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

Fast solvers for the high-order FEM simplicial de Rham complex

Pablo D. Brubeck Mathematical Institute, University of Oxford, Oxford, UK brubeckmarti@maths.ox.ac.uk Patrick E. Farrell Mathematical Institute, University of Oxford, Oxford, UK patrick.farrell@maths.ox.ac.uk Robert C. Kirby Department of Mathematics, Baylor University, Waco, TX, US robert_kirby@baylor.edu  and  Charles Parker Ridgway Scott Foundation, 2544 S Forest Ln, Cedarville, MI charles_parker@alumni.brown.edu
(Date: September 28, 2025)
Abstract.

We present new finite elements for solving the Riesz maps of the de Rham complex on triangular and tetrahedral meshes at high order. The finite elements discretize the same spaces as usual, but with different basis functions, so that the resulting matrices have desirable properties. These properties mean that we can solve the Riesz maps to a given accuracy in a pp-robust number of iterations with 𝒪(p6)\mathcal{O}(p^{6}) flops in three dimensions, rather than the naïve 𝒪(p9)\mathcal{O}(p^{9}) flops.

The degrees of freedom build upon an idea of Demkowicz et al., and consist of integral moments on an equilateral reference simplex with respect to a numerically computed polynomial basis that is orthogonal in two different inner products. As a result, the interior-interface and interior-interior couplings are provably weak, and we devise a preconditioning strategy by neglecting them. The combination of this approach with a space decomposition method on vertex and edge star patches allows us to efficiently solve the canonical Riesz maps at high order. We apply this to solving the Hodge Laplacians of the de Rham complex with novel augmented Lagrangian preconditioners.

2020 Mathematics Subject Classification:
Primary 65F08, 65N35, 65N55
PBD was supported by EPSRC grant EP/W026260/1
PEF was supported by EPSRC grant EP/W026260/1, the Donatio Universitatis Carolinae Chair “Mathematical modelling of multicomponent systems”, and the UKRI Digital Research Infrastructure Programme through the Science and Technology Facilities Council’s Computational Science Centre for Research Communities (CoSeC)
CP acknowledges this material is based upon work supported by the National Science Foundation under Award No. DMS-2201487

1. Introduction

High-order finite element discretizations have very attractive properties, such as rapid convergence and high arithmetic intensity. However, their implementation requires great care, as naïve approaches to operator evaluation and linear system solution can scale badly in terms of memory and floating-point operations (flops) as the polynomial degree pp increases. On a single dd-dimensional cell, the number of degrees of freedom scales like pdp^{d}; computing the 𝒪(p2d)\mathcal{O}(p^{2d}) entries of a dense stiffness matrix with standard Lagrange-type elements would cost 𝒪(p3d)\mathcal{O}(p^{3d}) flops. Making high-order discretizations competitive therefore requires better algorithms.

For tensor-product cells (quadrilaterals and hexahedra), a tensor-product basis naturally exposes the required structure to decouple multidimensional problems into a sequence of one-dimensional problems, reducing computational cost and storage. For example, the sum-factorization assembly strategy enables fast matrix-free operator evaluation in 𝒪(pd+1)\mathcal{O}(p^{d+1}) operations and 𝒪(pd)\mathcal{O}(p^{d}) storage [42], and the fast diagonalization method (FDM) addresses the solution of separable problems on structured domains with the same complexities [36]. The previous work of the first two authors employing FDM-inspired finite element bases with suitable space decompositions yields the solution of certain separable and non-separable problems on unstructured domains with the same complexities [14, 15]. In particular, these complexities are achieved on tensor-product cells for the solution of the following partial differential equations (PDEs), posed on a bounded Lipschitz domain Ωd\Omega\subset\mathbb{R}^{d} with d=3d=3 for concreteness:

(1.1) βu(αu)\displaystyle\beta u-\nabla\cdot\left(\alpha\nabla u\right) =f in Ω,\displaystyle=f\mbox{\leavevmode\nobreak\ in\leavevmode\nobreak\ }\Omega,\quad u\displaystyle u =0 on ΓD,\displaystyle=0\mbox{\leavevmode\nobreak\ on\leavevmode\nobreak\ }\Gamma_{D},\quad αu𝐧\displaystyle\alpha\nabla u\cdot\mathbf{n} =0 on ΓN;\displaystyle=0\text{\leavevmode\nobreak\ on\leavevmode\nobreak\ }\Gamma_{N};
(1.2) β𝐮+×(α×𝐮)\displaystyle\beta\mathbf{u}+\nabla\times\left(\alpha\nabla\times\mathbf{u}\right) =𝐟 in Ω,\displaystyle=\mathbf{f}\mbox{\leavevmode\nobreak\ in\leavevmode\nobreak\ }\Omega, 𝐮×𝐧\displaystyle\mathbf{u}\times\mathbf{n} =0 on ΓD,\displaystyle=0\mbox{\leavevmode\nobreak\ on\leavevmode\nobreak\ }\Gamma_{D},\quad α×𝐮×𝐧\displaystyle\alpha\nabla\times\mathbf{u}\times\mathbf{n} =0 on ΓN;\displaystyle=0\text{\leavevmode\nobreak\ on\leavevmode\nobreak\ }\Gamma_{N};
(1.3) β𝐮(α𝐮)\displaystyle\beta\mathbf{u}-\nabla\left(\alpha\nabla\cdot\mathbf{u}\right) =𝐟 in Ω,\displaystyle=\mathbf{f}\mbox{\leavevmode\nobreak\ in\leavevmode\nobreak\ }\Omega,\quad 𝐮𝐧\displaystyle\mathbf{u}\cdot\mathbf{n} =0 on ΓD,\displaystyle=0\mbox{\leavevmode\nobreak\ on\leavevmode\nobreak\ }\Gamma_{D},\quad α𝐮\displaystyle\alpha\nabla\cdot\mathbf{u} =0 on ΓN.\displaystyle=0\text{\leavevmode\nobreak\ on\leavevmode\nobreak\ }\Gamma_{N}.

Here α,β:Ω+\alpha,\beta:\Omega\to\mathbb{R}_{+} are given coefficients, f:Ωf:\Omega\to\mathbb{R} and 𝐟:Ω3\mathbf{f}:\Omega\to\mathbb{R}^{3} are given source functions, ΓDΩ\Gamma_{D}\subseteq\partial\Omega, and ΓN=ΩΓD\Gamma_{N}=\partial\Omega\setminus\Gamma_{D}. For α=β=1\alpha=\beta=1, these equations are the so-called Riesz maps associated with subsets of the spaces H(grad,Ω)=H1(Ω)H(\operatorname{grad},\Omega)=H^{1}(\Omega), H(curl,Ω)H(\operatorname{curl},\Omega), and H(div,Ω)H(\operatorname{div},\Omega), respectively. For brevity we shall write H(grad)=H(grad,Ω)H(\operatorname{grad})=H(\operatorname{grad},\Omega) etc. where there is no potential confusion. These equations 1.1, 1.2 and 1.3 often arise as subproblems in the construction of preconditioners for more complex systems involving solution variables in H(grad)H(\operatorname{grad}), H(curl)H(\operatorname{curl}), and H(div)H(\operatorname{div}) [26, 31, 38, 37].

Achieving 𝒪(pd+1)\mathcal{O}(p^{d+1}) operations and 𝒪(pd)\mathcal{O}(p^{d}) storage on simplices is more challenging. Bases constructed from collapsed-coordinate mappings [30] give one possible approach, allowing a similar sum-factorization evaluation strategy. Perhaps surprisingly, Bernstein–Bézier expansions also admit sum-factorized matrix-free operator evaluation in 𝒪(pd+1)\mathcal{O}(p^{d+1}) operations and 𝒪(pd)\mathcal{O}(p^{d}) storage [1, 32], and such techniques have been generalized to the entire de Rham complex [2, 33]. For the H(grad)H(\operatorname{grad}) mass matrix (α=0\alpha=0) in 2D, there are 𝒪(p3)\mathcal{O}(p^{3}) solvers available [3]. However, there is no known solver for the linear systems arising from FEM discretizations of 1.1, 1.2 and 1.3 with these complexities. This manuscript makes a step towards addressing this challenge.

The essential idea of our approach is to devise new finite elements for the L2L^{2} de Rham complex with favorable properties for solvers. With the basis functions we propose, the interior and interface degrees of freedom are provably weakly coupled. This motivates the use of block preconditioning strategies that decouple the interface and interior degrees of freedom. The degrees of freedom yielding this weak coupling are designed within the framework of Demkowicz et al. [18], by making a clever choice for bases of bubble spaces that arise in the construction of the degrees of freedom. These bases for the bubble spaces are constructed by solving a handful of offline eigenproblems on the reference cell for each polynomial degree pp. This combination of finite elements and block preconditioning strategy solves the Riesz maps of the L2L^{2} de Rham complex in 𝒪(p3(d1))\mathcal{O}(p^{3(d-1)}) operations and 𝒪(p2(d1))\mathcal{O}(p^{2(d-1)}) storage, worse than the optimal complexities achieved on tensor-product cells, but orders better than the naïve approach. This is borne out in numerical experiments, as shown in Figure 1.

22 33 44 55 66 77 88 99 1010 10210^{-2}10110^{-1}10010^{0}10110^{1}10210^{2}6191ppGflopCGp(facet)\mathrm{CG}_{p}(\mathrm{facet})CGp(full)\mathrm{CG}_{p}(\mathrm{full})
(a)
22 33 44 55 66 77 88 99 1010 10210^{-2}10110^{-1}10010^{0}4161ppGbyte
(b)
22 33 44 55 66 77 88 99 1010 10410^{4}10510^{5}10610^{6}10710^{7}10810^{8}4161pp# nonzeros
(c)
22 44 33 55 66 77 88 99 1010 10110^{-1}10010^{0}10110^{1}10210^{2}10310^{3}6191ppGflopNedp1(facet)\mathrm{Ned}^{1}_{p}(\mathrm{facet})Nedp2(facet)\mathrm{Ned}^{2}_{p}(\mathrm{facet})Nedp1(full)\mathrm{Ned}^{1}_{p}(\mathrm{full})Nedp2(full)\mathrm{Ned}^{2}_{p}(\mathrm{full})
(d)
22 44 33 55 66 77 88 99 1010 10210^{-2}10110^{-1}10010^{0}10110^{1}4161ppGbyte
(e)
22 44 33 55 66 77 88 99 1010 10510^{5}10610^{6}10710^{7}10810^{8}10910^{9}4161pp# nonzeros
(f)
22 33 44 55 66 77 88 99 1010 10110^{-1}10010^{0}10110^{1}10210^{2}10310^{3}6191ppGflopRTp(facet)\mathrm{RT}_{p}(\mathrm{facet})BDMp(facet)\mathrm{BDM}_{p}(\mathrm{facet})RTp(full)\mathrm{RT}_{p}(\mathrm{full})BDMp(full)\mathrm{BDM}_{p}(\mathrm{full})
(g)
22 33 44 55 66 77 88 99 1010 10210^{-2}10110^{-1}10010^{0}10110^{1}4161ppGbyte
(h)
22 33 44 55 66 77 88 99 1010 10510^{5}10610^{6}10710^{7}10810^{8}10910^{9}4161pp# nonzeros
(i)
Figure 1. Flop counts, peak memory usage, and nonzeros in the sparse matrices and patch factors in the solution of the Riesz maps (α=β=1\alpha=\beta=1) on a Freudenthal mesh with 3 cells in each direction.

Several other works have addressed the same problem of solving the Riesz maps at high order. For 1.1, Šolín and Vejchodský [50] introduced the idea of using generalized eigenfunctions as interior basis functions, which we also employ in our construction, but they do not address how to construct efficient solvers. Also for 1.1, Casarin and Sherwin [49] constructed nonoverlapping Schwarz methods for the Schur complement system requiring exact interior solves; however, these interior solves require 𝒪(p3d)\mathcal{O}(p^{3d}) operations. Low-order refined preconditioning is provably effective on tensor-product cells [44], but the extension to simplicial cells is delicate and has only been studied for 1.1 on triangles [16]. Beuchler and coauthors construct sparse hierarchical bases for discretizations of the de Rham complex using Jacobi polynomials [7, 8, 9, 10]; this approach yields sparser matrices than ours, but the bases here concentrate the nonzero entries in the interface block, enabling the efficient interior-interface space decomposition described in Theorem 6.1 below.

A disadvantage of our approach is that the resulting preconditioners are not parameter-robust in α\alpha and β\beta. We design elements suitable for the stiffness-dominated case, so that the preconditioner is provably robust for β/α[0,1)\beta/\alpha\in[0,1). This is the more difficult case of wider interest (e.g. arising in augmented Lagrangian preconditioning). We anticipate that it is possible to design a basis for the mass-dominated case (β/α1\beta/\alpha\gg 1), but with very different properties.

2. Finite element discretization

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded Lipschitz domain with boundary ΓΩ\Gamma\coloneqq\partial\Omega. We denote the spaces consisting of kk-forms as H(dk,Ω)={vL2(Ω):dkvL2(Ω)}H(\operatorname{d}^{k},\Omega)=\{v\in L^{2}(\Omega):\operatorname{d}^{k}v\in L^{2}(\Omega)\}, where dk\operatorname{d}^{k} denotes the exterior derivative on kk-forms (d0=grad\operatorname{d}^{0}=\operatorname{grad}, d1=curl\operatorname{d}^{1}=\operatorname{curl}, d2=div\operatorname{d}^{2}=\operatorname{div}, and d3=0\operatorname{d}^{3}=0). The trace operator on theses spaces trΓk:H(dk,Ω)H1/2(Γ)\operatorname{tr}^{k}_{\Gamma}:H(\operatorname{d}^{k},\Omega)\to H^{-1/2}(\Gamma) is defined by

(2.1) trΓkv{v|Γif k=0,ΠΓv|Γif k=1,𝐧v|Γif k=2,\operatorname{tr}^{k}_{\Gamma}v\coloneqq\begin{cases}v|_{\Gamma}&\text{if }k=0,\\ \Pi_{\Gamma}v|_{\Gamma}&\text{if }k=1,\\ \mathbf{n}\cdot v|_{\Gamma}&\text{if }k=2,\end{cases}

where ΠΓ\Pi_{\Gamma} denotes the projection of a vector onto the tangent space of Γ\Gamma. The trace operator is continuous, but not surjective if k<2k<2 [12].

On a shape-regular, conforming simplicial mesh of Ω\Omega, denoted by 𝒯h\mathcal{T}_{h}, we consider well-known finite element subcomplexes of the L2L^{2} de Rham complex. For concreteness, we will restrict our discussion to the case d=3d=3:

CGp{\mathrm{CG}_{p}}Nedp1{\mathrm{Ned}^{1}_{p}}RTp{\mathrm{RT}_{p}}DGp1{\mathrm{DG}_{p-1}}H(grad){H(\operatorname{grad})}H(curl){H(\operatorname{curl})}H(div){H(\operatorname{div})}L2{L^{2}}CGp{\mathrm{CG}_{p}}Nedp12{\mathrm{Ned}^{2}_{p-1}}BDMp2{\mathrm{BDM}_{p-2}}DGp3{\mathrm{DG}_{p-3}}grad\scriptstyle{\operatorname{grad}}curl\scriptstyle{\operatorname{curl}}div\scriptstyle{\operatorname{div}}grad\scriptstyle{\operatorname{grad}}curl\scriptstyle{\operatorname{curl}}div\scriptstyle{\operatorname{div}}grad\scriptstyle{\operatorname{grad}}curl\scriptstyle{\operatorname{curl}}div\scriptstyle{\operatorname{div}}
Figure 2. The L2L^{2} de Rham complex in three dimensions (middle), and the finite element subcomplexes of the first (above) and second (below) kinds.

The top and bottom row of Figure 2 correspond to the subcomplexes of first and second kinds, respectively. In particular, CGp\mathrm{CG}_{p} and DGp\mathrm{DG}_{p} denote the usual spaces of continuous and discontinuous piecewise polynomials, Nedp1\mathrm{Ned}^{1}_{p} [40] and Nedp2\mathrm{Ned}^{2}_{p} [41] correspond to the H(curl)H(\operatorname{curl})-conforming spaces of the first and second kinds, and RTp\mathrm{RT}_{p} [45, 40] and BDMp\mathrm{BDM}_{p} [13, 41] correspond to the H(div)H(\operatorname{div})-conforming spaces of the first and second kinds. One can also form a hybrid complex by interchanging kinds as in Figure 3:

CGp{\mathrm{CG}_{p}}Nedp1{\mathrm{Ned}^{1}_{p}}BDMp1{\mathrm{BDM}_{p-1}}DGp2{\mathrm{DG}_{p-2}}CGp{\mathrm{CG}_{p}}Nedp12{\mathrm{Ned}^{2}_{p-1}}RTp1{\mathrm{RT}_{p-1}}DGp1{\mathrm{DG}_{p-1}}grad\scriptstyle{\operatorname{grad}}curl\scriptstyle{\operatorname{curl}}div\scriptstyle{\operatorname{div}}grad\scriptstyle{\operatorname{grad}}curl\scriptstyle{\operatorname{curl}}div\scriptstyle{\operatorname{div}}
Figure 3. Hybrid subcomplexes of the L2L^{2} de Rham complex.

To unify our discussion, we employ the same notation Xpk(T^)X^{k}_{p}(\hat{T}) to denote the discrete spaces of kk-forms of the first or second kind on a reference equilateral simplex T^\hat{T}. Here, pp denotes the degree of the polynomial spaces. In particular, for the first kind, we have

(2.2) Xpk(T^){CGp(T^)Pp(T^)if k=0,Nedp1(T^)Pp1(T^)3+Pp1(T^)3×𝐱if k=1,RTp(T^)Pp1(T^)3+Pp1(T^)𝐱if k=2,DGp(T^)Pp(T^)if k=3,X^{k}_{p}(\hat{T})\coloneqq\left\{\begin{array}[]{rlll}\mathrm{CG}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p}(\hat{T})&\text{if }k=0,\\ \mathrm{Ned}^{1}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p-1}(\hat{T})^{3}+\mathrm{P}_{p-1}(\hat{T})^{3}\times\mathbf{x}&\text{if }k=1,\\ \mathrm{RT}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p-1}(\hat{T})^{3}+\mathrm{P}_{p-1}(\hat{T})\mathbf{x}&\text{if }k=2,\\ \mathrm{DG}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p}(\hat{T})&\text{if }k=3,\end{array}\right.

while for the second kind we have

(2.3) Xpk(T^){CGp(T^)Pp(T^)if k=0,Nedp2(T^)Pp(T^)3if k=1,BDMp(T^)Pp(T^)3if k=2,DGp(T^)Pp(T^)if k=3.X^{k}_{p}(\hat{T})\coloneqq\left\{\begin{array}[]{rlll}\mathrm{CG}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p}(\hat{T})&\text{if }k=0,\\ \mathrm{Ned}^{2}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p}(\hat{T})^{3}&\text{if }k=1,\\ \mathrm{BDM}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p}(\hat{T})^{3}&\text{if }k=2,\\ \mathrm{DG}_{p}(\hat{T})&\coloneqq&\mathrm{P}_{p}(\hat{T})&\text{if }k=3.\end{array}\right.

We define the global finite element space Xpk(𝒯h)X^{k}_{p}(\mathcal{T}_{h}) in terms of the space on the reference cell Xpk(T^)X^{k}_{p}(\hat{T}) via the standard pullback:

(2.4) Xpk(𝒯h){vH(dk,Ω):T𝒯hv^Xpk(T^) s.t.v|T=Tk(v^)},X^{k}_{p}(\mathcal{T}_{h})\coloneqq\{v\in H(\operatorname{d}^{k},\Omega):\operatorname{\forall}T\in\mathcal{T}_{h}\ \operatorname{\exists}\hat{v}\in X^{k}_{p}(\hat{T})\text{ s.t.}\left.v\right|_{T}=\mathcal{F}^{k}_{T}(\hat{v})\},

where for a cell T𝒯hT\in\mathcal{T}_{h}, the pullback Tk:H(dk,T^)H(dk,T)\mathcal{F}^{k}_{T}:H(\operatorname{d}^{k},\hat{T})\to H(\operatorname{d}^{k},T) is defined by:

(2.5) Tk(v^){v^FT1if k=0,JTv^FT1if k=1,(detJT)1JTv^FT1if k=2,(detJT)1v^FT1if k=3,\mathcal{F}^{k}_{T}(\hat{v})\coloneqq\begin{cases}\hat{v}\circ F_{T}^{-1}&\text{if }k=0,\\ J_{T}^{-\top}\hat{v}\circ F_{T}^{-1}&\text{if }k=1,\\ (\det J_{T})^{-1}J_{T}\hat{v}\circ F_{T}^{-1}&\text{if }k=2,\\ (\det J_{T})^{-1}\hat{v}\circ F_{T}^{-1}&\text{if }k=3,\end{cases}

where FT:T^TF_{T}:\hat{T}\to T is a fixed bijective mapping from the reference cell T^\hat{T} to the physical cell TT and JTJ_{T} denotes its Jacobian matrix. Denoting by Δl(𝒯h)\Delta_{l}(\mathcal{T}_{h}) the set of ll-dimensional subsimplices of 𝒯h\mathcal{T}_{h}, we note that a discrete kk-form vXpk(𝒯h)v\in X^{k}_{p}(\mathcal{T}_{h}) and any subset SΔl(𝒯h)S\subset\Delta_{l}(\mathcal{T}_{h}) with l=k:d1l=k:d-1, the trace trSkv\operatorname{tr}^{k}_{S}v, given by formula 2.1 with Γ\Gamma replaced by SS is well-defined (see e.g. [6, Lemma 5.1]). We also define the trivial trace trSk\operatorname{tr}^{k}_{S} for a dd-dimensional subset S𝒯hS\subset\mathcal{T}_{h} by trSkv=v|S\operatorname{tr}^{k}_{S}v=v|_{S}.

