is said to admit extensions and be compatible. In such cases, we can apply the following result, specialized to the case of cubical meshes.

Lemma 4.3 (Reference 14, Corollary 3.2).

Suppose that is a finite element system on and that is a compatible finite element system containing . Suppose that

Then is minimal among compatible finite element systems containing .

In Equation 4.6, denotes the homology group of the system ; the subscript again indicates vanishing trace on all dimensional subfaces. Note that the system need not be locally exact and hence need not have vanishing homology. We can compute the dimension of the homology group by

We apply the lemma as follows.

Theorem 4.4 (Minimality).

For and , the system is a minimal compatible finite element system containing .

Remark 4.5.

Theorem 4.4 is stated as applying only to dimensions and , however, it holds in any setting for which the number of degrees of freedom equals the dimension of the associated trimmed serendipity spaces. This includes at least all values up to 100 for and . We discuss this point further in Section 5 and Appendix B.

Proof.

We set and and show that Lemma 4.3 applies. Note that is a non-compatible finite element system as it satisfies the subcomplex and trace properties but is not locally exact. We have already discussed why is a compatible finite element system and shown that . By Reference 14, Proposition 4.5,

In the proof of Reference 14, Lemma 4.13, it is shown that

By Reference 6, Theorem 3.3,

Replacing by and by , we have

Applying Theorem 4.2, we have

Therefore, Lemma 4.3 applies and minimality is proved.

We close this section with an examination of the computational benefit that using a minimal compatible finite element system can provide by comparing the use of trimmed serendipity elements in place of regular serendipity or tensor product (Nédélec) elements for a simple problem. Consider the standard mixed formulation of the Dirichlet problem for the Poisson equation on a cubical domain : Given , find and such that:

We remark that Equation 4.7Equation 4.8 is one instance of the Hodge Laplacian problem studied in finite element exterior calculus Reference 6Reference 7 and its analysis serves as the foundation for many applications, such as the movement of a fluid through porous media via Darcy flow Reference 3, diffusion via the heat equation Reference 9, wave propagation Reference 18, and various non-linear partial differential equations.

A finite element method for Equation 4.7Equation 4.8 is determined by selecting finite-dimensional subspaces and and solving the problem: find and such that:

Supposing that is meshed by cubes, we compare three choices for the pair with at least decay in the approximation of , , and in the appropriate norms: tensor product elements , serendipity elements , and trimmed serendipity elements . We report the number of degrees of freedom associated to a single mesh element in Table 2. As is evident from the table, the trimmed serendipity elements require the fewest degrees of freedom of any of the three choices. Notably, the trimmed serendipity choice uses the same number of degrees of freedom as the tensor product elements in the lowest order case () while using strictly fewer than either of the other choices in all other cases.

5. Summary, outlook, and future directions

In this paper, we have defined spaces of trimmed serendipity finite element differential forms on -dimensional cubes and demonstrated how their relation to the non-trimmed serendipity spaces are, in all essential ways, analogous to the relation of the trimmed and non-trimmed polynomial differential form spaces on simplices. Accordingly, it is natural to treat them as a “fifth column” of the “Periodic table of finite elements” Reference 8. The ease with which the trimmed serendipity spaces arise in the exterior calculus setting echoes the fact that instances of their vector calculus analogues have been discovered from the related but distinct frameworks of Arbogast and Correa Reference 2 and Cockburn and Fu Reference 17, as detailed in Appendix A.

A minor point mentioned after the proof of Theorem 4.2 hints at an important direction for future research. While we have shown that the degrees of freedom given in Equation 4.1 are unisolvent for , this only establishes that the number of degrees of freedom is greater than or equal to the dimension of . Using Mathematica, we verified that formula Equation 4.3 and the closed-form expression for from Theorem 3.8 are in fact equal for , , and , covering many more than the cases of practical relevance to modern applications. In the cases and , we also confirmed by direct proof that the spaces have the equal dimension for any ; these proofs appear in Appendix B.

A more promising approach toward the same goal is to construct a basis for , count its dimension, and sum over subfaces of . Such an approach is used by Arbogast and Correa Reference 2 in their study of mixed methods on quadrilaterals, but extending it to hexahedra, or even just to cubes, introduces significant additional subtleties regarding the linear independence of spanning sets of specific sets of polynomial differential forms. We plan to explore this approach in future work, not only to establish this particular equality, but also to provide a practical computational basis so that the trimmed serendipity spaces can begin to see their benefits realized in practical application settings.

Appendix A. Trimmed serendipity spaces in vector calculus notation

To characterize the relationship between the trimmed serendipity spaces of differential forms and finite element families described in traditional vector calculus notation, we will need some additional notation. First we recall classical notation for spaces of polynomials and polynomial vector fields as used in Reference 2. Let denote the space polynomials of degree at most and the space of homogeneous polynomials of degree exactly . The number of variables (typically two or three) is implied from context. For , define

The above definition extends to by using as the third component of the vector. For , define a “2D curl” operator on a scalar field as the gradient operator followed by a rotation of clockwise, i.e.,

Thus, we recover the statement for any .