With these spaces, the discrete weak formulation of problems 1.1, 1.2 and 1.3 is to find uXp,Dk(𝒯h)u\in X^{k}_{p,D}(\mathcal{T}_{h}) such that

(2.6) ak(u,v)(u,βv)+(dku,αdkv)=F(v)vXp,Dk(𝒯h),a^{k}(u,v)\coloneqq(u,\beta v)+(\operatorname{d}^{k}u,\alpha\operatorname{d}^{k}v)=F(v)\qquad\operatorname{\forall}v\in X^{k}_{p,D}(\mathcal{T}_{h}),

where Xp,Dk(𝒯h)X^{k}_{p,D}(\mathcal{T}_{h}) is the subset of Xpk(𝒯h)X^{k}_{p}(\mathcal{T}_{h}) with zero trace on ΓD\Gamma_{D}:

(2.7) Xp,Dk(𝒯h){uXpk(𝒯h):trΓDku=0}.X^{k}_{p,D}(\mathcal{T}_{h})\coloneqq\{u\in X^{k}_{p}(\mathcal{T}_{h}):\operatorname{tr}_{\Gamma_{D}}^{k}u=0\}.

For standard bases for Xp,Dk(𝒯h)X^{k}_{p,D}(\mathcal{T}_{h}), the matrix corresponding to 2.6 will have dense 𝒪(pd)×𝒪(pd)\mathcal{O}(p^{d})\times\mathcal{O}(p^{d}) blocks.

3. Orthogonality-promoting discretizations

We wish to construct bases for Xpk(𝒯h)X^{k}_{p}(\mathcal{T}_{h}) that enable fast solvers for problem 2.6 at high order. In particular, we wish to promote orthogonality of the basis in the L2L^{2} and H(dk)H(\operatorname{d}^{k}) inner products. However, constructing such bases would involve the solution of global eigenproblems, which is not appealing. Instead, we devise bases so that the interior basis functions are orthogonal in these inner products on the reference cell. This will ensure that the interior-interior block of the mass and stiffness matrices associated with the L2(T^)L^{2}(\hat{T}) and H(dk,T^)H(\operatorname{d}^{k},\hat{T}) inner products are diagonal on the reference cell.

To form bases with this interior orthogonality property, we follow the definition of interpolation degrees of freedom from Demkowicz et al. [18]. Demkowicz et al. employed these degrees of freedom to define interpolation operators with commuting diagram properties; here, we will use them to define finite elements, through the Ciarlet dual basis construction [17]. In particular, given a set of degrees of freedom {j}\{\ell_{j}\}, the basis {ϕj}\{\phi_{j}\} is constructed so that j(ϕi)=δij\ell_{j}(\phi_{i})=\delta_{ij}.

For the remainder of the manuscript, we select one of the discrete subcomplexes in Figures 2 and 3 and drop the subscript “pp” so that the complexes read

(3.1) X0(T^){X^{0}(\hat{T})}X1(T^){X^{1}(\hat{T})}{\cdots}Xd(T^){X^{d}(\hat{T})}0.{0.}d0\scriptstyle{\operatorname{d}^{0}}d1\scriptstyle{\operatorname{d}^{1}}dd1\scriptstyle{\operatorname{d}^{d-1}}

In particular, Xk(T^)X^{k}(\hat{T}) may be a first or second kind space provided that 3.1 is exact. Although we will first proceed in this abstract setting, the degrees of freedom for each of the spaces in 2.2 and 2.3 and the corresponding spaces in 2D are listed explicitly in section 3.4.

3.1. Degrees of freedom with commuting diagram properties

Let k0:d1k\in 0:d-1 be fixed. The bubble spaces on a cell subentity SΔl(T^)S\in\Delta_{l}(\hat{T}) with lk+1:dl\in k+1:d, which play a key role in the degrees of freedom in [18], are defined by

(3.2) X̊k(S){trSkv:vXk(T^) and trSkv=0}.\mathring{X}^{k}(S)\coloneqq\left\{\operatorname{tr}_{S}^{k}v:v\in X^{k}(\hat{T})\text{ and }\operatorname{tr}_{\partial S}^{k}v=0\right\}.

We will later construct bases for these bubble spaces so as to promote sparsity. We also define differential operators on the trace subspace trSkXk(T^)\operatorname{tr}_{S}^{k}X^{k}(\hat{T}) for SΔl(T^)S\in\Delta_{l}(\hat{T}) with lk+1:d1l\in k+1:d-1, as follows:

(3.3) dS0v\displaystyle\operatorname{d}_{S}^{0}v gradSvanddS1vcurlSv,\displaystyle\coloneqq\operatorname{grad}_{S}v\quad\text{and}\quad\operatorname{d}_{S}^{1}v\coloneqq\operatorname{curl}_{S}v,

where gradS\operatorname{grad}_{S} denotes the tangential surface gradient taking values in l\mathbb{R}^{l} and curlS\operatorname{curl}_{S} is the surface curl/rot operator taking values in \mathbb{R}. Crucially, these operators are defined so that

(3.4) dSktrSkv\displaystyle\operatorname{d}_{S}^{k}\operatorname{tr}_{S}^{k}v =trSk+1dkv\displaystyle=\operatorname{tr}_{S}^{k+1}\operatorname{d}^{k}v\qquad vXk(T^).\displaystyle\operatorname{\forall}v\in X^{k}(\hat{T}).

The commuting interpolation procedure given in [18, 53] is based on the following linear functionals

(3.5a) (q,trSkv)S,\displaystyle(q,\operatorname{tr}^{k}_{S}v)_{S},\qquad SΔk(T^),\displaystyle\operatorname{\forall}S\in\Delta_{k}(\hat{T}),\ qtrSkXk(T^),\displaystyle\operatorname{\forall}q\in\operatorname{tr}^{k}_{S}X^{k}(\hat{T}),
(3.5b) (dSkq,dSktrSkv)S\displaystyle(\operatorname{d}_{S}^{k}q,\operatorname{d}^{k}_{S}\operatorname{tr}^{k}_{S}v)_{S}\qquad Sl=k+1dΔl(T^),\displaystyle\operatorname{\forall}S\in\bigcup_{l=k+1}^{d}\Delta_{l}(\hat{T}),\ qX̊k(S),\displaystyle\operatorname{\forall}q\in\mathring{X}^{k}(S),
(3.5c) (dSk1q,trSkv)S\displaystyle(\operatorname{d}_{S}^{k-1}q,\operatorname{tr}^{k}_{S}v)_{S}\qquad Sl=k+1dΔl(T^),\displaystyle\operatorname{\forall}S\in\bigcup_{l=k+1}^{d}\Delta_{l}(\hat{T}),\ qX̊k1(S),\displaystyle\operatorname{\forall}q\in\mathring{X}^{k-1}(S),

which we will use to construct a set of degrees of freedom by fixing bases for each of the subspaces above.

3.2. Local space decomposition

We observe that in the linear functionals 3.5b and 3.5c, only dSkq\operatorname{d}^{k}_{S}q and dSk1q\operatorname{d}^{k-1}_{S}q appear, and so the construction of a unisolvent set of degrees of freedom requires finding minimal sets {q}\{q\} such that {dSkq}\{\operatorname{d}^{k}_{S}q\} forms a basis for dSkX̊k(S)\operatorname{d}^{k}_{S}\mathring{X}^{k}(S). Since dSk\operatorname{d}^{k}_{S} may have a nontrivial kernel, it will be convenient to characterize this kernel. This characterization can be readily achieved by considering the bubble subcomplex for SΔl(T^)S\in\Delta_{l}(\hat{T}) with lk+1:dl\in k+1:d:

(3.6) X̊k1(S){\mathring{X}^{k-1}(S)}X̊k(S){\mathring{X}^{k}(S)}dSkX̊k(S){\operatorname{d}^{k}_{S}\mathring{X}^{k}(S)}0.{0.}dSk1\scriptstyle{\operatorname{d}^{k-1}_{S}}dSk\scriptstyle{\operatorname{d}^{k}_{S}}

We define the subspaces

(3.7a) X̊k,I(S)\displaystyle\mathring{X}^{k,\mathrm{I}}(S) {vX̊k(S):(v,w)S=0wker(dSk;X̊k)},\displaystyle\coloneqq\left\{v\in\mathring{X}^{k}(S):(v,w)_{S}=0\ \operatorname{\forall}w\in\ker(\operatorname{d}_{S}^{k};\mathring{X}^{k})\right\},
(3.7b) X̊k,II(S)\displaystyle\mathring{X}^{k,\mathrm{II}}(S) ker(dSk;X̊k),\displaystyle\coloneqq\ker(\operatorname{d}_{S}^{k};\mathring{X}^{k}),

Since the complex 3.6 is exact thanks to Lemma A.1, the kernel of dSk\operatorname{d}_{S}^{k} acting on X̊k(S)\mathring{X}^{k}(S) is simply dSk1X̊k1(S)\operatorname{d}^{k-1}_{S}\mathring{X}^{k-1}(S) and so X̊k,II(S)=dSk1X̊k1(S)\mathring{X}^{k,\mathrm{II}}(S)=\operatorname{d}^{k-1}_{S}\mathring{X}^{k-1}(S). In particular, we have the following L2(S)L^{2}(S)-orthogonal decomposition:

(3.8) X̊k(S)=X̊k,I(S)X̊k,II(S)=X̊k,I(S)dSk1X̊k1,I(S).\mathring{X}^{k}(S)=\mathring{X}^{k,\mathrm{I}}(S)\oplus\mathring{X}^{k,\mathrm{II}}(S)=\mathring{X}^{k,\mathrm{I}}(S)\oplus\operatorname{d}_{S}^{k-1}\mathring{X}^{k-1,\mathrm{I}}(S).

Therefore, we may select the functions in 3.5b and 3.5c from X̊k,I(S)\mathring{X}^{k,\mathrm{I}}(S) and X̊k1,I(S)\mathring{X}^{k-1,\mathrm{I}}(S), respectively.

We now perform a similar decomposition of the trace space trSkXk(T^)\operatorname{tr}^{k}_{S}X^{k}(\hat{T}), SΔk(T^)S\in\Delta_{k}(\hat{T}), to select the functions in 3.5a. We adopt the convention that for EΔ1(T^)E\in\Delta_{1}(\hat{T}) and vX1(T^)v\in X^{1}(\hat{T}), the projection onto the tangent space of EE is scalar-valued, and thus trSkXk(T^)\operatorname{tr}^{k}_{S}X^{k}(\hat{T}), SΔk(K^)S\in\Delta_{k}(\hat{K}), is scalar-valued for k0:d1k\in 0:d-1. Thanks to Lemma A.2, we have the following L2(S)L^{2}(S)-orthogonal decomposition of the trace space trSkXk(T^)\operatorname{tr}^{k}_{S}X^{k}(\hat{T}):

(3.9) trSkXk(T^)=trSkWk(T^)dSk1X̊k1,I(S)=dSk1X̊k1,I(S)\operatorname{tr}^{k}_{S}X^{k}(\hat{T})=\operatorname{tr}^{k}_{S}W^{k}(\hat{T})\oplus\operatorname{d}_{S}^{k-1}\mathring{X}^{k-1,\mathrm{I}}(S)=\mathbb{R}\oplus\operatorname{d}_{S}^{k-1}\mathring{X}^{k-1,\mathrm{I}}(S)

where Wk(T^)W^{k}(\hat{T}) denotes the space of Whitney forms, defined as the lowest-order finite element space of the first kind; e.g., if d=3d=3, then

(3.10) Wk(T^){CG1(T^)if k=0,Ned11(T^)if k=1,RT1(T^)if k=2,DG0(T^)if k=3.W^{k}(\hat{T})\coloneqq\begin{cases}\mathrm{CG}_{1}(\hat{T})&\text{if }k=0,\\ \mathrm{Ned}^{1}_{1}(\hat{T})&\text{if }k=1,\\ \mathrm{RT}_{1}(\hat{T})&\text{if }k=2,\\ \mathrm{DG}_{0}(\hat{T})&\text{if }k=3.\end{cases}

In contrast to [18], we further decompose 3.5a using 3.9 by choosing qq in 3.5a to be the constant 1 and of the form dSk1q~\operatorname{d}_{S}^{k-1}\tilde{q}, where q~X̊pk1,I(S)\tilde{q}\in\mathring{X}^{k-1,\mathrm{I}}_{p}(S). As we will see in Lemma 4.1, this choice ensures that the canonical basis for the Whitney forms is contained in the high-order basis. We will achieve an interior orthogonality property through a careful and novel choice of basis for the bubble spaces X̊k,I(S)\mathring{X}^{k,\mathrm{I}}(S).

3.3. Abstract construction of orthogonality-promoting basis

We define the finite elements via a Ciarlet triple (T^,Xk(T^),)(\hat{T},X^{k}(\hat{T}),\mathcal{L}), where the set ={jk}\mathcal{L}=\{\ell_{j}^{k}\} is a basis for the dual space (Xk(T^))(X^{k}(\hat{T}))^{*}, with degrees of freedom jk\ell_{j}^{k} that map a kk-form vXk(T^)v\in X^{k}(\hat{T}) to a real number jk(v)\ell_{j}^{k}(v). Let NSkdimX̊k,I(S)N^{k}_{S}\coloneqq\dim\mathring{X}^{k,\mathrm{I}}(S) denote the dimension of the type-I\mathrm{I} bubble spaces for SΔl(T^)S\in\Delta_{l}(\hat{T}), lk+1:dl\in k+1:d. For notational convenience, we also define NSkdimtrSkWk(T^)=1N^{k}_{S}\coloneqq\dim\operatorname{tr}^{k}_{S}W^{k}(\hat{T})=1 for SΔk(T^)S\in\Delta_{k}(\hat{T}). Thanks to 3.8 and 3.9, there holds

(3.11a) dimX̊k(S)\displaystyle\dim\mathring{X}^{k}(S) =NSk+NSk1\displaystyle=N^{k}_{S}+N^{k-1}_{S}\qquad Sl=k+1dΔl(T^),\displaystyle\operatorname{\forall}S\in\bigcup_{l=k+1}^{d}\Delta_{l}(\hat{T}),
(3.11b) dimtrSkXk(T^)\displaystyle\dim\operatorname{tr}_{S}^{k}X^{k}(\hat{T}) =1+NSk1\displaystyle=1+N^{k-1}_{S}\qquad SΔk(T^).\displaystyle\operatorname{\forall}S\in\Delta_{k}(\hat{T}).

For k0:d1k\in 0:d-1 and lk+1:dl\in k+1:d and each subsimplex SΔl(T^)S\in\Delta_{l}(\hat{T}), we compute an eigenbasis of the bubbles {ψS,jk:j1:NSk}X̊k,I(S)\{\psi_{S,j}^{k}:j\in 1:N^{k}_{S}\}\subseteq\mathring{X}^{k,\mathrm{I}}(S) that solves the following eigenvalue problem:

(3.12) (dSkψS,jk,dSkψS,ik)S=δijand(ψS,jk,ψS,ik)S=λS,jδiji,j1:NSk.(\operatorname{d}_{S}^{k}\psi_{S,j}^{k},\operatorname{d}_{S}^{k}\psi_{S,i}^{k})_{S}=\delta_{ij}\quad\text{and}\quad(\psi_{S,j}^{k},\psi_{S,i}^{k})_{S}=\lambda_{S,j}\delta_{ij}\qquad\operatorname{\forall}\leavevmode\nobreak\ i,j\in 1:N^{k}_{S}.

The eigenbases {ψS,jk}\{\psi_{S,j}^{k}\} are numerically computed offline and only once for each relevant reference subsimplex. Then, by construction, the set {ψS,jk}\{\psi_{S,j}^{k}\} forms a basis for X̊k,I(S)\mathring{X}^{k,\mathrm{I}}(S), and the set {dSkψS,jk}\{\operatorname{d}_{S}^{k}\psi_{S,j}^{k}\} forms a basis for dSkX̊k(S)\operatorname{d}_{S}^{k}\mathring{X}^{k}(S).

Here, we choose T^\hat{T} to be an equilateral simplex so that only one facet, edge, etc. needs to be computed, and the rest are constructed by the appropriate pullback. We compute the solution to 3.12 by first solving the following augmented eigenvalue problem on the whole bubble space X̊k\mathring{X}^{k}:

(3.13) (dSkψ^S,jk,dSkψ^S,ik)S=μS,jδijand(ψ^S,jk,ψ^S,ik)S=δiji,j1:dimX̊k.(\operatorname{d}_{S}^{k}\hat{\psi}_{S,j}^{k},\operatorname{d}_{S}^{k}\hat{\psi}_{S,i}^{k})_{S}=\mu_{S,j}\delta_{ij}\quad\text{and}\quad(\hat{\psi}_{S,j}^{k},\hat{\psi}_{S,i}^{k})_{S}=\delta_{ij}\qquad\operatorname{\forall}\leavevmode\nobreak\ i,j\in 1:\dim\mathring{X}^{k}.

Then, we renormalize the subset of eigenfunctions for which μS,j0\mu_{S,j}\neq 0 to form the solution to 3.12.

Our degrees of freedom for the discrete spaces in the L2L^{2}-de Rham complex Xk(T^)X^{k}(\hat{T}) are defined as follows. On each subsimplex we prescribe two types of degrees of freedom if k>0k>0. On each kk-simplex SΔk(T^)S\in\Delta_{k}(\hat{T}), we augment the degrees of freedom for Whitney forms 3.14a with integral moments against the exterior derivative of the basis for the bubble space of (k1)(k-1)-forms X̊k1(S)\mathring{X}^{k-1}(S). On the remaining sub-simplices of dimension lk+1:dl\in k+1:d, SΔl(T^)S\in\Delta_{l}(\hat{T}), we prescribe integral moments of dSktrSkv\operatorname{d}^{k}_{S}\operatorname{tr}^{k}_{S}v against the eigenbasis for dSkX̊k\operatorname{d}^{k}_{S}\mathring{X}^{k}, and integral moments of trSkv\operatorname{tr}^{k}_{S}v against the eigenbasis for dSk1X̊k1(S)\operatorname{d}^{k-1}_{S}\mathring{X}^{k-1}(S):

(3.14a) S,1k,I(v)\displaystyle\ell^{k,\mathrm{I}}_{S,1}(v) (1,trSkv)S,\displaystyle\coloneqq(1,\operatorname{tr}^{k}_{S}v)_{S},\qquad SΔk(T^),\displaystyle\operatorname{\forall}S\in\Delta_{k}(\hat{T}),
(3.14b) S,jk,I(v)\displaystyle\ell^{k,\mathrm{I}}_{S,j}(v) (dSkψS,jk,dSktrSkv)S\displaystyle\coloneqq(\operatorname{d}_{S}^{k}\psi_{S,j}^{k},\operatorname{d}_{S}^{k}\operatorname{tr}^{k}_{S}v)_{S}\qquad Sl=k+1dΔl(T^),j1:NSk,\displaystyle\operatorname{\forall}S\in\bigcup_{l=k+1}^{d}\Delta_{l}(\hat{T}),\ \operatorname{\forall}j\in 1:N^{k}_{S},
(3.14c) S,jk,II(v)\displaystyle\ell^{k,\mathrm{II}}_{S,j}(v) (dSk1ψS,jk1,trSkv)S\displaystyle\coloneqq(\operatorname{d}_{S}^{k-1}\psi_{S,j}^{k-1},\operatorname{tr}^{k}_{S}v)_{S}\qquad Sl=kdΔl(T^),j1:NSk1.\displaystyle\operatorname{\forall}S\in\bigcup_{l=k}^{d}\Delta_{l}(\hat{T}),\ \operatorname{\forall}j\in 1:N^{k-1}_{S}.

We collect the degrees of freedom into a single set

(3.15) :=l=kdSΔl(T^){S,jk,I:j1:NSk}{S,jk,II:j1:NSk1}.\mathcal{L}:=\bigcup_{l=k}^{d}\bigcup_{S\in\Delta_{l}(\hat{T})}\{\ell^{k,\mathrm{I}}_{S,j}:j\in 1:N^{k}_{S}\}\cup\{\ell^{k,\mathrm{II}}_{S,j}:j\in 1:N^{k-1}_{S}\}.

The unisolvence of the degrees of freedom \mathcal{L} follows from standard arguments.

Lemma 3.1.

The degrees of freedom 3.15 are unisolvent on Xk(T^)X^{k}(\hat{T}).

Proof.

We first show that dim=dimXk(T^)\dim\mathcal{L}=\dim X^{k}(\hat{T}). Note that

dim\displaystyle\dim\mathcal{L} =SΔk(T^)(1+NSk1)+l=k+1dSΔl(T^)(NSk+NSk1)\displaystyle=\sum_{S\in\Delta_{k}(\hat{T})}\left(1+N_{S}^{k-1}\right)+\sum_{l=k+1}^{d}\sum_{S\in\Delta_{l}(\hat{T})}\left(N_{S}^{k}+N_{S}^{k-1}\right)
=SΔk(T^)dimtrSkXk(T^)+l=k+1dSΔl(T^)dimX̊k(S),\displaystyle=\sum_{S\in\Delta_{k}(\hat{T})}\dim\operatorname{tr}_{S}^{k}X^{k}(\hat{T})+\sum_{l=k+1}^{d}\sum_{S\in\Delta_{l}(\hat{T})}\dim\mathring{X}^{k}(S),

where we used 3.11. Thanks to [5, Theorem 5.5 & Lemma 5.6], we conclude that dim=dimXk(T^)\dim\mathcal{L}=\dim X^{k}(\hat{T}).

Now suppose that vXk(T^)v\in X^{k}(\hat{T}) satisfies (v)=0\ell(v)=0 for all \ell\in\mathcal{L}. Since 3.14a vanishes, the decomposition 3.9 shows that for all SΔk(T^)S\in\Delta_{k}(\hat{T}), trSkv=dSk1wS\operatorname{tr}_{S}^{k}v=\operatorname{d}_{S}^{k-1}w_{S} for some wSX̊k1,I(S)w_{S}\in\mathring{X}^{k-1,\mathrm{I}}(S). However, the degrees of freedom 3.14c vanish, and so trSkv0\operatorname{tr}_{S}^{k}v\equiv 0.

We may now proceed inductively on lk+1:dl\in k+1:d. For SΔl(T^)S\in\Delta_{l}(\hat{T}) with lk+1:dl\in k+1:d, trSkvX̊k(S)\operatorname{tr}_{S}^{k}v\in\mathring{X}^{k}(S), and so dSktrSkv0\operatorname{d}_{S}^{k}\operatorname{tr}_{S}^{k}v\equiv 0 by 3.14b. By the exactness of 3.6, trSkv=dSk1wS\operatorname{tr}_{S}^{k}v=\operatorname{d}_{S}^{k-1}w_{S} for some wSX̊k1,I(S)w_{S}\in\mathring{X}^{k-1,\mathrm{I}}(S). The degrees of freedom 3.14c then give wS0w_{S}\equiv 0. Since this holds with S=T^S=\hat{T}, v0v\equiv 0. The unisolvence of \mathcal{L} now follows. ∎

3.4. Concrete definition of orthogonality-promoting degrees of freedom

For the sake of readability, we have simplified our notation slightly by dropping the kk-form superscript and explicitly writing out the trace operators.

3.4.1. Degrees of freedom for H(grad)H(\operatorname{grad})

We discretize H(grad,T^)H(\operatorname{grad},\hat{T}) with the space CGp(T^)\mathrm{CG}_{p}(\hat{T}). For each subsimplex SΔl(T^)S\in\Delta_{l}(\hat{T}) with l1:dl\in 1:d, we compute the eigenbasis {ψS,j}\{\psi_{S,j}\} for X̊0,I(S)=X̊0(S)\mathring{X}^{0,\mathrm{I}}(S)=\mathring{X}^{0}(S) satisfying

(3.16) (gradSψS,j,gradSψS,i)S=δij,(ψS,j,ψS,i)S=λjδij,(\operatorname{grad}_{S}\psi_{S,j},\operatorname{grad}_{S}\psi_{S,i})_{S}=\delta_{ij},\quad(\psi_{S,j},\psi_{S,i})_{S}=\lambda_{j}\delta_{ij},

where we recall that gradS\operatorname{grad}_{S} is the tangential gradient on SS. The degrees of freedom for X0(T^)X^{0}(\hat{T}) in 3.15 are point evaluations at vertices and integral moments of surface gradients on each higher-dimensional subsimplex:

(3.17a) V(v)\displaystyle\ell_{V}(v) =v(V)\displaystyle=v(V)\qquad VΔ0(T^),\displaystyle\operatorname{\forall}V\in\Delta_{0}(\hat{T}),
(3.17b) S,j(v)\displaystyle\ell_{S,j}(v) =(gradSψS,j,gradSv)S\displaystyle=(\operatorname{grad}_{S}\psi_{S,j},\operatorname{grad}_{S}v)_{S}\qquad Sl=1dΔl(T^).\displaystyle\operatorname{\forall}S\in\bigcup_{l=1}^{d}\Delta_{l}(\hat{T}).

3.4.2. Degrees of freedom for H(curl)H(\operatorname{curl})

We discretize H(curl,T^)H(\operatorname{curl},\hat{T}) with X1(T^)X^{1}(\hat{T}), which we recall is either Nedp1(T^)\mathrm{Ned}^{1}_{p}(\hat{T}) or Nedp2(T^)\mathrm{Ned}^{2}_{p}(\hat{T}), so that the corresponding space X0(T^)X^{0}(\hat{T}) is either CGp(T^)\mathrm{CG}_{p}(\hat{T}) or CGp+1(T^)\mathrm{CG}_{p+1}(\hat{T}), respectively. We define a basis for the dual of X1(T^)X^{1}(\hat{T}) by using the eigenbasis {ψS,j}\{\psi_{S,j}\} constructed in 3.16 for the H(grad)H(\operatorname{grad}) bubbles X̊0(S)\mathring{X}^{0}(S). For each subsimplex SΔl(T^)S\in\Delta_{l}(\hat{T}) with l2:dl\in 2:d, we compute the eigenbasis {ΨS,j}\{\Psi_{S,j}\} for X̊1,I(S)\mathring{X}^{1,\mathrm{I}}(S) satisfying

(3.18) (curlSΨS,j,curlSΨS,i)S=δij,(ΨS,j,ΨS,i)S=λjδij,(\operatorname{curl}_{S}\Psi_{S,j},\operatorname{curl}_{S}\Psi_{S,i})_{S}=\delta_{ij},\quad(\Psi_{S,j},\Psi_{S,i})_{S}=\lambda_{j}\delta_{ij},

where we recall that curlS\operatorname{curl}_{S} is the tangential curl on SS. As mentioned above, {curlSΨS,j}\{\operatorname{curl}_{S}\Psi_{S,j}\} is a basis for curlSX̊1(S)\operatorname{curl}_{S}\mathring{X}^{1}(S).

The degrees of freedom for X1(T^)X^{1}(\hat{T}) in 3.15 are tangential moments along edges, moments of curl against curls and moments against gradients on each higher-dimensional subsimplex:

(3.19a) E,1I(v)\displaystyle\ell^{\mathrm{I}}_{E,1}(v) =(1,𝐭v)E\displaystyle=(1,\mathbf{t}\cdot v)_{E}\qquad EΔ1(T^),\displaystyle\operatorname{\forall}E\in\Delta_{1}(\hat{T}),
(3.19b) S,jI(v)\displaystyle\ell^{\mathrm{I}}_{S,j}(v) =(curlSΨS,j,curlSΠSv)S\displaystyle=(\operatorname{curl}_{S}\Psi_{S,j},\operatorname{curl}_{S}\Pi_{S}v)_{S}\qquad Sl=2dΔl(T^),\displaystyle\operatorname{\forall}S\in\bigcup_{l=2}^{d}\Delta_{l}(\hat{T}),
(3.19c) S,jII(v)\displaystyle\ell^{\mathrm{II}}_{S,j}(v) =(gradSψS,j,ΠSv)S\displaystyle=(\operatorname{grad}_{S}\psi_{S,j},\Pi_{S}v)_{S}\qquad Sl=1dΔl(T^),\displaystyle\operatorname{\forall}S\in\bigcup_{l=1}^{d}\Delta_{l}(\hat{T}),

where we recall that ΠS\Pi_{S} is the projection operator onto the tangent space of SS. In particular, the only difference between Nedp1(T^)\mathrm{Ned}^{1}_{p}(\hat{T}) and Nedp2(T^)\mathrm{Ned}^{2}_{p}(\hat{T}) is in 3.19c. For Nedp1(T^)\mathrm{Ned}^{1}_{p}(\hat{T}), the eigenfunctions ψS,j\psi_{S,j} defined in 3.16 and appearing in 3.19c are degree pp, while for Nedp2(T^)\mathrm{Ned}^{2}_{p}(\hat{T}), they are degree p+1p+1.

3.4.3. Degrees of freedom for H(div)H(\operatorname{div})

We discretize H(div,T^)H(\operatorname{div},\hat{T}) with X2(T^)X^{2}(\hat{T}), which we recall is either RTp(T^)\mathrm{RT}_{p}(\hat{T}) or BDMp(T^)\mathrm{BDM}_{p}(\hat{T}). If X2(T^)=RTp(T^)X^{2}(\hat{T})=\mathrm{RT}_{p}(\hat{T}), then we can choose X1(T^)X^{1}(\hat{T}) as either Nedp1\mathrm{Ned}^{1}_{p} or Nedp2\mathrm{Ned}^{2}_{p}, while if X2(T^)=BDMp(T^)X^{2}(\hat{T})=\mathrm{BDM}_{p}(\hat{T}), we can choose Xp1(T^)X^{1}_{p}(\hat{T}) as either Nedp+11\mathrm{Ned}^{1}_{p+1} or Nedp+12\mathrm{Ned}^{2}_{p+1}. The resulting basis will depend on this particular choice, but the properties of the basis in the remainder of the manuscript are independent of this choice. For the implementation, we always choose X1(T^)X^{1}(\hat{T}) to be the corresponding space of the second kind.

We define a basis for the dual of X2(T^)X^{2}(\hat{T}) by using the eigenbasis {ΨF,j}\{\Psi_{F,j}\} constructed in 3.18 for the H(curl)H(\operatorname{curl}) bubbles X̊1,I(F)\mathring{X}^{1,\mathrm{I}}(F). On the interior of T^\hat{T}, we compute the eigenbasis {ΦC,j}\{\Phi_{C,j}\} for X̊2,I(T^)\mathring{X}^{2,\mathrm{I}}(\hat{T}) satisfying

(3.20) (divΦC,j,divΦC,i)T^=δij,(ΦC,j,ΦC,i)T^=λjδij.(\operatorname{div}\Phi_{C,j},\operatorname{div}\Phi_{C,i})_{\hat{T}}=\delta_{ij},\quad(\Phi_{C,j},\Phi_{C,i})_{\hat{T}}=\lambda_{j}\delta_{ij}.

As mentioned above, {divΨC,j}\{\operatorname{div}\Psi_{C,j}\} is a basis for divX̊2(T^)\operatorname{div}\mathring{X}^{2}(\hat{T}).

The degrees of freedom for X2(T^)X^{2}(\hat{T}) in 3.15 are normal moments on faces, moments of divergence against divergences and moments against curls:

(3.21a) F,1I(v)\displaystyle\ell^{\mathrm{I}}_{F,1}(v) =(1,𝐧v)F\displaystyle=(1,\mathbf{n}\cdot v)_{F}\qquad FΔd1(T^),\displaystyle\operatorname{\forall}F\in\Delta_{d-1}(\hat{T}),
(3.21b) F,jII(v)\displaystyle\ell^{\mathrm{II}}_{F,j}(v) =(curlFΨF,j,𝐧v)F,\displaystyle=(\operatorname{curl}_{F}\Psi_{F,j},\mathbf{n}\cdot v)_{F},\qquad FΔd1(T^),\displaystyle\operatorname{\forall}F\in\Delta_{d-1}(\hat{T}),
(3.21c) C,jI(v)\displaystyle\ell^{\mathrm{I}}_{C,j}(v) =(divΦC,j,divv)T^,\displaystyle=(\operatorname{div}\Phi_{C,j},\operatorname{div}v)_{\hat{T}},\qquad
(3.21d) C,jII(v)\displaystyle\ell^{\mathrm{II}}_{C,j}(v) =(curlΨC,j,v)T^.\displaystyle=(\operatorname{curl}\Psi_{C,j},v)_{\hat{T}}.

Note that the only difference between RTp(T^)\mathrm{RT}_{p}(\hat{T}) or BDMp(T^)\mathrm{BDM}_{p}(\hat{T}) are the eigenfunctions ΨF,j\Psi_{F,j} and ΨC,j\Psi_{C,j} defined in 3.18 appearing in 3.21b and 3.21d. For RTp(T^)\mathrm{RT}_{p}(\hat{T}), these eigenfunctions are Nedp1\mathrm{Ned}^{1}_{p} or Nedp2\mathrm{Ned}^{2}_{p} functions (or their trace), while for BDMp(T^)\mathrm{BDM}_{p}(\hat{T}), they are Nedp+11(T^)\mathrm{Ned}^{1}_{p+1}(\hat{T}) or Nedp+12(T^)\mathrm{Ned}^{2}_{p+1}(\hat{T}) functions (or their trace).

Remark 3.2.

The degrees of freedom in 3.17 and 3.19 for discretizations of H(grad)H(\operatorname{grad}) and H(curl)H(\operatorname{curl}) are well-defined for d=2d=2. To discretize H(div)H(\operatorname{div}) in 2D, which corresponds to k=1k=1, we choose X0(T^)X^{0}(\hat{T}) to be CGp\mathrm{CG}_{p} if X1(T^)=RTpX^{1}(\hat{T})=\mathrm{RT}_{p} or CGp+1\mathrm{CG}_{p+1} if X1(T^)=BDMpX^{1}(\hat{T})=\mathrm{BDM}_{p} and the operator curlF\operatorname{curl}_{F} in 3.21b is the tangential derivative operator and curl\operatorname{curl} in 3.21d corresponds to a π/2\pi/2 rotation of the gradient operator.

4. Properties of the basis on the reference cell

For k0:d1k\in 0:d-1, let

(4.1) {ϕS,jk,I,ϕS,nk,II:j=1:NSk,n=1:NSk1,SΔl(T^),lk:d}\displaystyle\left\{\phi_{S,j}^{k,\mathrm{I}},\phi_{S,n}^{k,\mathrm{II}}:j=1:N^{k}_{S},\ n=1:N^{k-1}_{S},\ S\in\Delta_{l}(\hat{T}),\ l\in k:d\right\}

be the basis dual to the degrees of freedom 3.15:

S,ik,I(ϕS,jk,I)\displaystyle\ell^{k,\mathrm{I}}_{S,i}(\phi_{S,j}^{k,\mathrm{I}}) =δij,\displaystyle=\delta_{ij},\qquad S,mk,II(ϕS,jk,I)\displaystyle\ell^{k,\mathrm{II}}_{S,m}(\phi_{S,j}^{k,\mathrm{I}}) =0,\displaystyle=0,
S,ik,I(ϕS,nk,II)\displaystyle\ell^{k,\mathrm{I}}_{S,i}(\phi_{S,n}^{k,\mathrm{II}}) =0,\displaystyle=0,\qquad S,mk,II(ϕS,nk,II)\displaystyle\ell^{k,\mathrm{II}}_{S,m}(\phi_{S,n}^{k,\mathrm{II}}) =δmn,\displaystyle=\delta_{mn},

for all i,j=1:NSki,j=1:N^{k}_{S}, m,n=1:NSk1m,n=1:N^{k-1}_{S}, and SΔl(T^)S\in\Delta_{l}(\hat{T}) with l=k:dl=k:d.

4.1. Relations to Whitney forms and eigenfunctions

The first result shows that this basis contains the canonical basis for the Whitney forms.

Lemma 4.1.

For k=0:d1k=0:d-1 and SΔk(T^)S\in\Delta_{k}(\hat{T}), ϕS,1k,I\phi_{S,1}^{k,\mathrm{I}} coincides with the Whitney form associated to SS.

Proof.

We prove the result for d=3d=3; the case d=2d=2 is similar. Let k0:d1k\in 0:d-1 and, for SΔk(T^)S\in\Delta_{k}(\hat{T}), let wSkWk(T^)w_{S}^{k}\in W^{k}(\hat{T}) denote the Whitney form associated to SS: (1,trSkwSk)=δSS(1,\operatorname{tr}_{S^{\prime}}^{k}w_{S}^{k})=\delta_{SS^{\prime}} for all SΔk(T^)S^{\prime}\in\Delta_{k}(\hat{T}).

Step 1: Type I degrees of freedom 3.14b vanish. For Sl=k+1dΔk(T^)S^{\prime}\in\cup_{l=k+1}^{d}\Delta_{k}(\hat{T}) and j1:NSkj\in 1:N_{S^{\prime}}^{k}, there holds

S,jk,I(wSk)=(dSkψS,jk,dSktrSkwSk)S\displaystyle\ell^{k,\mathrm{I}}_{S^{\prime},j}(w_{S}^{k})=(\operatorname{d}_{S^{\prime}}^{k}\psi_{S^{\prime},j}^{k},\operatorname{d}_{S^{\prime}}^{k}\operatorname{tr}^{k}_{S^{\prime}}w_{S}^{k})_{S^{\prime}} ={(ψS,j0,divSgradStrSkwS0)Sif k=0,(ψS,j1,curlScurlStrSkwS1)if k=1,(ψS,j2,gradSdivStrSkwS2)Sif k=2,\displaystyle=\begin{cases}-(\psi_{S^{\prime},j}^{0},\operatorname{div}_{S^{\prime}}\operatorname{grad}_{S^{\prime}}\operatorname{tr}^{k}_{S^{\prime}}w_{S}^{0})_{S^{\prime}}&\text{if }k=0,\\ (\psi_{S^{\prime},j}^{1},\operatorname{curl}_{S^{\prime}}\operatorname{curl}_{S^{\prime}}\operatorname{tr}^{k}_{S^{\prime}}w_{S}^{1})&\text{if }k=1,\\ -(\psi_{S^{\prime},j}^{2},\operatorname{grad}_{S^{\prime}}\operatorname{div}_{S^{\prime}}\operatorname{tr}^{k}_{S^{\prime}}w_{S}^{2})_{S^{\prime}}&\text{if }k=2,\end{cases}

where divS\operatorname{div}_{S^{\prime}} denotes surface divergence. Since wSkP1(T^)w_{S}^{k}\in\mathrm{P}_{1}(\hat{T}), S,jk,I(wSk)=0\ell^{k,\mathrm{I}}_{S^{\prime},j}(w_{S}^{k})=0.

Step 2: Type II degrees of freedom 3.14c vanish. Let Sl=kdΔl(T^)S^{\prime}\in\cup_{l=k}^{d}\Delta_{l}(\hat{T}) and j1:NSk1j\in 1:N_{S^{\prime}}^{k-1}. Then, there holds

S,jk,II(wSk)=(dSk1ψS,jk1,trSkwSk)S\displaystyle\ell^{k,\mathrm{II}}_{S^{\prime},j}(w_{S}^{k})=(\operatorname{d}_{S^{\prime}}^{k-1}\psi_{S^{\prime},j}^{k-1},\operatorname{tr}^{k}_{S^{\prime}}w_{S}^{k})_{S^{\prime}} ={(ψS,j1,divStrSkwS1)if k=1,(ψS,j2,curlStrSkwS2)Sif k=2,.\displaystyle=\begin{cases}(\psi_{S^{\prime},j}^{1},\operatorname{div}_{S^{\prime}}\operatorname{tr}^{k}_{S^{\prime}}w_{S}^{1})&\text{if }k=1,\\ -(\psi_{S^{\prime},j}^{2},\operatorname{curl}_{S^{\prime}}\operatorname{tr}^{k}_{S^{\prime}}w_{S}^{2})_{S^{\prime}}&\text{if }k=2,\end{cases}.

If SΔk(T^)S^{\prime}\in\Delta_{k}(\hat{T}), then trSkwSk\operatorname{tr}_{S^{\prime}}^{k}w_{S}^{k}\in\mathbb{R}, and so S,jk,II(wSk)=0\ell^{k,\mathrm{II}}_{S^{\prime},j}(w_{S}^{k})=0.

Now suppose that SΔl(T^)S^{\prime}\in\Delta_{l}(\hat{T}) with lk+1:dl\in k+1:d. For k=1k=1, we have

trSwS1{P0(S)2+P0(S)(ΠS𝐱)if l=2,P0(S)3+P0(S)3×𝐱if l=3,\displaystyle\operatorname{tr}_{S^{\prime}}w_{S}^{1}\in\begin{cases}\mathrm{P}_{0}(S^{\prime})^{2}+\mathrm{P}_{0}(S)(\Pi_{S^{\prime}}\mathbf{x})^{\perp}&\text{if }l=2,\\ \mathrm{P}_{0}(S^{\prime})^{3}+\mathrm{P}_{0}(S)^{3}\times\mathbf{x}&\text{if }l=3,\end{cases}

and direct computation shows that trSwS1=0\operatorname{tr}_{S^{\prime}}w_{S}^{1}=0. E.g. for l=3l=3, we have

div(𝐚+𝐛×𝐱)=div(𝐛×𝐱)=𝐛curl𝐱=0𝐚,𝐛3.\displaystyle\operatorname{div}(\mathbf{a}+\mathbf{b}\times\mathbf{x})=\operatorname{div}(\mathbf{b}\times\mathbf{x})=-\mathbf{b}\cdot\operatorname{curl}\mathbf{x}=0\qquad\operatorname{\forall}\mathbf{a},\mathbf{b}\in\mathbb{R}^{3}.

Similarly, for k=2k=2, we have

curl(𝐚+b𝐱)=bcurl𝐱=0𝐚3,b.\displaystyle\operatorname{curl}(\mathbf{a}+b\mathbf{x})=b\operatorname{curl}\mathbf{x}=0\qquad\operatorname{\forall}\mathbf{a}\in\mathbb{R}^{3},\ \operatorname{\forall}b\in\mathbb{R}.

Thus, 3.14c vanish. By the unisolvence of 3.15, ϕS,1k,I=wSk\phi_{S,1}^{k,\mathrm{I}}=w_{S}^{k}. ∎

The next result shows that the traces of the type-I\mathrm{I} basis functions coincide with the associated eigenfunctions, while the traces of the type-II\mathrm{II} basis functions coincide with the exterior derivative of the (k1)(k-1)-form eigenfunctions.

Lemma 4.2.

For k0:d1k\in 0:d-1, there holds

(4.2a) trSkϕS,jk,I\displaystyle\operatorname{tr}^{k}_{S}\phi_{S,j}^{k,\mathrm{I}} =ψS,jk\displaystyle=\psi_{S,j}^{k}\qquad Sl=k+1dΔl(T^),\displaystyle\operatorname{\forall}S\in\bigcup_{l=k+1}^{d}\Delta_{l}(\hat{T}),\ j1:NSk,\displaystyle\operatorname{\forall}j\in 1:N_{S}^{k},
(4.2b) trSkϕS,nk,II\displaystyle\operatorname{tr}^{k}_{S}\phi_{S,n}^{k,\mathrm{II}} =dSk1ψS,nk1\displaystyle=\operatorname{d}_{S}^{k-1}\psi_{S,n}^{k-1}\qquad Sl=kdΔl(T^),\displaystyle\operatorname{\forall}S\in\bigcup_{l=k}^{d}\Delta_{l}(\hat{T}),\ n1:NSk1,\displaystyle\operatorname{\forall}n\in 1:N_{S}^{k-1},

where {ψS,jk}\{\psi_{S,j}^{k}\} is defined in 3.12.

Proof.