To convert a vector field to its corresponding differential form, we use the flat operator, , following the conventions of Abraham et al. Reference 1; see also Hirani Reference 20. Given a scalar field on , there is an associated 0-form and an associated 2-form . It will be clear from context whether should be interpreted as a 0-form or a 2-form. Given a vector field on , define and . In , given a scalar field on , there is an associated 0-form and an associated 3-form . Given a vector field

on , the associated 1-form is defined by and the associated 2-form by . Again, the kind of flat operator to be used will be obvious from context.

Proof of Proposition 2.2.

The space from Reference 2 refers to a pair of spaces , where is the reference element. These spaces are defined to be

Since here corresponds to in finite element exterior calculus notation, we observe that . Further, we have , which is identical to by Lemma 3.4.

Turning to , observe that . Note that is an automorphism on , i.e., . Using and in place of and and omitting “span” notation for ease of reading, we also see that

Hence, we have .

The space is a space of “supplemental” vectors that satisfy certain conditions described in Reference 2. Foremost, the space satisfies the containment

Further, the elements of are required to have normal components on that are polynomials of degree . The authors present the following basis for for :

Looking at the homogeneous degree part of , we see that

Applying to both sides, we have that

Note that Footnote1 and . Hence, and similarly, . Observe that and are linearly independent and have distinct, non-zero projections onto . Thus, given a basis for , the set is a basis for and the set

1

Recall that if ; the relevant case here is .

is a basis for . This proves Proposition 2.2.

Proof of Proposition 2.3.

We now turn to the paper by Cockburn and Fu Reference 17. Rather than restate all their definitions, we translate their notation to the Arbogast-Correa notation or finite element exterior calculus notation as we analyze their spaces. First, we look at the sequence from their Theorem 3.3. Applying the flat operator to the first space, we get

Recall, by Lemma 3.4, that . The second space is written as a direct sum of three components. Applying the flat operator to each, we find that

By the direct sum decomposition Equation 3.4, we recognize that

since . For the third space, taking the operator for 2-forms, we have . This proves the first statement of Proposition 2.3.

The second statement of Proposition 2.3 can be established similarly. We state all the equivalencies first, then provide some details for the subtler cases:

The first and last statements are straightforward. For the -form case, Equation A.2, we first recognize that

We now claim that . It suffices to show that The space is defined as the span of polynomials of the form or , where denotes homogeneous polynomial of degree in variables and only, or of similar forms with the variables permuted. We have

We can write as the span of elements of the form or for any , or of similar forms with the variables permuted. Observe that and , both of which belong to . By similar analysis, after permuting variables, we have that . Next, the space is spanned by the set , . Taking of this set we get , establishing that . Since applying the flat operator to the elements of the spanning set for produces a spanning set for , we have established the claim.

Finally, we show that . The space is defined as the span of elements of the form or of two similar forms, with the variables permuted. Let , i.e., is a homogeneous polynomial of degree in variables and only. Observe that

Since is spanned by form monomials that can be written as and similar form monomials with the variables permuted, we have that

The last equality follows from Equation 2.11, since any element of has linear degree at most 1. By Equation 3.4, we have established Equation A.2. The final equality, Equation A.3, can be confirmed by similar analysis.

Appendix B. Proofs of dimension equality

We now prove that the number of degrees of freedom defined for the trimmed serendipity elements is equal to the dimension of the corresponding polynomial differential form space for and . In our experience, all intuition for the cardinalities of these sets comes from the geometry of the -cubes to which they are associated more so than the algebra of binomial coefficients required for their computation. Additional cases beyond those proved here can easily be checked using Mathematica or similar software, as we have done for and for .

Let denote the number of degrees of freedom associated to ; its value is defined by the formula Equation 4.3. Recall that can be computed using Equation 3.10 and that can be computed using Lemma 3.7.

Remark B.1.

In the following proofs, we adopt the convention:

This convention is strictly for notational convenience as we frequently encounter summations whose upper index limit depends on . For instance, the term appears in an expression for only when . By our convention, this summand is 0 when , whereas converting it to the polynomial and evaluating at gives a value of 1. Hence, we will only convert binomial coefficients to functions when doing so preserves the value according to the above convention. As we will see, this approach simplifies the presentation of the proofs.

Proposition B.1.

For , .

Proof.

We start with . Expanding Equation 4.3, we get

Note that our convention Equation B.1 applies to the third term in the sum, corresponding exactly to the summation index going from to . By Equation 3.4, we have that ; the term is the empty set. Using Equation 3.11 and Lemma 3.7, we compute