Let k0:d1k\in 0:d-1, SΔl(T^)S\in\Delta_{l}(\hat{T}) for some lk+1:dl\in k+1:d, and j1:NSkj\in 1:N_{S}^{k}. Arguing as in the proof of Lemma 3.1, we may show that trSkϕS,jk,IX̊k,I(S)\operatorname{tr}_{S}^{k}\phi_{S,j}^{k,\mathrm{I}}\in\mathring{X}^{k,\mathrm{I}}(S), and so 4.2a follows by the uniqueness of the eigendecomposition of X̊k(S)\mathring{X}^{k}(S) 3.12. Equation 4.2b follows from a similar argument. ∎

In fact, we can improve 4.2b and show that each type-II\mathrm{II} basis functions is the exterior derivative of a type-I\mathrm{I} basis function for (k1)(k-1)-forms.

Lemma 4.3.

For k1:d1k\in 1:d-1, there holds

(4.3) ϕS,nk,II=dk1ϕS,nk1,ISl=kdΔl(T^),n1:NSk1.\displaystyle\phi_{S,n}^{k,\mathrm{II}}=\operatorname{d}^{k-1}\phi_{S,n}^{k-1,\mathrm{I}}\qquad\operatorname{\forall}S\in\bigcup_{l=k}^{d}\Delta_{l}(\hat{T}),\ \operatorname{\forall}n\in 1:N_{S}^{k-1}.
Proof.

Let SΔk(T^)S^{\prime}\in\Delta_{k}(\hat{T}). If SSS\neq S^{\prime}, then trSk1ϕS,nk1,I0\operatorname{tr}^{k-1}_{S^{\prime}}\phi_{S,n}^{k-1,\mathrm{I}}\equiv 0 by the same arguments in the proof of Lemma 3.1. Moreover, if S=SS=S^{\prime}, then ϕS,nk1,IX̊k1,I\phi_{S,n}^{k-1,\mathrm{I}}\in\mathring{X}^{k-1,\mathrm{I}} and 3.9 shows that (1,dSk1trSk1ϕS,nk1,I)S=0(1,\operatorname{d}_{S}^{k-1}\operatorname{tr}^{k-1}_{S}\phi_{S,n}^{k-1,\mathrm{I}})_{S}=0, and so 3.4 gives

(1,trSkdk1ϕS,nk1,I)S=(1,dSk1trSk1ϕS,nk1,I)S=0.\displaystyle(1,\operatorname{tr}^{k}_{S}\operatorname{d}^{k-1}\phi_{S,n}^{k-1,\mathrm{I}})_{S}=(1,\operatorname{d}_{S}^{k-1}\operatorname{tr}^{k-1}_{S}\phi_{S,n}^{k-1,\mathrm{I}})_{S}=0.

Now let Sl=k+1dΔl(T^)S^{\prime}\in\cup_{l=k+1}^{d}\Delta_{l}(\hat{T}) and j1:NSkj\in 1:N^{k}_{S^{\prime}}. Equation 3.4 again gives

(dSkψS,jk,dSktrSkdk1ϕS,nk1,I)S=(dSkψS,jk,dSkdSk1trSk1ϕS,nk1,I)S=0\displaystyle(\operatorname{d}_{S^{\prime}}^{k}\psi_{S^{\prime},j}^{k},\operatorname{d}_{S^{\prime}}^{k}\operatorname{tr}_{S^{\prime}}^{k}\operatorname{d}^{k-1}\phi_{S,n}^{k-1,\mathrm{I}})_{S^{\prime}}=(\operatorname{d}_{S^{\prime}}^{k}\psi_{S^{\prime},j}^{k},\operatorname{d}_{S^{\prime}}^{k}\operatorname{d}_{S^{\prime}}^{k-1}\operatorname{tr}_{S^{\prime}}^{k-1}\phi_{S,n}^{k-1,\mathrm{I}})_{S^{\prime}}=0

by the complex property dSkdSk1=0\operatorname{d}_{S^{\prime}}^{k}\circ\operatorname{d}_{S^{\prime}}^{k-1}=0. Thus, the degrees of freedom 3.14a and 3.14b vanish.

Finally, let Sl=kdΔl(T^)S^{\prime}\in\cup_{l=k}^{d}\Delta_{l}(\hat{T}) and m1:NSk1m\in 1:N^{k-1}_{S}. By virtue of ϕS,nk1,I\phi_{S,n}^{k-1,\mathrm{I}} being a basis function of Xpk1(T^)X^{k-1}_{p}(\hat{T}) dual to the degrees of freedom \mathcal{L}, there holds

(dSk1ψS,jk1,trSkdk1ϕS,nk1,I)S=(dSk1ψS,jk1,dSk1trSk1ϕS,nk1,I)S=δSSδmn.\displaystyle(\operatorname{d}_{S^{\prime}}^{k-1}\psi_{S^{\prime},j}^{k-1},\operatorname{tr}^{k}_{S^{\prime}}\operatorname{d}^{k-1}\phi_{S,n}^{k-1,\mathrm{I}})_{S^{\prime}}=(\operatorname{d}_{S^{\prime}}^{k-1}\psi_{S^{\prime},j}^{k-1},\operatorname{d}_{S^{\prime}}^{k-1}\operatorname{tr}^{k-1}_{S^{\prime}}\phi_{S,n}^{k-1,\mathrm{I}})_{S^{\prime}}=\delta_{SS^{\prime}}\delta_{mn}.

Equation 4.3 now follows from the unisolvence of \mathcal{L} on Xk(T^)X^{k}(\hat{T}). ∎

Remark 4.4.

The degrees of freedom \mathcal{L} are hierarchical in dimension in the sense that the degrees of freedom for d=2d=2 are present on each face of the reference tetrahedron for d=3d=3. Consequently, for SΔ2(T^)S\in\Delta_{2}(\hat{T}), the trace of the 3D basis functions associated to SS, {trSϕS,jk,I,trSϕS,nk,II}\{\operatorname{tr}_{S}\phi_{S,j}^{k,\mathrm{I}},\operatorname{tr}_{S}\phi_{S,n}^{k,\mathrm{II}}\}, coincides with the 2D basis construction on SS.

4.2. Algebraic properties

We now turn to properties of the resulting mass and stiffness matrices on the reference cell. In particular, let {ϕjk}\{\phi_{j}^{k}\} denote the basis for Xk(T^)X^{k}(\hat{T}) in 4.1 and let the mass matrix M^=(M^ij)\hat{M}=(\hat{M}_{ij}) and stiffness matrix K^=(K^ij)\hat{K}=(\hat{K}_{ij}) be given by

M^ij=(ϕik,ϕjk)T^andK^ij=(dkϕik,dkϕjk)T^.\displaystyle\hat{M}_{ij}=(\phi_{i}^{k},\phi_{j}^{k})_{\hat{T}}\quad\text{and}\quad\hat{K}_{ij}=(\operatorname{d}^{k}\phi_{i}^{k},\operatorname{d}^{k}\phi_{j}^{k})_{\hat{T}}.

We partition M^\hat{M} and K^\hat{K} into the contribution from the interior and interface basis functions:

M^=[M^M^ΓM^ΓM^ΓΓ]andK^=[K^K^ΓK^ΓK^ΓΓ],\displaystyle\hat{M}=\begin{bmatrix}\hat{M}_{\mathcal{I}\mathcal{I}}&\hat{M}_{\mathcal{I}\Gamma}\\ \hat{M}_{\Gamma\mathcal{I}}&\hat{M}_{\Gamma\Gamma}\end{bmatrix}\quad\text{and}\quad\hat{K}=\begin{bmatrix}\hat{K}_{\mathcal{I}\mathcal{I}}&\hat{K}_{\mathcal{I}\Gamma}\\ \hat{K}_{\Gamma\mathcal{I}}&\hat{K}_{\Gamma\Gamma}\end{bmatrix},

where the subscript “\mathcal{I}\mathcal{I}” denotes the contribution from the interior basis functions (ϕS,jk,I,ϕS,jk,II\phi_{S,j}^{k,\mathrm{I}},\phi_{S,j}^{k,\mathrm{II}} with S=T^S=\hat{T}), “ΓΓ\Gamma\Gamma” the contribution from the remaining (interface) functions, and “Γ\mathcal{I}\Gamma” and “Γ\Gamma\mathcal{I}” the interaction between the interior and interface functions. We further partition each block into contributions from the type-I\mathrm{I} and type-II\mathrm{II} basis functions, using the superscripts “I,I\mathrm{I},\mathrm{I}”, “I,II\mathrm{I},\mathrm{II}”, etc.

Lemma 4.5.

M^\hat{M} and K^\hat{K} have the following structure:

M^=[ΛI]andK^=[I],\displaystyle\hat{M}=\left[\begin{array}[]{cc|cc}\Lambda&&\bullet&\bullet\\ &I&&\\ \hline\cr\bullet&&\bullet&\bullet\\ \bullet&&\bullet&\bullet\end{array}\right]\quad\text{and}\quad\hat{K}=\left[\begin{array}[]{cc|cc}I&&&\\ &&&\\ \hline\cr&&\bullet&\\ &&&\end{array}\right],

where Λ\Lambda is a diagonal matrix whose entries are the eigenvalues {λT,j}j=1NTk\{\lambda_{T,j}\}_{j=1}^{N_{T}^{k}} 3.12.

Proof.

First note that the degrees of freedom 3.14b with S=T^S=\hat{T} mean that (dkϕ,dkv)T^=0(\operatorname{d}^{k}\phi,\operatorname{d}^{k}v)_{\hat{T}}=0 for any interface basis function ϕ\phi and any interior bubble vX̊k(T^)v\in\mathring{X}^{k}(\hat{T}), and so K^Γ=0\hat{K}_{\mathcal{I}\Gamma}=0 and K^Γ=0\hat{K}_{\Gamma\mathcal{I}}=0. Similarly, 4.3 shows that dkϕ=0\operatorname{d}^{k}\phi=0 for any type-II\mathrm{II} basis function, and so the only nonzero blocks of K^\hat{K} are K^I,I\hat{K}_{\mathcal{I}\mathcal{I}}^{\mathrm{I},\mathrm{I}} and K^Γ,ΓI,I\hat{K}_{\Gamma,\Gamma}^{\mathrm{I},\mathrm{I}}. Thanks to 4.2a and 3.12, K^I,I=I\hat{K}_{\mathcal{I}\mathcal{I}}^{\mathrm{I},\mathrm{I}}=I.

Using 4.2 and 3.12 once again gives that M^\hat{M}_{\mathcal{I}\mathcal{I}} is diagonal. More precisely, Equations 4.2a and 3.12 mean that M^I,I=Λ\hat{M}_{\mathcal{I}\mathcal{I}}^{\mathrm{I},\mathrm{I}}=\Lambda. Additionally, 4.3 and the choice of degrees of freedom 3.14c show that (i) M^ΓII,I=0\hat{M}_{\mathcal{I}\Gamma}^{\mathrm{II},\mathrm{I}}=0 and M^ΓII,II=0\hat{M}_{\mathcal{I}\Gamma}^{\mathrm{II},\mathrm{II}}=0 and (ii) the typeII-\mathrm{II}/type-II\mathrm{II} block of the mass matrix of a kk-form is the typeI-\mathrm{I}/type-I\mathrm{I} block of the stiffness matrix of a (k1)(k-1)-form. Thus, M^II,II=I\hat{M}_{\mathcal{I}\mathcal{I}}^{\mathrm{II},\mathrm{II}}=I. ∎

4.3. Comparison to existing bases

In Figure 4 we compare the H(grad)H(\operatorname{grad}) interior basis functions of this construction on the equilateral reference simplex to the hierarchical basis proposed by Sherwin & Karniadakis [48]. The basis proposed by this work has been constructed to produce diagonal stiffness and mass interior submatrices KK_{\mathcal{I}\mathcal{I}} and MM_{\mathcal{I}\mathcal{I}}. Thus, diagonal scaling will ensure a condition number of one. On the other hand, for the hierarchical basis these condition numbers grow with pp, indicating that a point-Jacobi preconditioner is not a suitable approach for this basis. A sparse-direct method would be required to eliminate the interiors.

We also consider the number of nonzeros per row of the stiffness matrix. In our construction, the number of nonzeros per row of the interior mass and stiffness blocks is exactly one. For the hierarchical basis of [48], Beuchler & Pillwein [7] proved that the number of nonzeros per row asymptotes to a constant of approximately 8 in two dimensions, and is proven to be bounded above by 105 in three dimensions.

44 66 88 1010 1212 1616 2020 2424 10010^{0}10110^{1}10210^{2}10310^{3}p2p^{2}p4p^{4}Degree, ppκ2(diag(A)1A)\kappa_{2}(\operatorname{diag}(A)^{-1}A)Hierarchical, KK_{\mathcal{I}\mathcal{I}}Hierarchical, MM_{\mathcal{I}\mathcal{I}}This work, KK_{\mathcal{I}\mathcal{I}}This work, MM_{\mathcal{I}\mathcal{I}}
(a)
44 66 88 1010 1212 1414 10010^{0}10110^{1}10210^{2}10310^{3}p4p^{4}p5p^{5}Degree, ppκ2(diag(A)1A)\kappa_{2}(\operatorname{diag}(A)^{-1}A)Hierarchical, KK_{\mathcal{I}\mathcal{I}}Hierarchical, MM_{\mathcal{I}\mathcal{I}}This work, KK_{\mathcal{I}\mathcal{I}}This work, MM_{\mathcal{I}\mathcal{I}}
(b)
Figure 4. Condition number of the diagonally-scaled stiffness and mass interior submatrices for H(grad)H(\operatorname{grad}) on an equilateral reference simplex. The proposed basis in this work yields diagonal submatrices, while the sparsity-optimized hierarchical bases studied in [7] are less suitable for diagonal scaling.

5. Properties of the global basis functions

Continuing with the notation from the previous sections, we let XkX^{k} denote the global discrete space Xp,Dk(𝒯h)X^{k}_{p,D}(\mathcal{T}_{h}). The global basis for XkX^{k} is constructed from the local basis on Xk(T^)X^{k}(\hat{T}) as in [46]. In particular, for each element T𝒯hT\in\mathcal{T}_{h}, the bijective mapping FT:T^TF_{T}:\hat{T}\to T is chosen to preserve fixed global orientations of subsimplices, and each global basis function supported on TT is of the form Tk(ϕ^)\mathcal{F}_{T}^{k}(\hat{\phi}), where {ϕ^}\{\hat{\phi}\} is a basis for Xk(T^)X^{k}(\hat{T}) and Tk\mathcal{F}_{T}^{k} is defined in 2.5. For the remainder of the manuscript, we will assume that FTF_{T} is chosen in this way. Boundary conditions are then enforced by removing all functions associated to subsimplices SΔl(𝒯h)S\in\Delta_{l}(\mathcal{T}_{h}), lk:d1l\in k:d-1, lying on ΓD\Gamma_{D}.

5.1. Key subspaces

For k0:d1k\in 0:d-1 and each subsimplex SΔl(𝒯h)S\in\Delta_{l}(\mathcal{T}_{h}), lk:dl\in k:d, let {ϕS,jk,I}\{\phi_{S,j}^{k,\mathrm{I}}\} and {ϕS,nk,II}\{\phi_{S,n}^{k,\mathrm{II}}\} denote the corresponding global basis functions associated to SS. We collect a number of subspaces related to the basis functions. First, we separate the subspaces corresponding to the type-I\mathrm{I} basis functions into the Whitney forms, the high-order interface bubble functions, and the interior functions supported on one element:

(5.1a) Wk\displaystyle W^{k} :=span{ϕS,1k,I:SΔk(𝒯h),trΓDkϕS,1k,I=0},\displaystyle:=\operatorname{span}\left\{\phi_{S,1}^{k,\mathrm{I}}:\ S\in\Delta_{k}(\mathcal{T}_{h}),\ \operatorname{tr}_{\Gamma_{D}}^{k}\phi_{S,1}^{k,\mathrm{I}}=0\right\},
(5.1b) X̊Γk,I\displaystyle\mathring{X}^{k,\mathrm{I}}_{\Gamma} :=span{ϕS,jk,I:Sl=k+1d1Δl(𝒯h),j1:NSk,trΓDkϕS,jk,I=0},\displaystyle:=\operatorname{span}\left\{\phi_{S,j}^{k,\mathrm{I}}:S\in\bigcup_{l=k+1}^{d-1}\Delta_{l}(\mathcal{T}_{h}),\ j\in 1:N^{k}_{S},\ \operatorname{tr}_{\Gamma_{D}}^{k}\phi_{S,j}^{k,\mathrm{I}}=0\right\},
(5.1c) X̊k,I\displaystyle\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}} :=span{ϕT,jk,I:TΔd(𝒯h),j1:NTk},\displaystyle:=\operatorname{span}\left\{\phi_{T,j}^{k,\mathrm{I}}:\ \forall T\in\Delta_{d}(\mathcal{T}_{h}),\ j\in 1:N^{k}_{T}\right\},
(5.1d) Xk,I\displaystyle X^{k,\mathrm{I}} :=WkX̊Γk,IX̊k,I,\displaystyle:=W^{k}\oplus\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}},

where spaces with the “̊\mathring{\phantom{n}}” overset vanish on SΔk(𝒯h)S\in\Delta_{k}(\mathcal{T}_{h}). Similarly, we separate the type-II\mathrm{II} subspaces:

(5.2a) XΓk,II\displaystyle X^{k,\mathrm{II}}_{\Gamma} :=span{ϕS,jk,II:Sl=kd1Δl(𝒯h),j1:NSk1,trΓDkϕS,jk,II=0},\displaystyle:=\operatorname{span}\left\{\phi_{S,j}^{k,\mathrm{II}}:S\in\bigcup_{l=k}^{d-1}\Delta_{l}(\mathcal{T}_{h}),\ j\in 1:N^{k-1}_{S},\operatorname{tr}_{\Gamma_{D}}^{k}\phi_{S,j}^{k,\mathrm{II}}=0\right\},
(5.2b) X̊k,II\displaystyle\mathring{X}^{k,\mathrm{II}}_{\mathcal{I}} :=span{ϕT,jk,II:TΔd(𝒯h),j1:NTk1},\displaystyle:=\operatorname{span}\left\{\phi_{T,j}^{k,\mathrm{II}}:\forall T\in\Delta_{d}(\mathcal{T}_{h}),\ j\in 1:N^{k-1}_{T}\right\},
(5.2c) Xk,II\displaystyle X^{k,\mathrm{II}} :=XΓk,IIX̊k,II.\displaystyle:=X^{k,\mathrm{II}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{II}}_{\mathcal{I}}.

By construction, the type-I\mathrm{I} and type-II\mathrm{II} spaces decompose XkX^{k}:

(5.3) Xk=Xk,IXk,II.\displaystyle X^{k}=X^{k,\mathrm{I}}\oplus X^{k,\mathrm{II}}.

Moreover, since dkTk(ϕ)=Tk+1(dkϕ)\operatorname{d}^{k}\mathcal{F}_{T}^{k}(\phi)=\mathcal{F}_{T}^{k+1}(\operatorname{d}^{k}\phi), we also have the global analogue of Lemma 4.3 for k1:d1k\in 1:d-1:

(5.4) ϕS,nk,II=dk1ϕS,nk1,ISl=kdΔl(𝒯h),n1:NSk1.\displaystyle\phi_{S,n}^{k,\mathrm{II}}=\operatorname{d}^{k-1}\phi_{S,n}^{k-1,\mathrm{I}}\qquad\operatorname{\forall}S\in\bigcup_{l=k}^{d}\Delta_{l}(\mathcal{T}_{h}),\ \operatorname{\forall}n\in 1:N_{S}^{k-1}.

5.2. Decomposing the range and kernel of dk\operatorname{d}^{k}

We now decompose the range of kernel of dk\operatorname{d}^{k} using the subspaces defined above. The first result shows that the image of dk\operatorname{d}^{k} acting on the high-order type-I\mathrm{I} spaces is disjoint from dkWk\operatorname{d}^{k}W^{k}.

Lemma 5.1.

For k0:d1k\in 0:d-1, if uX̊Γk,IX̊k,Iu\in\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}} satisfies dkudkWk\operatorname{d}^{k}u\in\operatorname{d}^{k}W^{k}, then u0u\equiv 0. Consequently, dk(X̊Γk,IX̊k,I)dkWk={0}\operatorname{d}^{k}(\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}})\cap\operatorname{d}^{k}W^{k}=\{0\}.

Proof.

Let vX̊Γk,IX̊k,Iv\in\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}} and wWkw\in W^{k} satisfy dkv=dkw\operatorname{d}^{k}v=\operatorname{d}^{k}w. Let T𝒯hT\in\mathcal{T}_{h}, v^X̊Γk,I(T^)\hat{v}\in\mathring{X}^{k,\mathrm{I}}_{\Gamma}(\hat{T}), and w^Wk(T^)\hat{w}\in W^{k}(\hat{T}) be such that v|T=Tk(v^)v|_{T}=\mathcal{F}^{k}_{T}(\hat{v}) and w|T=Tk(w^)w|_{T}=\mathcal{F}^{k}_{T}(\hat{w}), where X̊Γk,I(T^)\mathring{X}^{k,\mathrm{I}}_{\Gamma}(\hat{T}), etc. denotes the spans of the corresponding basis functions on the reference cell. Since

Tk+1(dkw^)=dkw|T=dkv|T=Tk+1(dkv^)\displaystyle\mathcal{F}^{k+1}_{T}(\operatorname{d}^{k}\hat{w})=\operatorname{d}^{k}w|_{T}=\operatorname{d}^{k}v|_{T}=\mathcal{F}^{k+1}_{T}(\operatorname{d}^{k}\hat{v})

and Tk+1\mathcal{F}^{k+1}_{T} is invertible, there holds dkw^=dkv^\operatorname{d}^{k}\hat{w}=\operatorname{d}^{k}\hat{v}. However, for SΔl(T^)S\in\Delta_{l}(\hat{T}), lk+1:dl\in k+1:d, and j1:NSkj\in 1:N_{S}^{k}, there holds

(dSk,ψS,j,dSktrSkv^)S=(dSk,ψS,j,trSk+1dkv^)S\displaystyle(\operatorname{d}^{k}_{S},\psi_{S,j},\operatorname{d}^{k}_{S}\operatorname{tr}_{S}^{k}\hat{v})_{S}=(\operatorname{d}^{k}_{S},\psi_{S,j},\operatorname{tr}_{S}^{k+1}\operatorname{d}^{k}\hat{v})_{S} =(dSk,ψS,j,trSk+1dkw^)S\displaystyle=(\operatorname{d}^{k}_{S},\psi_{S,j},\operatorname{tr}_{S}^{k+1}\operatorname{d}^{k}\hat{w})_{S}
=(dSk,ψS,j,dSktrSkw^)S=0\displaystyle=(\operatorname{d}^{k}_{S},\psi_{S,j},\operatorname{d}^{k}_{S}\operatorname{tr}_{S}^{k}\hat{w})_{S}=0

by Lemma 4.1. Thus,

v^\displaystyle\hat{v} =l=k+1dSΔl(T^)j=1NSk(dSk,ψS,j,dSktrSkv^)SϕS,jk,I=0v0.\displaystyle=\sum_{l=k+1}^{d}\sum_{S\in\Delta_{l}(\hat{T})}\sum_{j=1}^{N_{S}^{k}}(\operatorname{d}^{k}_{S},\psi_{S,j},\operatorname{d}^{k}_{S}\operatorname{tr}_{S}^{k}\hat{v})_{S}\phi_{S,j}^{k,\mathrm{I}}=0\implies v\equiv 0.

The next result shows the kernel of dk\operatorname{d}^{k} consists of a low-order component from the Whitney forms plus the (high-order) type-II\mathrm{II} space, while the image of dk\operatorname{d}^{k} also contains a low-order component from the Whitney forms and a component from the high-order type-I\mathrm{I} space.

Lemma 5.2.

For k0:d1k\in 0:d-1, there holds

(5.5) ker(dk;Xk)=XΓk,IIX̊k,IIker(dk;Wk)=dk1X̊Γk1,Idk1X̊k1,Idk1Wk1k,\displaystyle\begin{aligned} \ker(\operatorname{d}^{k};X^{k})&=X^{k,\mathrm{II}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{II}}_{\mathcal{I}}\oplus\ker(\operatorname{d}^{k};W^{k})\\ &=\operatorname{d}^{k-1}\mathring{X}^{k-1,\mathrm{I}}_{\Gamma}\oplus\operatorname{d}^{k-1}\mathring{X}^{k-1,\mathrm{I}}_{\mathcal{I}}\oplus\operatorname{d}^{k-1}W^{k-1}\oplus\mathfrak{H}^{k},\end{aligned}

where k\mathfrak{H}^{k} denote the space of lowest-order harmonic forms:

(5.6) k:={𝔥ker(dk;Wk):(𝔥,dk1w)=0wWk1}.\displaystyle\mathfrak{H}^{k}:=\{\mathfrak{h}\in\ker(\operatorname{d}^{k};W^{k}):(\mathfrak{h},\operatorname{d}^{k-1}w)=0\ \operatorname{\forall}w\in W^{k-1}\}.

Moreover, there holds

(5.7) dkXk\displaystyle\operatorname{d}^{k}X^{k} =dkX̊Γk,IdkX̊k,IdkWk.\displaystyle=\operatorname{d}^{k}\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus\operatorname{d}^{k}\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}}\oplus\operatorname{d}^{k}W^{k}.
Proof.

Step 1: 5.5. Thanks to 5.4, XΓk,II=dk1X̊Γk1,IX^{k,\mathrm{II}}_{\Gamma}=\operatorname{d}^{k-1}\mathring{X}^{k-1,\mathrm{I}}_{\Gamma} and X̊k,II=dk1X̊k1,I\mathring{X}^{k,\mathrm{II}}_{\mathcal{I}}=\operatorname{d}^{k-1}\mathring{X}^{k-1,\mathrm{I}}_{\mathcal{I}}, and so XΓk,IIX̊k,IIker(dk;Xk)X^{k,\mathrm{II}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{II}}_{\mathcal{I}}\subset\ker(\operatorname{d}^{k};X^{k}).

Now suppose that uker(dk;Xk)u\in\ker(\operatorname{d}^{k};X^{k}). Then, u=uI+uII+wu=u_{\mathrm{I}}+u_{\mathrm{II}}+w, where uIX̊Γk,IX̊k,Iu_{\mathrm{I}}\in\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}}, uIIXΓk,IIX̊k,IIu_{\mathrm{II}}\in X^{k,\mathrm{II}}_{\Gamma}\oplus\mathring{X}^{k,\mathrm{II}}_{\mathcal{I}}, and wWkw\in W^{k}. Then,

0=dku=dkuI+dkuII+dkw=dkuI+dkw,\displaystyle 0=\operatorname{d}^{k}u=\operatorname{d}^{k}u_{\mathrm{I}}+\operatorname{d}^{k}u_{\mathrm{II}}+\operatorname{d}^{k}w=\operatorname{d}^{k}u_{\mathrm{I}}+\operatorname{d}^{k}w,

and so dkuIdkWk\operatorname{d}^{k}u_{\mathrm{I}}\in\operatorname{d}^{k}W^{k}. By Lemma 5.1, uI0u_{\mathrm{I}}\equiv 0, and so dkw0\operatorname{d}^{k}w\equiv 0. Equation 5.5 now follows.

Step 2: 5.7. Thanks to Step 1 and Lemma 5.1, we have dkXk=dkX̊Γk,IdkX̊k,IdkWk\operatorname{d}^{k}X^{k}=\operatorname{d}^{k}\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus\operatorname{d}^{k}\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}}\oplus\operatorname{d}^{k}W^{k}, and so it suffices to show that dkX̊Γk,IdkX̊k,I={0}\operatorname{d}^{k}\mathring{X}^{k,\mathrm{I}}_{\Gamma}\cap\operatorname{d}^{k}\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}}=\{0\}. Let uΓX̊Γk,Iu_{\Gamma}\in\mathring{X}^{k,\mathrm{I}}_{\Gamma} and uX̊k,Iu_{\mathcal{I}}\in\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}} satisfy dk(uΓu)=0\operatorname{d}^{k}(u_{\Gamma}-u_{\mathcal{I}})=0. By Lemma 5.1, uΓ=uu_{\Gamma}=u_{\mathcal{I}}, and so uΓ=u=0u_{\Gamma}=u_{\mathcal{I}}=0. ∎

5.3. Matrix properties on a physical cell

Let T𝒯hT\in\mathcal{T}_{h} and consider the local matrix AA corresponding to the restriction of ak(,)a^{k}(\cdot,\cdot) 2.6 to the cell TT:

Aij=(ϕi,βϕj)T+(dkϕi,αdkϕj)T,i,j1:dimXk(T),\displaystyle A_{ij}=(\phi_{i},\beta\phi_{j})_{T}+(\operatorname{d}^{k}\phi_{i},\alpha\operatorname{d}^{k}\phi_{j})_{T},\qquad i,j\in 1:\dim X^{k}(T),

where {ϕi}\{\phi_{i}\} consists of all global basis functions supported on TT. As in section 4.2, we partition AA into the contribution from interior and interface basis functions:

A=[AAΓAΓAΓΓ].\displaystyle A=\begin{bmatrix}A_{\mathcal{I}\mathcal{I}}&A_{\mathcal{I}\Gamma}\\ A_{\Gamma\mathcal{I}}&A_{\Gamma\Gamma}\end{bmatrix}.

If TT is an equilateral simplex, then Lemma 4.5 shows that AA_{\mathcal{I}\mathcal{I}} is diagonal and that the AΓA_{\mathcal{I}\Gamma} and AΓA_{\Gamma\mathcal{I}} blocks only contain a mass term. This structure is lost if TT is not an equilateral simplex due to the mapping Tk\mathcal{F}_{T}^{k}. However, both the coupling and the mass terms only introduce weak coupling in the sense that if we define

P:=[diag(A)00AΓΓ],\displaystyle P:=\begin{bmatrix}\operatorname{diag}(A_{\mathcal{I}\mathcal{I}})&0\\ 0&A_{\Gamma\Gamma}\end{bmatrix},

then PP is spectrally equivalent to AA, as the following result quantifies.

Lemma 5.3.

For k0:d1k\in 0:d-1 and T𝒯hT\in\mathcal{T}_{h}, there exists a constant C>0C>0 depending only on shape regularity, dd, kk, and diam(Ω)\mathrm{diam}(\Omega) such that

(5.8) κ2(P1A)Cmax{1,βαhT2}\displaystyle\kappa_{2}\left(P^{-1}A\right)\leq C\max\left\{1,\frac{\beta}{\alpha}h_{T}^{2}\right\}

where κ2()\kappa_{2}(\cdot) denotes the condition number with respect to the 2-norm.

The proof of Lemma 5.3 is given in section B.3.

Provided that β/α\beta/\alpha is not too large, 5.8 suggests that we may construct efficient preconditioners from existing ones by dropping the interior-interface coupling and by using a point-Jacobi (diagonal scaling) preconditioner on the interior basis functions. We combine this strategy with space decomposition preconditioners in the next section.

6. Preconditioning

In this section, we propose preconditioners for the Riesz maps 1.1, 1.2 and 1.3. We describe our preconditioners in terms of space decompositions [51, 52]. To capitalize on the weak coupling between interior and interface functions described in Lemma 5.3, we group the global interior and global interface functions together in their own subspaces:

(6.1) Xk\displaystyle X^{k}_{\mathcal{I}} :=X̊k,IX̊k,IIandXΓk:=WkX̊Γk,IXΓk,II.\displaystyle:=\mathring{X}^{k,\mathrm{I}}_{\mathcal{I}}\oplus\mathring{X}^{k,\mathrm{II}}_{\mathcal{I}}\quad\text{and}\quad X^{k}_{\Gamma}:=W^{k}\oplus\mathring{X}^{k,\mathrm{I}}_{\Gamma}\oplus X^{k,\mathrm{II}}_{\Gamma}.

We also collect all Nk:=dimXkN^{k}_{\mathcal{I}}:=\dim X^{k}_{\mathcal{I}} interior basis functions {ϕ,k:1:Nk}\{\phi_{\mathcal{I},\ell}^{k}:\ell\in 1:N^{k}_{\mathcal{I}}\}.

To illustrate how the interior-interface decoupling strategy in Lemma 5.3 may be applied to arbitrary space decompositions of XkX^{k}, we present the corresponding result for Schwarz methods with exact solvers. To this end, let {Xmk}m=1M\{X_{m}^{k}\}_{m=1}^{M} be a subspace decomposition of XkX^{k}:

(6.2) Xk=m=1MXmk,whereXmXkm1:M,\displaystyle X^{k}=\sum_{m=1}^{M}X_{m}^{k},\qquad\text{where}\quad X_{m}\subseteq X^{k}\quad\forall m\in 1:M,

where each space is equipped with the ak(,)a^{k}(\cdot,\cdot) inner product. Then, we may decouple the interior and interface functions to arrive at the finer decomposition

(6.3) Xk=m=1MXmkXΓk+Xk,\displaystyle X^{k}=\sum_{m=1}^{M}X^{k}_{m}\cap X^{k}_{\Gamma}+X^{k}_{\mathcal{I}},

where XmkXΓkX^{k}_{m}\cap X^{k}_{\Gamma} is equipped with ak(,)a^{k}(\cdot,\cdot) for m1:Mm\in 1:M and XkX^{k}_{\mathcal{I}} is equipped with a point-Jacobi solver

ak(u,v):==1Nkak(u,v)u,vXk,\displaystyle a_{\mathcal{I}}^{k}(u,v):=\sum_{\ell=1}^{N^{k}_{\mathcal{I}}}a^{k}(u_{\ell},v_{\ell})\qquad\forall u,v\in X^{k}_{\mathcal{I}},

where u==1Nkuu=\sum_{\ell=1}^{N^{k}_{\ell}}u_{\ell} with uspan{ϕk}u_{\ell}\in\operatorname{span}\{\phi^{k}_{\ell}\} and similar notation for vv. Note that the decomposition 6.3 and associated bilinear forms is equivalent to the decomposition

(6.4) Xk=m=1MXmkXΓk+=1Nkspan{ϕk},\displaystyle X^{k}=\sum_{m=1}^{M}X^{k}_{m}\cap X^{k}_{\Gamma}+\sum_{\ell=1}^{N^{k}_{\mathcal{I}}}\operatorname{span}\{\phi^{k}_{\ell}\},

where each space is equipped with ak(,)a^{k}(\cdot,\cdot); we will use 6.3 and 6.4 interchangeably. Then, the following result shows that the interior-interface splitting is, up to a constant independent of hh, pp, α\alpha, and β\beta, as stable as the original decomposition 6.2 provided that βh2/α1\beta h^{2}/\alpha\leq 1:

Theorem 6.1.

Suppose that the decomposition 6.2 is energy stable: For every uXku\in X^{k}, there exists umXmku_{m}\in X^{k}_{m}, m1:Mm\in 1:M such that

(6.5) m=1Mum=uandm=1Mak(um,um)\displaystyle\sum_{m=1}^{M}u_{m}=u\quad\text{and}\quad\sum_{m=1}^{M}a^{k}(u_{m},u_{m}) Cstable(α,β)ak(u,u),\displaystyle\leq C_{\mathrm{stable}}(\alpha,\beta)a^{k}(u,u),

where Cstable(α,β)1C_{\mathrm{stable}}(\alpha,\beta)\geq 1 is independent of uu. Then, the decomposition 6.4 is also energy stable: For every uXku\in X^{k}, there exists umXmkXΓku_{m}\in X^{k}_{m}\cap X^{k}_{\Gamma}, m1:Mm\in 1:M, and uXku_{\mathcal{I}}\in X^{k}_{\mathcal{I}}, such that

(6.6a) m=1Mum+u\displaystyle\sum_{m=1}^{M}u_{m}+u_{\mathcal{I}} =u,\displaystyle=u,
(6.6b) m=1Mak(um,um)+ak(u,u)\displaystyle\sum_{m=1}^{M}a^{k}(u_{m},u_{m})+a^{k}_{\mathcal{I}}(u_{\mathcal{I}},u_{\mathcal{I}}) CCstable(α,β)max{1,βαh2}a(u,u),\displaystyle\leq CC_{\mathrm{stable}}(\alpha,\beta)\max\left\{1,\frac{\beta}{\alpha}h^{2}\right\}a(u,u),

where CC depends only on shape regularity, dd, and kk. Moreover, ak(,)a^{k}_{\mathcal{I}}(\cdot,\cdot) is locally stable:

(6.7) ak(v,v)Cak(v,v)vXk.\displaystyle a^{k}(v,v)\leq Ca_{\mathcal{I}}^{k}(v,v)\qquad\forall v\in X^{k}_{\mathcal{I}}.

The proof of Theorem 6.1 appears in appendix C.

One can apply Theorem 6.1 to obtain condition number bounds on Schwarz methods associated to the subspace decomposition 6.4, where the spaces are equipped with the solvers described above. For example, if the subspaces {Xmk}m=1M\{X_{m}^{k}\}_{m=1}^{M} can be colored with NcolorN_{\mathrm{color}} colors, then {XmkXΓk}m=1M\{X_{m}^{k}\cap X^{k}_{\Gamma}\}_{m=1}^{M} and XkX^{k}_{\mathcal{I}} can be be colored with Ncolor+1N_{\mathrm{color}}+1 colors. Thanks to 6.6, 6.7, and standard theory [51, Theorem 2.7], the additive Schwarz preconditioner P1P^{-1} associated to the subspace decomposition 6.4 satisfies

κ2(P1A)CCstable(α,β)(Ncolor+1)max{1,βαh2},\displaystyle\kappa_{2}(P^{-1}A)\leq CC_{\mathrm{stable}}(\alpha,\beta)(N_{\mathrm{color}}+1)\max\left\{1,\frac{\beta}{\alpha}h^{2}\right\},

where CC depends only on shape regularity, dd, and kk. Similar results can be obtained for multiplicative and hybrid methods.

Note that a Schwarz method associated to the decomposition 6.4 is substantially cheaper than the Schwarz method associated to the original decomposition 6.2, as the interior and interface functions are completely decoupled and a point-Jacobi preconditioner is employed on the interior basis functions. Decoupling the interior and interface functions is reminiscent of static condensation in which the 𝒪(p3)\mathcal{O}(p^{3}) interior degrees of freedom are exactly eliminated from the global system at a cost of 𝒪(p9)\mathcal{O}(p^{9}) operations, and problem is reformulated only on the interface degrees of freedom. The key advantage of the decomposition 6.4 is that we retain the decoupling of the interior and interface functions without actually performing static condensation, and we only require a point-Jacobi preconditioner for the interior degrees of freedom.

We now turn to specific space decompositions.

6.1. Pavarino–Arnold–Falk–Winther decompositions

To start, we consider the Pavarino–Arnold–Falk–Winther (PAFW) decomposition [43, 4, 53, 47, 15]. The decomposition for kk-forms is given by

(6.8) Xk=Wk+VΔ0(𝒯h)Xk|V,X^{k}=W^{k}+\sum_{V\in\Delta_{0}(\mathcal{T}_{h})}\left.X^{k}\right|_{\star V},

which consists of two components. The first summand is the lowest-degree space and plays the role of the coarse grid, propagating coarse-scale information globally. The second summand restricts the fine space to small subspaces, each supported on the star of a vertex V\star V, where VΔ0(𝒯h)V\in\Delta_{0}(\mathcal{T}_{h}) is a vertex of the mesh. The star operation is well-known in algebraic topology [39, §2][19, 22]; it gathers all entities that contain the given entity as a subentity. In particular, the star of a vertex returns all cells, faces, and edges that are adjacent to this vertex, as well as the vertex itself, and functions in Xk|V\left.X^{k}\right|_{\star V} have vanishing trace on (V)\partial(\star V). This second summand plays the role of a multigrid relaxation, or the subdomain solves in a domain decomposition solver.

We denote the decomposition 6.8 by PAFW0(Xk)\mathrm{PAFW}_{0}({X^{k}}) to indicate that the space in the summation is XkX^{k} restricted to the star of a 0-dimensional subsimplex. The decomposition satisfies the hypothesis of Theorem 6.1 with CstableC_{\mathrm{stable}} independent of α\alpha, β\beta, hh, and pp [21]. However, on a regular (Freudenthal [11]) mesh, an interior vertex star contains 14 edges, 36 faces, and 24 cells, and the vertex-star patch problems become very expensive to solve, particularly at high order. Thus, we employ the subspace splitting in 6.4:

(6.9) Xk=Wk+VΔ0(𝒯h)XΓk|V+=1Nkspan{ϕk},\displaystyle X^{k}=W^{k}+\sum_{V\in\Delta_{0}(\mathcal{T}_{h})}\left.X^{k}_{\Gamma}\right|_{\star V}+\sum_{\ell=1}^{N^{k}_{\mathcal{I}}}\operatorname{span}\left\{\phi_{\ell}^{k}\right\},

which we refer to as PAFW0(XΓk)+J(Xk)\mathrm{PAFW}_{0}({X^{k}_{\Gamma}})+\mathrm{J}({X^{k}_{\mathcal{I}}}), where J(Xk)\mathrm{J}({X^{k}_{\mathcal{I}}}) indicates that a point-Jacobi preconditioner is employed on the space of interior functions. Theorem 6.1 shows that the decomposition 6.9 is also uniformly stable with respect to hh and pp and stable in α\alpha and β\beta provided that the ratio βh2/α\beta h^{2}/\alpha is bounded. Crucially, the dimension of the vertex-star problems scales like 𝒪(pd1)\mathcal{O}(p^{d-1}) for 6.9, while it scales like 𝒪(pd)\mathcal{O}(p^{d}) for 6.8. For example, at degree p=4p=4 and p=10p=10, the vertex star X0|V\left.X^{0}\right|_{\star V} has dimension 175 and 3439, respectively, while XΓ0|V\left.X_{\Gamma}^{0}\right|_{\star V} has dimension 151 and 1423, respectively.

6.2. Hiptmair–Toselli space decompositions

For k>0k>0, finer space decompositions are possible and more efficient. Arnold, Falk & Winther [4] prove that one can employ stars over entities of dimension k1k-1 instead of dimension 0 , with the resulting patches being much smaller for k>1k>1. Furthermore, as observed by Hiptmair and Toselli [28, 25, 29], one can even use stars over entities of dimension kk by introducing a potential space. This approach splits XkX^{k} as the sum of the image of the exterior derivative dk1\operatorname{d}^{k-1} on a potential space Yk1Y^{k-1}, and some other space ZkXkZ^{k}\subset X^{k}:

(6.10) Xk=dk1Yk1+Zk.X^{k}=\operatorname{d}^{k-1}Y^{k-1}+Z^{k}.

The challenge here is to choose Yk1Y^{k-1} and ZkZ^{k} as small as possible; e.g. excluding as much of the kernel of dk1\operatorname{d}^{k-1} as possible from Yk1Y^{k-1} and as much of the kernel of dk\operatorname{d}^{k} from ZkZ^{k}.

6.2.1. Standard Hiptmair–Toselli decomposition

The largest choice Yk1=Xk1Y^{k-1}=X^{k-1} and Zk=XkZ^{k}=X^{k} motivates the Hiptmair–Toselli space decomposition

(6.11) Xk=Wk+JΔk1(𝒯h)dk1Xk1|J+LΔk(𝒯h)Xk|L,X^{k}=W^{k}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}\left.\operatorname{d}^{k-1}X^{k-1}\right|_{\star J}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}\left.X^{k}\right|_{\star L},

where introducing the potential space in 6.10 enables the splitting over stars of dimension k1k-1 and kk, respectively. We denote this decomposition by HT(Xk1,Xk)\mathrm{HT}({X^{k-1}},{X^{k}}) to indicate that potential space is Xk1X^{k-1} and that the kk-dimensional stars are taken over XkX^{k}. As shown in [28, 25, 29], the HT(Xk1,Xk)\mathrm{HT}({X^{k-1}},{X^{k}}) decomposition satisfies the hypothesis of Theorem 6.1 with CstableC_{\mathrm{stable}} independent of hh, α\alpha, and β\beta. As before, the HT(Xk1,Xk)\mathrm{HT}({X^{k-1}},{X^{k}}) decomposition can be combined with Theorem 6.1, yielding the HT(XΓk1,XΓk)+J(Xk)\mathrm{HT}({X^{k-1}_{\Gamma}},{X^{k}_{\Gamma}})+\mathrm{J}({X^{k}_{\mathcal{I}}}) decomposition.

Note that the HT(Xk1,Xk)\mathrm{HT}({X^{k-1}},{X^{k}}) decomposition involves an auxiliary problem on the local potential space Xk1|JX^{k-1}|_{\star J}: find ψXk1|J\psi\in X^{k-1}|_{\star J} such that

(6.12) ak(dk1ϕ,dk1ψ)=(dk1ϕ,βdk1ψ)J=F(dk1ϕ)ϕXk1|J,a^{k}(\operatorname{d}^{k-1}\phi,\operatorname{d}^{k-1}\psi)=(\operatorname{d}^{k-1}\phi,\beta\operatorname{d}^{k-1}\psi)_{\star J}=F(\operatorname{d}^{k-1}\phi)\quad\operatorname{\forall}\phi\in\left.X^{k-1}\right|_{\star J},

since dkdk1=0\operatorname{d}^{k}\circ\operatorname{d}^{k-1}=0. For k>1k>1, problem 6.12 is singular since dk2Xk2ker(dk1;Xk1)\operatorname{d}^{k-2}X^{k-2}\subset\ker(\operatorname{d}^{k-1};X^{k-1}). To address this, a pragmatic choice is to solve the regularized problem

(6.13) (ϕ,ϵψ)J+(dk1ϕ,βdk1ψ)J=F(dk1ϕ)ϕXk1|J,(\phi,\epsilon\psi)_{\star J}+(\operatorname{d}^{k-1}\phi,\beta\operatorname{d}^{k-1}\psi)_{\star J}=F(\operatorname{d}^{k-1}\phi)\quad\operatorname{\forall}\phi\in\left.X^{k-1}\right|_{\star J},

for small ϵ\epsilon; in this work ϵ=108β\epsilon=10^{-8}\beta is used. This regularization can be avoided by posing 6.12 on a reduced space, as discussed in the next section.

6.2.2. Type-I\mathrm{I} based decomposition

The finest possible choice for Yk1Y^{k-1} and ZkZ^{k} in 6.10 is to ensure

Yk1ker(dk1;Xk1)={0}andZkdk1Xk1={0}.\displaystyle Y^{k-1}\cap\ker(\operatorname{d}^{k-1};X^{k-1})=\{0\}\quad\text{and}\quad Z^{k}\cap\operatorname{d}^{k-1}X^{k-1}=\{0\}.

However, this is not practical. As shown in 5.5, one component of ker(dk1,Xk1)\ker(\operatorname{d}^{k-1},X^{k-1}) is the space of low-order harmonic forms k\mathfrak{H}^{k} 5.6 which depends on the topology of Ω\Omega, and forms in k\mathfrak{H}^{k} have global support owing to the orthogonality condition in 5.6. Instead, a pragmatic intermediate choice motivated by Lemma 5.2 is to employ

(6.14) Yk1=Xk1,IandZk=Xk,I,Y^{k-1}=X^{k-1,\mathrm{I}}\quad\text{and}\quad Z^{k}=X^{k,\mathrm{I}},

i.e. to use only the span of the type-I basis functions. This choice excludes all type-II\mathrm{II} basis functions in the potential space Yk1Y^{k-1} and in the space ZkZ^{k}, and this choice is simple to construct because our basis is already split in this fashion. Since X0,I=X0X^{0,\mathrm{I}}=X^{0}, we only obtain a reduction compared to the largest choice of the potential space Yk1=Xk1Y^{k-1}=X^{k-1} if k>1k>1. We then arrive at the HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}}) decomposition:

(6.15) Xk=Wk+JΔk1(𝒯h)dk1Xk1,I|J+LΔk(𝒯h)Xk,I|L,X^{k}=W^{k}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}\left.\operatorname{d}^{k-1}X^{k-1,\mathrm{I}}\right|_{\star J}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}\left.X^{k,\mathrm{I}}\right|_{\star L},

which we believe to be novel.

The local potential problems for the HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}}) decomposition take the following form for JΔk1(𝒯h)J\in\Delta_{k-1}(\mathcal{T}_{h}): Find ψXk1,I|J\psi\in X^{k-1,\mathrm{I}}|_{\star J} such that

(6.16) ak(dk1ϕ,dk1ψ)=(dk1ϕ,βdk1ψ)J=F(dk1ϕ)ϕXk1,I|J.a^{k}(\operatorname{d}^{k-1}\phi,\operatorname{d}^{k-1}\psi)=(\operatorname{d}^{k-1}\phi,\beta\operatorname{d}^{k-1}\psi)_{\star J}=F(\operatorname{d}^{k-1}\phi)\quad\operatorname{\forall}\phi\in\left.X^{k-1,\mathrm{I}}\right|_{\star J}.

Note that if wWk1|Jw\in W^{k-1}|_{\star J} satisfies dk1w=0\operatorname{d}^{k-1}w=0, then w0w\equiv 0, and so problem 6.16 is invertible thanks to Lemmas 5.1 and 5.2. In particular, problem 6.16 involves fewer degrees of freedom than problem 6.13. A similar approach was used by Zaglmayr [53, §6.3] to solve the curl-curl problem on the whole domain Ω\Omega by posing the problem only the type-I\mathrm{I} space X1,IX^{1,\mathrm{I}}.

Remark 6.2.

Since the type-I\mathrm{I} space Xk1,IX^{k-1,\mathrm{I}} captures the image of the exterior derivative dk1\operatorname{d}^{k-1} (i.e. dk1Xk1=dk1Xk1,I\operatorname{d}^{k-1}X^{k-1}=\operatorname{d}^{k-1}X^{k-1,\mathrm{I}}), the space decompositions HT(Xk1,Xk)\mathrm{HT}({X^{k-1}},{X^{k}}) and HT(Xk1,I,Xk)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k}}) are identical and thus satisfy the hypotheses of Theorem 6.1 with the same stability constant. The only distinction we make in the notation is the form of the potential problems 6.13 and 6.16. So, one could always use the type-I\mathrm{I} space in the potential space and avoid solving the singular problem 6.12.

Combining 6.15 with Theorem 6.1 yields the HT(XΓk1,I,XΓk,I)+J(Xk)\mathrm{HT}({X^{k-1,\mathrm{I}}_{\Gamma}},{X^{k,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{k}_{\mathcal{I}}}) decomposition:

(6.17) Xk=Wk+JΔk1(𝒯h)dk1XΓk1,I|J+LΔk(𝒯h)XΓk,I|L+=1Nkspan{ϕk}.X^{k}=W^{k}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}\left.\operatorname{d}^{k-1}X^{k-1,\mathrm{I}}_{\Gamma}\right|_{\star J}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}\left.X^{k,\mathrm{I}}_{\Gamma}\right|_{\star L}+\sum_{\ell=1}^{N^{k}_{\mathcal{I}}}\operatorname{span}\left\{\phi_{\ell}^{k}\right\}.

In Table 1 we record the dimension of the largest subspace on vertices and edges in the proposed space decompositions for H(curl)H(\operatorname{curl}). We again observe that applying Theorem 6.1 reduces patch size, that HT(X0,X1)\mathrm{HT}({X^{0}},{X^{1}}) and HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) involve much smaller patches than PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) and PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}), respectively, and that using only the type-I\mathrm{I} spaces as in HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) and HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) yields a further substantial improvement to the size of the edge problems solved.

Table 1. Maximum subspace sizes for solving the H(curl)H(\operatorname{curl}) Riesz map with the PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) 6.8, PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) 6.9, HT(X0,X1)\mathrm{HT}({X^{0}},{X^{1}}) 6.11, HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}), HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) 6.15, and HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) 6.17 decompositions on a regular mesh for the Nedéléc space of the first kind. We distinguish between the maximal vertex-star patch and maximal edge-star patch. The edge patch includes 6 cell interiors, 6 faces, and one edge.
decomposition X1X^{1} max. vertex dim max. edge dim
PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) Ned41\mathrm{Ned}^{1}_{4} 776 -
PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) Ned41\mathrm{Ned}^{1}_{4} 488 -
HT(X0,X1)\mathrm{HT}({X^{0}},{X^{1}}) Ned41\mathrm{Ned}^{1}_{4} 175 148
HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) Ned41\mathrm{Ned}^{1}_{4} 151 76
HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) Ned41\mathrm{Ned}^{1}_{4} 175 121
PH(XΓ0X_{\Gamma}^{0}, XΓ1,IX_{\Gamma}^{1,\mathrm{I}})++J(X1X^{1}_{\mathcal{I}}) Ned41\mathrm{Ned}^{1}_{4} 151 55
PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) Ned101\mathrm{Ned}^{1}_{10} 12020 -
PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) Ned101\mathrm{Ned}^{1}_{10} 3380 -
HT(X0,X1)\mathrm{HT}({X^{0}},{X^{1}}) Ned101\mathrm{Ned}^{1}_{10} 3439 2710
HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) Ned101\mathrm{Ned}^{1}_{10} 1423 550
HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) Ned101\mathrm{Ned}^{1}_{10} 3439 1981
PH(XΓ0X_{\Gamma}^{0}, XΓ1,IX_{\Gamma}^{1,\mathrm{I}})++J(X1X^{1}_{\mathcal{I}}) Ned101\mathrm{Ned}^{1}_{10} 1423 325

6.2.3. Stability of HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}})

We arrived at the HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}}) decomposition by an algebraic decomposition of dk1Xk1\operatorname{d}^{k-1}X^{k-1} and XkX^{k}, which does not necessarily imply that HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}}) is a stable decomposition of XkX^{k}. The next result shows that HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}}) is indeed a stable decomposition of XkX^{k} with respect to the mesh size hh.

Lemma 6.3.

Let k0:d1k\in 0:d-1 and CHTC_{\mathrm{HT}} denote the smallest constant independent of α\alpha and β\beta such that the HT(Xk1,Xk)\mathrm{HT}({X^{k-1}},{X^{k}}) decomposition 6.11 satisfies the stability hypothesis of Theorem 6.1. Then, there exists a constant CI/IIC_{\mathrm{I}/\mathrm{II}} independent of hh, α\alpha, and β\beta such that the HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}}) decomposition 6.15 satisfies the stability hypothesis of Theorem 6.1 with Cstable(α,β)CI/IICHTC_{\mathrm{stable}}(\alpha,\beta)\leq C_{\mathrm{I}/\mathrm{II}}C_{\mathrm{HT}}.

The proof of Lemma 6.3 is given in section D.1.

The constant CI,IIC_{\mathrm{I},\mathrm{II}} in Lemma 6.3 may depend on the polynomial degree. Indeed, the numerical experiments below in section 7 suggest that using the Hiptmair–Toselli HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}}) decomposition to construct preconditioners does not provide pp-independent iteration counts, which in turn suggests that CI,IIC_{\mathrm{I},\mathrm{II}} depends on pp. A more detailed discussion of CI,IIC_{\mathrm{I},\mathrm{II}} is given in section D.2, where further experiments suggest that CI,IIC_{\mathrm{I},\mathrm{II}}\to\infty as pp\to\infty.

6.3. Decomposition for H(div)H(\operatorname{div})

For d=3d=3 and k=2k=2, the HT(X1,I,X2,I)\mathrm{HT}({X^{1,\mathrm{I}}},{X^{2,\mathrm{I}}}) decomposition 6.15 simplifies as 3.21c and 5.1d show that XΓ2,I=W2X^{2,\mathrm{I}}_{\Gamma}=W^{2}. The last term in 6.17 then becomes a splitting of W2W^{2}, which is not necessary as W2W^{2} is already included in 6.15. Combining this fact with Theorem 6.1 leads to the following decomposition:

(6.18) X2=W2+EΔ1(𝒯h)curlXΓ1,I|E+=1N2span{ϕ2}.\displaystyle X^{2}=W^{2}+\sum_{E\in\Delta_{1}(\mathcal{T}_{h})}\left.\operatorname{curl}X_{\Gamma}^{1,\mathrm{I}}\right|_{\star E}+\sum_{\ell=1}^{N^{2}_{\mathcal{I}}}\operatorname{span}\left\{\phi_{\ell}^{2}\right\}.

Moreover, the interface potential problem on the edge star EΔ1(𝒯h)E\in\Delta_{1}(\mathcal{T}_{h})

(6.19) ϕXΓ1,I|E:(curlϕ,βcurlψ)=F(curlψ)ψXΓ1,I|E\displaystyle\phi\in\left.X_{\Gamma}^{1,\mathrm{I}}\right|_{\star E}:\qquad(\operatorname{curl}\phi,\beta\operatorname{curl}\psi)=F(\operatorname{curl}\psi)\qquad\forall\psi\in\left.X_{\Gamma}^{1,\mathrm{I}}\right|_{\star E}

and the full interface problem on the edge star

(6.20) uXΓ2|E:(u,βv)+(divu,αdivv)=F(v)vXΓ2|E\displaystyle u\in\left.X_{\Gamma}^{2}\right|_{\star E}:\qquad(u,\beta v)+(\operatorname{div}u,\alpha\operatorname{div}v)=F(v)\qquad\forall v\in\left.X_{\Gamma}^{2}\right|_{\star E}

have a comparable number of interface unknowns. In particular, 5.4 and 5.7 give

XΓ2|E=XΓ2,I|EXΓ2,II|E=W2|E+curlXΓ1,I|E=W2|EcurlX̊Γ1,I|E,\displaystyle\left.X_{\Gamma}^{2}\right|_{\star E}=\left.X_{\Gamma}^{2,\mathrm{I}}\right|_{\star E}\oplus\left.X_{\Gamma}^{2,\mathrm{II}}\right|_{\star E}=\left.W^{2}\right|_{\star E}+\left.\operatorname{curl}X_{\Gamma}^{1,\mathrm{I}}\right|_{\star E}=\left.W^{2}\right|_{\star E}\oplus\left.\operatorname{curl}\mathring{X}_{\Gamma}^{1,\mathrm{I}}\right|_{\star E},

and so 6.20 has only |Δ2(E)|1|\Delta_{2}(\star E)|-1 more unknowns than 6.19. We can thus avoid using the potential space XΓ1,IX^{1,\mathrm{I}}_{\Gamma} altogether without substantially increasing the patch size by replacing curlXΓ1,I|E\left.\operatorname{curl}X_{\Gamma}^{1,\mathrm{I}}\right|_{\star E} with XΓ2X_{\Gamma}^{2} to obtain

(6.21) X2=W2+EΔ1(𝒯h)XΓ2|E+=1N2span{ϕ2}.X^{2}=W^{2}+\sum_{E\in\Delta_{1}(\mathcal{T}_{h})}\left.X^{2}_{\Gamma}\right|_{\star E}+\ \sum_{\ell=1}^{N^{2}_{\mathcal{I}}}\operatorname{span}\{\phi_{\ell}^{2}\}.

We denote this decomposition by PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}), as it is of the form 6.9, with the only difference being that edge stars are used instead of vertex stars. As shown in [4], PAFW1(X2)\mathrm{PAFW}_{1}({X^{2}}) satisfies the assumptions of Theorem 6.1 with CstableC_{\mathrm{stable}} independent of hh, α\alpha, and β\beta, and so 6.21 is uniformly stable with respect to hh, α\alpha, and β\beta provided that βh2/α\beta h^{2}/\alpha is bounded. We remark that the pp-stability of PAFW1(X2)\mathrm{PAFW}_{1}({X^{2}}), and PAFWk1(Xk)\mathrm{PAFW}_{k-1}({X^{k}}) more generally, remains an open problem for k>1k>1.

In Table 2 we record the dimension of the largest subspace on vertices, edges, and faces in the proposed space decompositions for H(div)H(\operatorname{div}). We observe that applying Theorem 6.1 drastically reduces the patch size for each decomposition, while HT(XΓ1,I,XΓ2,I)+J(X2)\mathrm{HT}({X^{1,\mathrm{I}}_{\Gamma}},{X^{2,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) and PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) have the smallest subproblems of about the same size.

Table 2. Maximum subspace sizes for solving the H(div)H(\operatorname{div}) Riesz map with the PAFW0(X2)\mathrm{PAFW}_{0}({X^{2}}) 6.8, PAFW0(XΓ2)+J(X2)\mathrm{PAFW}_{0}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) 6.9, HT(X1,X2)\mathrm{HT}({X^{1}},{X^{2}}) 6.11, HT(XΓ1,I,XΓ2,I)+J(X2)\mathrm{HT}({X^{1,\mathrm{I}}_{\Gamma}},{X^{2,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) 6.18, and PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) 6.21 decompositions on a regular mesh for the Raviart–Thomas space. The face patch includes 2 cell interiors, and one face.
decomposition X2X^{2} max. vertex dim max. edge dim max. face dim
PAFW0(X2)\mathrm{PAFW}_{0}({X^{2}}) RT4\mathrm{RT}_{4} 1080 - -
PAFW0(XΓ2)+J(X2)\mathrm{PAFW}_{0}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) RT4\mathrm{RT}_{4} 360 - -
HT(X1,X2)\mathrm{HT}({X^{1}},{X^{2}}) RT4\mathrm{RT}_{4} - 148 70
HT(XΓ1,I,XΓ2,I)+J(X2)\mathrm{HT}({X^{1,\mathrm{I}}_{\Gamma}},{X^{2,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) RT4\mathrm{RT}_{4} - 55 -
PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) RT4\mathrm{RT}_{4} - 60 -
PAFW0(X2)\mathrm{PAFW}_{0}({X^{2}}) RT10\mathrm{RT}_{10} 13860 - -
PAFW0(XΓ2)+J(X2)\mathrm{PAFW}_{0}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) RT10\mathrm{RT}_{10} 1980 - -
HT(X1,X2)\mathrm{HT}({X^{1}},{X^{2}}) RT10\mathrm{RT}_{10} - 2710 1045
HT(XΓ1,I,XΓ2,I)+J(X2)\mathrm{HT}({X^{1,\mathrm{I}}_{\Gamma}},{X^{2,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) RT10\mathrm{RT}_{10} - 325 -
PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) RT10\mathrm{RT}_{10} - 330 -

6.4. Computational cost

Our solver is globally matrix-free, but requires assembly of interface submatrices over patches of mesh entities of dimension at most d1d-1. In order to avoid computing and storing the matrix entries ak(ϕjk,ϕik)a^{k}(\phi^{k}_{j},\phi^{k}_{i}), we instead employ a Krylov method only requiring the computation of ak(u,ϕik)a^{k}(u,\phi^{k}_{i}) for a given uXku\in X^{k}. Given a quadrature rule with 𝒪(pd)\mathcal{O}(p^{d}) quadrature points, and without sum-factorization, this can be done in 𝒪(p2d)\mathcal{O}(p^{2d}) operations and 𝒪(pd)\mathcal{O}(p^{d}) storage per cell. However, to form the basis there is an offline cost of 𝒪(p3d)\mathcal{O}(p^{3d}) flops and 𝒪(p2d)\mathcal{O}(p^{2d}) storage to solve the eigenproblem 3.13 and tabulate the basis at quadrature points. Nevertheless, the tabulation is precomputed and stored, so the 𝒪(p9)\mathcal{O}(p^{9}) flops are only incurred once on the reference cell, and the 𝒪(p6)\mathcal{O}(p^{6}) memory requirement does not scale with the number of cells.

The interface submatrices cost 𝒪(p3d2)\mathcal{O}(p^{3d-2}) flops to assemble and have 𝒪(p2(d1))\mathcal{O}(p^{2(d-1)}) entries. The solution of each interface problem involves the factorization of the submatrix, which requires 𝒪(p3(d1))\mathcal{O}(p^{3(d-1)}) flops and 𝒪(p2(d1))\mathcal{O}(p^{2(d-1)}) storage per patch. At each Krylov iteration, solving the interface subproblems incurs 𝒪(p2(d1))\mathcal{O}(p^{2(d-1)}) flops per patch. In three dimensions, the costs of operator application and preconditioner setup and application amount to 𝒪(p6)\mathcal{O}(p^{6}) flops and 𝒪(p4)\mathcal{O}(p^{4}) storage.

This is illustrated in Figure 1, where we record the total number of floating point operations, memory storage, and matrix nonzeros required to solve the Riesz maps using the space decompositions with and without the interior-interface decomposition. For H(grad)H(\operatorname{grad}) we use PAFW0(XΓ0)+J(X0)\mathrm{PAFW}_{0}({X^{0}_{\Gamma}})+\mathrm{J}({X^{0}_{\mathcal{I}}}), for H(curl)H(\operatorname{curl}) we use HT(XΓ0,XΓ1,I)+J(X2)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}), and for H(div)H(\operatorname{div}) we use PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}). The analogous decompositions before interior-interface splitting result in 𝒪(p9)\mathcal{O}(p^{9}) flops and 𝒪(p6)\mathcal{O}(p^{6}) storage and nonzeros.

7. Numerical results

All of the above bases and preconditioners have been implemented in Firedrake [24], which we use to perform the numerical experiments. For brevity we only present results for spaces of the first kind.

7.1. Riesz maps

We first discretize each of the Riesz maps 2.6 on an initial Freudenthal mesh of Ω=(0,1)3\Omega=(0,1)^{3} with three elements along each edge (see Figure 5(a) for the case of one element per edge) with pure Neumann boundary conditions (ΓN=Ω\Gamma_{N}=\partial\Omega). For each Riesz map, we fix β=1\beta=1, vary α{103,1,103}\alpha\in\{10^{3},1,10^{-3}\} and p3:7p\in 3:7, and perform two levels of mesh refinement following [11]. Note that the meshes resulting from refinement have a smaller shape regularity constant compared to the original Freudenthal mesh, but the shape regularity constant remains bounded away from zero as more refinements are performed. We take the right hand side of the resulting linear system to be a random vector and solve with the preconditioned conjugate gradient (PCG) method with a relative tolerance of 10810^{-8}.

(a)
Refer to caption
(b)
Figure 5. (a) Freudenthal subdivision of a cube and (b) mesh of the Fichera corner.

7.1.1. H(grad)H(\operatorname{grad}) solvers

We consider the subspace decomposition PAFW0(X0)\mathrm{PAFW}_{0}({X^{0}}) 6.8 and the corresponding interior-interface splitting PAFW0(XΓ0)+J(X0)\mathrm{PAFW}_{0}({X^{0}_{\Gamma}})+\mathrm{J}({X^{0}_{\mathcal{I}}}) 6.9. Here, and in the following examples, we use a symmetric hybrid Schwarz method preconditioner. In particular, the solvers from each group of spaces separated by a “++” in the decompositions

(7.1) Xk=Wk+VΔ0(𝒯h)Xk|VandXk=Wk+VΔ0(𝒯h)XΓk|V+=1Nkspan{ϕk}\displaystyle X^{k}=W^{k}+\sum_{V\in\Delta_{0}(\mathcal{T}_{h})}\left.X^{k}\right|_{\star V}\quad\text{and}\quad X^{k}=W^{k}+\sum_{V\in\Delta_{0}(\mathcal{T}_{h})}\left.X^{k}_{\Gamma}\right|_{\star V}+\sum_{\ell=1}^{N^{k}_{\mathcal{I}}}\operatorname{span}\left\{\phi_{\ell}^{k}\right\}

are weighted using estimated extremal eigenvalues and combined multiplicatively from right to left to right for symmetry. We equip W0W^{0} with a geometric multigrid V-cycle with vertex patch relaxation (equivalently point-Jacobi) and a direct Cholesky solve on the coarsest mesh, while the remaining subspaces are equipped with exact solvers.

To more precisely describe the multiplicative splitting and the computation of the weights, we rewrite the hybrid methods in 7.1 as a symmetric multiplicative Schwarz method with inexact solvers:

(7.2) Xk=m=1MXmk,\displaystyle X^{k}=\sum_{m=1}^{M}X^{k}_{m},

where each space XmkX^{k}_{m} is equipped with a weight ρm>0\rho_{m}>0 and the inexact solver ρma~m(,)\rho_{m}\tilde{a}_{m}(\cdot,\cdot), and the multiplicative sweep is done X1k,,Xmk,X1kX^{k}_{1},\ldots,X^{k}_{m},\ldots X^{k}_{1}. For example, the second decomposition in 7.1 satisfies

X10=X0,X20=XΓ0,andX30=W0,\displaystyle X^{0}_{1}=X^{0}_{\mathcal{I}},\quad X^{0}_{2}=X^{0}_{\Gamma},\quad\text{and}\quad X^{0}_{3}=W^{0},

where a~1(,)\tilde{a}_{1}(\cdot,\cdot) is a point-Jacobi solver, a~2(,)\tilde{a}_{2}(\cdot,\cdot) is the additive Schwarz method with decomposition VΔ0(𝒯h)XΓ0|V\sum_{V\in\Delta_{0}(\mathcal{T}_{h})}\left.X^{0}_{\Gamma}\right|_{\star V} and exact solvers (additive patch solver on 0-stars), and a~3(,)\tilde{a}_{3}(\cdot,\cdot) is a geometric multigrid V-cycle as above. The first decomposition in 7.1 can be described similarly. The weight ρm\rho_{m} is taken to be

(7.3) ρm=λ~min+3λ~max4\rho_{m}=\frac{\tilde{\lambda}_{\min}+3\tilde{\lambda}_{\max}}{4}

where λ~min\tilde{\lambda}_{\min} and λ~max\tilde{\lambda}_{\max} are estimates of the extreme generalized eigenvalues of

uXm0,λ:a(u,v)=λa~m(u,v)vXm0\displaystyle u\in X^{0}_{m},\lambda\in\mathbb{R}:\qquad a(u,v)=\lambda\tilde{a}_{m}(u,v)\qquad\forall v\in X^{0}_{m}

computed from 10 conjugate gradient iterations with a randomized right hand side [34]. With this language, the PAFW0(XΓ0)+J(X0)\mathrm{PAFW}_{0}({X^{0}_{\Gamma}})+\mathrm{J}({X^{0}_{\mathcal{I}}}) solver is summarized in Figure 6.

KSP: conjugate gradientsPC: symmetric multiplicative Schwarz XkX^{k}_{\mathcal{I}}: point-Jacobi on cell interiors XΓkX^{k}_{\Gamma}: interface additive patches on ll-stars WkW^{k}: geometric multigrid V-cycle Relaxation: patches on ll-stars hh-coarse: Cholesky
Figure 6. Solver diagram for PAFWl(XΓk)+J(Xk)\mathrm{PAFW}_{l}({X^{k}_{\Gamma}})+\mathrm{J}({X^{k}_{\mathcal{I}}}) used for H(grad)H(\operatorname{grad}) and H(curl)H(\operatorname{curl}) (l=0l=0) and H(div)H(\operatorname{div}) (l=1l=1).

The iteration counts for the two decompositions in 7.1 with p5:7p\in 5:7 are displayed in section 7.1.1. Here, and in the tables that follow, each column corresponds to a different value of α\alpha, and the number of iterations for the solver with interior-interface splitting are displayed on the left followed by the iteration counts for the corresponding solver without the interior-interface splitting in brackets. We observe that across all values of α\alpha and pp, the computationally cheaper PAFW0(XΓ0)+J(X0)\mathrm{PAFW}_{0}({X^{0}_{\Gamma}})+\mathrm{J}({X^{0}_{\mathcal{I}}}) solver performs within 3 iterations of the full PAFW0(XΓ0)\mathrm{PAFW}_{0}({X^{0}_{\Gamma}}) solver, and all of these iteration counts are robust with respect to the α\alpha, pp, and hh.

Table 3. PCG iteration counts for the H(grad)H(\operatorname{grad}) Riesz map, where Y0=XΓ0(Y0=X0)Y^{0}=X^{0}_{\Gamma}\ (Y^{0}=X^{0}).
\csvreader

[ head to column names, head to column names prefix=MY, tabular=rcccc, table head= PAFW0(Y0)\mathrm{PAFW}_{0}({Y^{0}})
ll pp α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree5 \equal\MYdegree6 \equal\MYdegree7, table foot=]results/CG.3d.demkowicz.afw.combined.csv \MYlevel \MYdegree \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations)

7.1.2. H(curl)H(\operatorname{curl}) solvers

We first consider the subspace decompositions based on vertex patches PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) and PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) using the same symmetric hybrid Schwarz solver described above and in Figure 6. Note that the vertex patch relaxation of the geometric multigrid V-cycle preconditioner on W1W^{1} is no longer equivalent to a point-Jacobi relaxation. The iteration counts for these two solvers with p{3,5,7}p\in\{3,5,7\} are displayed in the leftmost section of section 7.1.2. For α1\alpha\geq 1, PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) and PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) are within 7 iterations of one another, while for α=103\alpha=10^{-3}, PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) requires about double the number of iterations as PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}). The robustness of the interior-interface splitting for α1\alpha\geq 1 and its degradation for α=103\alpha=10^{-3} is consistent with the stability of the decomposition in Theorem 6.1.

We now consider two Hiptmair–Toselli variants HT(X0,X1)\mathrm{HT}({X^{0}},{X^{1}}) and HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) along with their interior-interface splitting counterparts HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) and HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}). We use an analogous hybrid Schwarz method preconditioner as for PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}). For example, the HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) solver can be expressed as a symmetric multiplicative Schwarz method as in 7.2 with weighted inexact solvers by choosing

X11=X1,X21=XΓ1,andX31=W1.\displaystyle X_{1}^{1}=X^{1}_{\mathcal{I}},\quad X^{1}_{2}=X^{1}_{\Gamma},\quad\text{and}\quad X^{1}_{3}=W^{1}.

As above, a~1(,)\tilde{a}_{1}(\cdot,\cdot) is point-Jacobi, while a~2(,)\tilde{a}_{2}(\cdot,\cdot) is the additive Schwarz method associated to the (Hiptmair–Toselli) decomposition

(7.4) VΔ0(𝒯h)gradXΓ0|V+EΔ1(𝒯h)XΓ1,I|E\displaystyle\sum_{V\in\Delta_{0}(\mathcal{T}_{h})}\left.\operatorname{grad}X^{0}_{\Gamma}\right|_{\star V}+\sum_{E\in\Delta_{1}(\mathcal{T}_{h})}\left.X^{1,\mathrm{I}}_{\Gamma}\right|_{\star E}

with exact solvers. The inexact solver a~3(,)\tilde{a}_{3}(\cdot,\cdot) is a geometric multigrid V-cycle preconditioner with Hiptmair–Jacobi smoothing meaning that the smoother is the additive Schwarz method associated to the decomposition 7.4 at lowest order (XΓ0=W0X^{0}_{\Gamma}=W^{0} and XΓ1,I=W1X^{1,\mathrm{I}}_{\Gamma}=W^{1}) with exact solvers. The solver is summarized in Figure 7. The other three hybrid Hiptmair–Toselli solvers are defined analogously.

KSP: conjugate gradientsPC: symmetric multiplicative Schwarz XkX^{k}_{\mathcal{I}}: point-Jacobi on cell interiors XΓkX^{k}_{\Gamma}: additive Hiptmair–Toselli decomposition dk1YΓk1\operatorname{d}^{k-1}Y^{k-1}_{\Gamma}: interface additive patches on (k1)(k-1)-stars ZΓkZ^{k}_{\Gamma}: interface additive patches on kk-stars WkW^{k}: geometric multigrid V-cycle Relaxation: Hiptmair–Jacobi hh-coarse: Cholesky
Figure 7. Solver diagram for HT(YΓk1,ZΓk,I)+J(Xk)\mathrm{HT}({Y_{\Gamma}^{k-1}},{Z_{\Gamma}^{k,\mathrm{I}}})+\mathrm{J}({X^{k}_{\mathcal{I}}}) used for H(curl)H(\operatorname{curl}) and H(div)H(\operatorname{div}).

The iteration counts for the four Hiptmair–Toselli decompositions are displayed in the two rightmost sections of section 7.1.2. For each fixed value of α\alpha, the HT(X0,X1)\mathrm{HT}({X^{0}},{X^{1}}) and HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) solvers give nearly identical iteration counts which are robust in hh and pp. Moreover, the iteration counts are robust across values of α\alpha. For α1\alpha\geq 1, the HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) and HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) solvers also give nearly identical iteration counts that are a few iterations less than the HT(X0,X1)\mathrm{HT}({X^{0}},{X^{1}}) and HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) solvers. For α=103\alpha=10^{-3}, HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) slightly outperforms HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}), but both require substantially more iterations than for α1\alpha\geq 1 and the number of iterations decreases as the mesh is refined. Overall, for α1\alpha\geq 1, the finest splitting HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1,\mathrm{I}}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) is within 6 iterations of PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}), which is the most computationally expensive solver and requires the smallest iteration counts, while the slightly coarser splitting HT(XΓ0,XΓ1)+J(X1)\mathrm{HT}({X^{0}_{\Gamma}},{X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) requires about 3 times the number of iterations as PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) for α=103\alpha=10^{-3}.

Table 4. PCG iteration counts for the H(curl)H(\operatorname{curl}) Riesz map, where Yk=XΓk(Yk=Xk)Y^{k}=X^{k}_{\Gamma}\ (Y^{k}=X^{k}).
\csvreader

[ head to column names, head to column names prefix=MY, tabular=rlccc—, table head= PAFW0(Y1)\mathrm{PAFW}_{0}({Y^{1}})  
ll pp α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7, table foot=]results/N1curl.3d.demkowicz.afw.combined.csv \MYlevel \MYdegree \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations) \csvreader[ head to column names, head to column names prefix=MY, tabular=—ccc, table head= HT(Y0,Y1)\mathrm{HT}({Y^{0}},{Y^{1}})
α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7, table foot=]results/N1curl.3d.demkowicz.hiptmair.combined.csv \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations)

\csvreader

[ head to column names, head to column names prefix=MY, tabular=rlccc, table head= HT(Y0,Y1,I)\mathrm{HT}({Y^{0}},{Y^{1,\mathrm{I}}})
ll pp α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7, table foot=]results/N1curl.3d.demkowicz.hiptmair-red.combined.csv \MYlevel \MYdegree \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations) \csvreader[ head to column names, head to column names prefix=MY, tabular=—ccc—, table head= HT(Y0,Y1)\mathrm{HT}({Y^{0}},{Y^{1}})  
α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7, table foot=]results/N1curl.3d.demkowicz.hiptmair.combined.csv \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations)

7.1.3. H(div)H(\operatorname{div}) solvers

We again first consider the vertex patch subspace decompositions PAFW0(X3)\mathrm{PAFW}_{0}({X^{3}}) and PAFW0(XΓ3)+J(X3)\mathrm{PAFW}_{0}({X_{\Gamma}^{3}})+\mathrm{J}({X_{\mathcal{I}}^{3}}) with the same symmetric hybrid Schwarz solver described in Figure 6. The iteration counts for these two solvers with p{3,5,7}p\in\{3,5,7\} are displayed in the upper left quadrant of section 7.1.3. After one or two levels of refinement, PAFW0(XΓ3)+J(X3)\mathrm{PAFW}_{0}({X_{\Gamma}^{3}})+\mathrm{J}({X_{\mathcal{I}}^{3}}) requires about double the number of iterations as PAFW0(X3)\mathrm{PAFW}_{0}({X^{3}}), but all iteration counts appear to be robust in α\alpha, hh, and pp. For the edge patch subspace decompositions PAFW1(X3)\mathrm{PAFW}_{1}({X^{3}}) and PAFW1(XΓ3)+J(X3)\mathrm{PAFW}_{1}({X_{\Gamma}^{3}})+\mathrm{J}({X_{\mathcal{I}}^{3}}), described in Figure 6 and displayed in the upper right of section 7.1.3, the iteration counts for α1\alpha\geq 1 differ by at most 2 and are insensitive to α\alpha and pp. For α=103\alpha=10^{-3}, the iteration counts decrease as the mesh is refined and on the finest mesh, the iteration counts are also within 2 and appear to be insensitive to pp.

The iteration counts for the Hiptmair–Toselli decompositions HT(X2,X3)\mathrm{HT}({X^{2}},{X^{3}}),
HT(XΓ2,XΓ3)+J(X3)\mathrm{HT}({X_{\Gamma}^{2}},{X_{\Gamma}^{3}})+\mathrm{J}({X_{\mathcal{I}}^{3}}), HT(X2,I,X3,I)\mathrm{HT}({X^{2,\mathrm{I}}},{X^{3,\mathrm{I}}}), and HT(XΓ2,I,XΓ3,I)+J(X3)\mathrm{HT}({X_{\Gamma}^{2,\mathrm{I}}},{X_{\Gamma}^{3,\mathrm{I}}})+\mathrm{J}({X_{\mathcal{I}}^{3}}), described in Figure 7, are displayed in the bottom half of section 7.1.3. Note that the results are similar to the edge patch decompositions, with the only exception being that the type-I/type-II splitting HT(X2,I,X3,I)\mathrm{HT}({X^{2,\mathrm{I}}},{X^{3,\mathrm{I}}}) performs worse than HT(X2,X3)\mathrm{HT}({X^{2}},{X^{3}}) on coarser meshes. Overall, PAFW0(X3)\mathrm{PAFW}_{0}({X^{3}}) and PAFW0(XΓ3)+J(X3)\mathrm{PAFW}_{0}({X_{\Gamma}^{3}})+\mathrm{J}({X_{\mathcal{I}}^{3}}) perform the best, but all solvers with the interior-interface split perform nearly identically for α1\alpha\geq 1 and on fine enough meshes for α=103\alpha=10^{-3}.

Table 5. PCG iteration counts for the H(div)H(\operatorname{div}) Riesz map, where Yk=XΓk(Yk=Xk)Y^{k}=X^{k}_{\Gamma}\ (Y^{k}=X^{k}).
\csvreader

[ head to column names, head to column names prefix=MY, tabular=rlccc—, table head= PAFW0(Y2)\mathrm{PAFW}_{0}({Y^{2}})  
ll pp α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7 ]results/N1div.3d.demkowicz.afw0.combined.csv \MYlevel \MYdegree \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations) \csvreader[ head to column names, head to column names prefix=MY, tabular=—ccc, table head= PAFW1(Y2)\mathrm{PAFW}_{1}({Y^{2}})
α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7 ]results/N1div.3d.demkowicz.afw.combined.csv \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations)

\csvreader

[ head to column names, head to column names prefix=MY, tabular=rlccc—, table head= HT(Y1,Y2)\mathrm{HT}({Y^{1}},{Y^{2}})  
ll pp α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7, table foot=]results/N1div.3d.demkowicz.hiptmair.combined.csv \MYlevel \MYdegree \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations) \csvreader[ head to column names, head to column names prefix=MY, tabular=—ccc, table head= HT(Y1,I,Y2,I)\mathrm{HT}({Y^{1,\mathrm{I}}},{Y^{2,\mathrm{I}}})
α=103\alpha=10^{3} α=1\alpha=1 α=103\alpha=10^{-3}
, late after line=
, filter ifthen=\equal\MYdegree3 \equal\MYdegree5 \equal\MYdegree7, table foot=]results/N1div.3d.demkowicz.hiptmair-red.combined.csv \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \MYFCiterations (\MYCiterations)

7.1.4. Summary remarks

For all of the Riesz maps considered, each space decomposition with and without the interior-interface splitting have the same behavior as h0h\to 0, pp\to\infty, and h2β/α0h^{2}\beta/\alpha\to 0, which is consistent with Theorem 6.1. The degradation of the splitting as h2β/αh^{2}\beta/\alpha\to\infty present in Theorem 6.1 is confirmed by a jump in iteration counts for l=0l=0 and α=103\alpha=10^{-3} between PAFW0(X1)\mathrm{PAFW}_{0}({X^{1}}) and PAFW0(XΓ1)+J(X1)\mathrm{PAFW}_{0}({X^{1}_{\Gamma}})+\mathrm{J}({X^{1}_{\mathcal{I}}}), between PAFW0(X2)\mathrm{PAFW}_{0}({X^{2}}) and PAFW0(XΓ2)+J(X2)\mathrm{PAFW}_{0}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}), between PAFW1(X2)\mathrm{PAFW}_{1}({X^{2}}) and PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}), and between HT(X1,X2)\mathrm{HT}({X^{1}},{X^{2}}) and HT(XΓ1,XΓ2)+J(X2)\mathrm{HT}({X^{1}_{\Gamma}},{X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}). That the remaining solvers appear to be unaffected by h2β/αh^{2}\beta/\alpha warrants further theoretical investigation. Also note that the effect of changing shape regularity is observed as there is a slight increase in the iteration counts after one level of mesh refinement. Moreover, Lemma 6.3 seems to partially capture the lack of pp-robustness of the type-I/type-II splitting in as observed in HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) and HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X_{\Gamma}^{0}},{X_{\Gamma}^{1,\mathrm{I}}})+\mathrm{J}({X_{\mathcal{I}}^{1}}) when α=103\alpha=10^{-3}. The apparent pp-robustness of all the Type-I/Type-II splittings for α1\alpha\geq 1 and for HT(X1,I,X2,I)\mathrm{HT}({X^{1,\mathrm{I}}},{X^{2,\mathrm{I}}}) for α=103\alpha=10^{-3} is not explained by our current theory.

We again emphasize that the cost of one iteration for the standard solvers, which include all interior degrees of freedom in the patch solves, is significantly more expensive than the solvers with the interior-interface splittings. In the case h2β/α1h^{2}\beta/\alpha\leq 1, all of the iteration counts observed are nearly identical between solvers with and without the interior-interface splitting. Moreover, all solvers with the interior-interface splitting considered for a particular Riesz map gave nearly identical iteration counts. Thus, in the case that h2β/α1h^{2}\beta/\alpha\leq 1, we recommend solvers for each of the Riesz maps as follows. For H(grad)H(\operatorname{grad}), there is only one choice, PAFW0(XΓ0)+J(X0)\mathrm{PAFW}_{0}({X^{0}_{\Gamma}})+\mathrm{J}({X^{0}_{\mathcal{I}}}). For H(curl)H(\operatorname{curl}), we take the finest splitting HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X_{\Gamma}^{0}},{X_{\Gamma}^{1,\mathrm{I}}})+\mathrm{J}({X^{1}_{\mathcal{I}}}). For H(div)H(\operatorname{div}), we take PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) because this decomposition involves patch problems that are only marginally more expensive than the patch problems in the finest splitting HT(X1,I,X2,I)+J(X2)\mathrm{HT}({X^{1,\mathrm{I}}},{X^{2,\mathrm{I}}})+\mathrm{J}({X^{2}_{\mathcal{I}}}), as discussed in section 6.3, but avoids the need for an auxiliary space. We will use these decompositions in the next section to precondition the Hodge–Laplace problem.

7.2. Hodge–Laplacians

The Riesz maps are very useful in the construction of block preconditioner for coupled systems of partial differential equations [38]. Here, we employ block diagonal preconditioners based on our developed space decomposition methods for the Riesz maps to solve Hodge–Laplacians [6] on the Fichera corner Ω=(0,1)3(0.5,1)3\Omega=(0,1)^{3}\setminus(0.5,1)^{3} with ΓN=Ω\Gamma_{N}=\partial\Omega as a prototypical example. For k1:3k\in 1:3, the problem is to find (σ,u)Xk1×Xk(\sigma,u)\in X^{k-1}\times X^{k} such that

(7.5a) (σ,τ)+(u,dk1τ)\displaystyle-(\sigma,\tau)+(u,\operatorname{d}^{k-1}\tau) =0\displaystyle=0\qquad τXk1\displaystyle\forall\tau\in X^{k-1}
(7.5b) (dk1σ,v)+(dku,dkv)\displaystyle(\operatorname{d}^{k-1}\sigma,v)+(\operatorname{d}^{k}u,\operatorname{d}^{k}v) =F(v)\displaystyle=F(v)\qquad vXk,\displaystyle\forall v\in X^{k},

where we recall that d3:=0d^{3}:=0. The cases k1:2k\in 1:2 correspond to a mixed formulation of the vector Poisson problem, while the case k=3k=3 yields the mixed formulation of the scalar Poisson problem in H(div)×L2H(\operatorname{div})\times L^{2}. We consider a single unstructured mesh, pictured in Figure 5(b). The mesh is refined towards the re-entrant vertex and the edges abutting the vertex, and the diameter of the refined tetrahedra are about 1/81/8 the diameter of the largest tetrahedra. We prescribe a constant source term F(v)=(f,v)F(v)=(f,v) for a constant f=1f=1 or f=(1,1,1)f=(1,1,1)^{\top}.

Since problem 7.5 is well-posed in H(dk1)×H(dk)H(\operatorname{d}^{k-1})\times H(\operatorname{d}^{k}) [6], we consider the following block diagonal (weighted) Riesz map preconditioner:

(7.6) (σ,τ)+(dk1σ,γdk1τ)+(u,γ1v)+(dku,dkv).(\sigma,\tau)+(\operatorname{d}^{k-1}\sigma,\gamma\operatorname{d}^{k-1}\tau)+(u,\gamma^{-1}v)+(\operatorname{d}^{k}u,\operatorname{d}^{k}v).

The case γ=1\gamma=1 corresponds to standard operator preconditioning [38], while k=3k=3 and γ1\gamma\gg 1 corresponds to augmented Lagrangian preconditioning [23, 27]. For k1:2k\in 1:2 and γ1\gamma\neq 1, the weighted Riesz map preconditioner 7.6 appears to be novel. We show in Theorem E.1 that if one uses 7.6 to precondition 7.5, then all of the eigenvalues of the preconditioned system converge to ±1\pm 1 as γ\gamma\to\infty.

Of course, the preconditioner 7.6 is too expensive to use directly, so we replace each of the weighted Reisz maps with the hybrid Schwarz method preconditioners we used above. Since we only encounter the case β/α1\beta/\alpha\leq 1, we restrict the discussion to the subspace decompositions we recommend in section 7.1.4 and the corresponding decompositions that do not split the interior-interface degrees of freedom:

  • CGp\mathrm{CG}_{p}: PAFW0(XΓ0)+J(X0)\mathrm{PAFW}_{0}({X^{0}_{\Gamma}})+\mathrm{J}({X^{0}_{\mathcal{I}}}) (and PAFW0(X0)\mathrm{PAFW}_{0}({X^{0}})),

  • Nedp1\mathrm{Ned}^{1}_{p}: HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X_{\Gamma}^{0}},{X_{\Gamma}^{1,\mathrm{I}}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) (and HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}})),

  • RTp\mathrm{RT}_{p}: PAFW1(XΓ2)+J(X2)\mathrm{PAFW}_{1}({X^{2}_{\Gamma}})+\mathrm{J}({X^{2}_{\mathcal{I}}}) (and PAFW1(X2)\mathrm{PAFW}_{1}({X^{2}})),

  • DGp1\mathrm{DG}_{p-1}: J(X3)\mathrm{J}({X^{3}}) (and J(X3)\mathrm{J}({X^{3}})), an exact solver as the basis is orthogonal.

Table 6. Preconditioned MINRES iteration counts to solve the Hodge Laplacians.
\csvreader

[ head to column names, head to column names prefix=MY, tabular=r@  c@  c@  c@  c—, table head= CGp×Nedp1\mathrm{CG}_{p}\times\mathrm{Ned}^{1}_{p}  
pp γ=1\gamma=1 γ=103\gamma=10^{3}
, table foot=]results/hodge.N1curl.3d.demkowicz.combined.csv \MYdegree \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \csvreader[ head to column names, head to column names prefix=MY, tabular=—c@  c@  c—, table head= Nedp1×RTp\mathrm{Ned}^{1}_{p}\times\mathrm{RT}_{p}  
γ=1\gamma=1 γ=103\gamma=10^{3}
, table foot=]results/hodge.N1div.3d.demkowicz.combined.csv \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations) \csvreader[ head to column names, head to column names prefix=MY, tabular=—c@  c@  c, table head= RTp×DGp1\mathrm{RT}_{p}\times\mathrm{DG}_{p-1}^{\phantom{1}}
γ=1\gamma=1 γ=103\gamma=10^{3}
, filter ifthen=/\equal\MYdegree7, table foot=]results/hodge.DG.3d.demkowicz.combined.csv \MYFAiterations (\MYAiterations) \MYFBiterations (\MYBiterations)

The preconditioned MINRES iteration counts for p1:6p\in 1:6 and γ{1,103}\gamma\in\{1,10^{3}\} with a relative tolerance of 10810^{-8} are displayed in Section 7.2. First note that for p=1p=1, the solver corresponds to using the exact weighted Riesz map preconditioner 7.6 as the multigrid V-cycle preconditioner reduces to an exact solver. Here, we observe that taking γ=103\gamma=10^{3} reduces the iteration counts from 5 or 6 to 3. For p2p\geq 2, we generally see that the solvers are robust in pp and that taking γ\gamma large reduces the iteration counts at no extra expense in the solver. The effect is minimal for the H(curl)×H(div)H(\operatorname{curl})\times H(\operatorname{div}) Hodge–Laplacian for high-order Schwarz solvers owing to difficulty of solving the individual Riesz maps. For example, one would expect to apply the Schwarz solvers 3-4 times the iteration counts for a single Riesz map solve if instead, one used the Schwarz solvers as an inner iteration to solve each block of the 7.6. For the H(curl)×H(div)H(\operatorname{curl})\times H(\operatorname{div}) Hodge–Laplacian, this amounts to about 70-100 iterations, which is in line with our observed iteration counts. We also note that the interior-interface splitting again gives comparable iteration counts within 4 for the H(curl)×H(div)H(\operatorname{curl})\times H(\operatorname{div}) and H(div)×L2H(\operatorname{div})\times L^{2} Hodge–Laplacians and within 16 for the H(grad)×H(curl)H(\operatorname{grad})\times H(\operatorname{curl}) Hodge–Laplacian.

Appendix A Bubble complexes and trace decomposition

Lemma A.1.

The complex 3.6 is exact for all SΔl(T^)S\in\Delta_{l}(\hat{T}) with lk+1:dl\in k+1:d.

Proof.

The case that S=T^S=\hat{T} follows from the exactness of the usual finite element de Rham complex with homogeneous boundary conditions (see e.g. [35, Corollary 2]).

The only remaining complex to verify is the case that SΔ2(T^)S\in\Delta_{2}(\hat{T}). One can readily verify that X̊p0(S)=Pp(S)H01(S)\mathring{X}_{p}^{0}(S)=\mathrm{P}_{p}(S)\cap H^{1}_{0}(S) and

(A.1) X̊p1(S)={(Pp1(S)2+Pp1(S)(ΠS𝐱))H0(curl;S)if Xp1(T^)=Nedp1(T^),Pp(S)2H0(curl;S)if Xp1(T^)=Nedp2(T^),\displaystyle\mathring{X}_{p}^{1}(S)=\begin{cases}(\mathrm{P}_{p-1}(S)^{2}+\mathrm{P}_{p-1}(S)(\Pi_{S}\mathbf{x})^{\perp})\cap H_{0}(\operatorname{curl}{};S)&\text{if }X^{1}_{p}(\hat{T})=\mathrm{Ned}^{1}_{p}(\hat{T}),\\ \mathrm{P}_{p}(S)^{2}\cap H_{0}(\operatorname{curl}{};S)&\text{if }X^{1}_{p}(\hat{T})=\mathrm{Ned}^{2}_{p}(\hat{T}),\end{cases}

which is the usual 2D Nedéléc space of first or second kind with vanishing tangential trace. The exactness of 3.6 again follows from [35, Corollary 2]. ∎

Lemma A.2.

The decomposition 3.9 holds.

Proof.

Let SΔk(T^)S\in\Delta_{k}(\hat{T}). Direct calculation revels that trSkWk(T^)=\operatorname{tr}^{k}_{S}W^{k}(\hat{T})=\mathbb{R}. Define τ{0,1}\tau\in\{0,1\} as follows: Let τ=1\tau=1 if k=0k=0 or Xk(T^)X^{k}(\hat{T}) is of the first kind and τ=0\tau=0 otherwise. Then, direct calculation shows that

trSkXk(T^)=Ppτ(S)SΔk(T^).\displaystyle\operatorname{tr}^{k}_{S}X^{k}(\hat{T})=\mathrm{P}_{p-\tau}(S)\qquad\operatorname{\forall}S\in\Delta_{k}(\hat{T}).

If k=1k=1, then X̊0(S)=𝒫p+1τ(S)H01(S)\mathring{X}^{0}(S)=\mathcal{P}_{p+1-\tau}(S)\cap H^{1}_{0}(S), while if k=2k=2, then X̊1(S)\mathring{X}^{1}(S) is given by A.1 with pp replaced by p+1τp+1-\tau. [35, Corollary 2] then implies that the following sequence is exact

X̊k1(S){\mathring{X}^{k-1}(S)}Ppτ(S)L02(S){\mathrm{P}_{p-\tau}(S)\cap L^{2}_{0}(S)}0,{0,}dSk1\scriptstyle{\operatorname{d}^{k-1}_{S}}

and the result now follows. ∎

Appendix B Energy stable splittings

By construction, the space XkX^{k} admits direct sum decompositions based on splitting the interior and interface basis functions and based on splitting the type-I\mathrm{I} and type-II\mathrm{II} basis functions:

(B.1) Xk=XΓkXkandXk=Xk,IXk,II.\displaystyle X^{k}=X^{k}_{\Gamma}\oplus X^{k}_{\mathcal{I}}\quad\text{and}\quad X^{k}=X^{k,\mathrm{I}}\oplus X^{k,\mathrm{II}}.

We now examine the energy stability of these two decompositions. Given a domain ω\omega, let

(u,v)H(dk;ω):=(u,v)ω+(dku,dkv)ωanduH(dk;ω)2:=(u,u)H(dk;ω).\displaystyle(u,v)_{H(\operatorname{d}^{k};\omega)}:=(u,v)_{\omega}+(\operatorname{d}^{k}u,\operatorname{d}^{k}v)_{\omega}\quad\text{and}\quad\|u\|_{H(\operatorname{d}^{k};\omega)}^{2}:=(u,u)_{H(\operatorname{d}^{k};\omega)}.

B.1. H(dk)H(\operatorname{d}^{k})-stability on the reference cell

We begin by analyzing the stability of the decompositions on the reference cell:

(B.2) Xk(T^)=XΓk(T^)Xk(T^)andXk(T^)=Xk,I(T^)Xk,II(T^),\displaystyle X^{k}(\hat{T})=X^{k}_{\Gamma}(\hat{T})\oplus X^{k}_{\mathcal{I}}(\hat{T})\quad\text{and}\quad X^{k}(\hat{T})=X^{k,\mathrm{I}}(\hat{T})\oplus X^{k,\mathrm{II}}(\hat{T}),

where XΓk(T^)X^{k}_{\Gamma}(\hat{T}), etc. denotes the space analogous to 5.1, 5.2 and 6.1 on the reference cell. In particular, given uXk(T^)u\in X^{k}(\hat{T}), there exist unique uΓXΓk(T^)u_{\Gamma}\in X^{k}_{\Gamma}(\hat{T}), uXk(T^)u_{\mathcal{I}}\in X^{k}_{\mathcal{I}}(\hat{T}), uIXk,I(T^)u_{\mathrm{I}}\in X^{k,\mathrm{I}}(\hat{T}), and uIIXk,II(T^)u_{\mathrm{II}}\in X^{k,\mathrm{II}}(\hat{T}) such that

(B.3) u=uΓ+uandu=uI+uII,\displaystyle u=u_{\Gamma}+u_{\mathcal{I}}\quad\text{and}\quad u=u_{\mathrm{I}}+u_{\mathrm{II}},

and the mappings u(uΓ,u)u\mapsto(u_{\Gamma},u_{\mathcal{I}}) and u(uI,uII)u\mapsto(u_{\mathrm{I}},u_{\mathrm{II}}) are linear. The first result shows that the interior-interface decomposition is uniformly stable with respect to the polynomial degree.

Theorem B.1.

There exists a constant CC depending only on the dimension dd such that for all uXk(T^)u\in X^{k}(\hat{T}), there holds

(B.4) 12uT^2uΓT^2+uT^2CuH(dk;T^)2\displaystyle\frac{1}{2}\|u\|_{\hat{T}}^{2}\leq\|u_{\Gamma}\|_{\hat{T}}^{2}+\|u_{\mathcal{I}}\|_{\hat{T}}^{2}\leq C\|u\|_{H(\operatorname{d}^{k};\hat{T})}^{2}

and

(B.5) dkuT^2=dkuΓT^2+dkuT^2.\displaystyle\|\operatorname{d}^{k}u\|_{\hat{T}}^{2}=\|\operatorname{d}^{k}u_{\Gamma}\|_{\hat{T}}^{2}+\|\operatorname{d}^{k}u_{\mathcal{I}}\|_{\hat{T}}^{2}.
Proof.

We prove the case d=3d=3; the case d=2d=2 follows from analogous arguments.

Step 1: dk\operatorname{d}^{k} stability. The degrees of freedom 3.14b and 3.14c show that uu_{\mathcal{I}} satisfies

(B.6a) (u,dk1y)T^\displaystyle(u_{\mathcal{I}},\operatorname{d}^{k-1}y)_{\hat{T}} =(u,dk1y)T^\displaystyle=(u,\operatorname{d}^{k-1}y)_{\hat{T}}\qquad yX̊k1(T^),\displaystyle\operatorname{\forall}y\in\mathring{X}^{k-1}(\hat{T}),
(B.6b) (dku,dkv)T^\displaystyle(\operatorname{d}^{k}u_{\mathcal{I}},\operatorname{d}^{k}v)_{\hat{T}} =(dku,dkv)T^\displaystyle=(\operatorname{d}^{k}u,\operatorname{d}^{k}v)_{\hat{T}}\qquad vX̊k(T^),\displaystyle\operatorname{\forall}v\in\mathring{X}^{k}(\hat{T}),

and in particular dku\operatorname{d}^{k}u_{\mathcal{I}} is the L2(T^)L^{2}(\hat{T})-projection of dku\operatorname{d}^{k}u onto dkX̊k(T^)\operatorname{d}^{k}\mathring{X}^{k}(\hat{T}). Equation B.5 immediately follows.

Step 2: L2L^{2}-stability. Suppose first that k=0k=0. Then, Poincaré’s inequality gives uT^CuH(dk;T^)\|u_{\mathcal{I}}\|_{\hat{T}}\leq C\|u\|_{H(\operatorname{d}^{k};\hat{T})}. Now suppose that k{1,2}k\in\{1,2\}. Then, conditions B.6 ensure that there exists qdkX̊k(T^)q\in\operatorname{d}^{k}\mathring{X}^{k}(\hat{T}) such that

(B.7a) (u,v)T^+(dkv,q)T^\displaystyle(u_{\mathcal{I}},v)_{\hat{T}}+(\operatorname{d}^{k}v,q)_{\hat{T}} =(u,v)T^\displaystyle=(u,v)_{\hat{T}}\qquad vX̊k(T^),\displaystyle\operatorname{\forall}v\in\mathring{X}^{k}(\hat{T}),
(B.7b) (dku,r)T^\displaystyle(\operatorname{d}^{k}u_{\mathcal{I}},r)_{\hat{T}} =(dku,r)T^\displaystyle=(\operatorname{d}^{k}u,r)_{\hat{T}}\qquad rdkX̊k(T^).\displaystyle\operatorname{\forall}r\in\operatorname{d}^{k}\mathring{X}^{k}(\hat{T}).

Applying standard estimates for saddle point systems gives

uH(dk;T^)C(βk)2uH(dk;T^),\displaystyle\|u_{\mathcal{I}}\|_{H(\operatorname{d}^{k};\hat{T})}\leq C(\beta^{k})^{-2}\|u\|_{H(\operatorname{d}^{k};\hat{T})},

where βk\beta^{k} is the inf-sup constant

βk:=inf0rdkX̊k(T^)sup0vX̊k(T^)(dkv,r)T^vH(dk,T^)r.\displaystyle\beta^{k}:=\inf_{0\neq r\in\operatorname{d}^{k}\mathring{X}^{k}(\hat{T})}\sup_{0\neq v\in\mathring{X}^{k}(\hat{T})}\frac{(\operatorname{d}^{k}v,r)_{\hat{T}}}{\|v\|_{H(\operatorname{d}^{k},\hat{T})}\|r\|}.

We now show that βkβ0>0\beta^{k}\geq\beta_{0}>0 for some β0\beta_{0} independent of pp. Let rdkX̊k(T^)r\in\operatorname{d}^{k}\mathring{X}^{k}(\hat{T}) be given. Thanks to [20, Theorem 5.1], there exists vX̊k(T^)v\in\mathring{X}^{k}(\hat{T}) such that dkv=r\operatorname{d}^{k}v=r and vT^CrT^\|v\|_{\hat{T}}\leq C\|r\|_{\hat{T}}, where CC is independent of pp and rr. Then, vH(dk,T^)CrT^\|v\|_{H(\operatorname{d}^{k},\hat{T})}\leq C\|r\|_{\hat{T}}, and so βkC1>0\beta^{k}\geq C^{-1}>0. Consequently, for k0:2k\in 0:2, there holds uH(dk,T^)CuH(dk,T^)\|u_{\mathcal{I}}\|_{H(\operatorname{d}^{k},\hat{T})}\leq C\|u\|_{H(\operatorname{d}^{k},\hat{T})}. Inequality B.4 then follows. ∎

We now turn to the type-I\mathrm{I}/type-II\mathrm{II} decomposition. Let {ujk,ωj,pk,I/II:j1:dimXk(T^)}Xk(T^)×\{u_{j}^{k},\omega_{j,p}^{k,\mathrm{I}/\mathrm{II}}:j\in 1:\dim X^{k}(\hat{T})\}\subset X^{k}(\hat{T})\times\mathbb{R} denote the eigenfunctions and eigenvalues (in increasing order) of the following eigenvalue problem:

(B.8a) (uik,ujk)T^\displaystyle(u_{i}^{k},u_{j}^{k})_{\hat{T}} =ωi,pk,I/IIδij\displaystyle=\omega_{i,p}^{k,\mathrm{I}/\mathrm{II}}\delta_{ij}\qquad i,j1:dimXk(T^),\displaystyle\operatorname{\forall}i,j\in 1:\dim X^{k}(\hat{T}),
(B.8b) (ui,Ik,uj,Ik)T^+(ui,IIk,uj,IIk)T^\displaystyle(u_{i,\mathrm{I}}^{k},u^{k}_{j,\mathrm{I}})_{\hat{T}}+(u_{i,\mathrm{II}}^{k},u_{j,\mathrm{II}}^{k})_{\hat{T}} =δij\displaystyle=\delta_{ij}\qquad i,j1:dimXk(T^).\displaystyle\operatorname{\forall}i,j\in 1:\dim X^{k}(\hat{T}).

Then, the following result shows that the L2(K^)L^{2}(\hat{K}) stability of the type-I\mathrm{I}/type-II\mathrm{II} decomposition is determined by the minimal eigenvalue ω1,pk,I/II\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}.

Lemma B.2.

For all uXk(T^)u\in X^{k}(\hat{T}), dku=dkuI\operatorname{d}^{k}u=\operatorname{d}^{k}u_{\mathrm{I}} and there holds

(B.9) 12uT^2uIT^2+uIIT^21ω1,pk,I/IIuT^2.\displaystyle\frac{1}{2}\|u\|_{\hat{T}}^{2}\leq\|u_{\mathrm{I}}\|_{\hat{T}}^{2}+\|u_{\mathrm{II}}\|_{\hat{T}}^{2}\leq\frac{1}{\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}}\|u\|_{\hat{T}}^{2}.
Proof.

Since Xk,II(T^)ker(dk;Xk(T^))X^{k,\mathrm{II}}(\hat{T})\subset\ker(\operatorname{d}^{k};X^{k}(\hat{T})) thanks to Lemma 4.3, dku=dkuI\operatorname{d}^{k}u=\operatorname{d}^{k}u_{\mathrm{I}}. Moreover, let cjc_{j}\in\mathbb{R} be such that u=j=1dimXk(T^)cjujku=\sum_{j=1}^{\dim X^{k}(\hat{T})}c_{j}u_{j}^{k}. Then, applying B.8 gives

uIT^2+uIIT^2=j=1dimXk(T^)cj21ω1,pk,I/IIj=1dimXk(T^)cj2ωj,pk,I/II=uT^2.\displaystyle\|u_{\mathrm{I}}\|_{\hat{T}}^{2}+\|u_{\mathrm{II}}\|_{\hat{T}}^{2}=\sum_{j=1}^{\dim X^{k}(\hat{T})}c_{j}^{2}\leq\frac{1}{\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}}\sum_{j=1}^{\dim X^{k}(\hat{T})}c_{j}^{2}\omega_{j,p}^{k,\mathrm{I}/\mathrm{II}}=\|u\|_{\hat{T}}^{2}.

B.2. Energy stability on physical cell

Given an element in T𝒯hT\in\mathcal{T}_{h} and uXku\in X^{k}, define

aTk(u,v):=(u,βv)T+(dku,αdkv)Tandua,T2:=aTk(u,u)u,vXk(T).\displaystyle a_{T}^{k}(u,v):=(u,\beta v)_{T}+(\operatorname{d}^{k}u,\alpha\operatorname{d}^{k}v)_{T}\ \ \text{and}\ \ \|u\|_{a,T}^{2}:=a_{T}^{k}(u,u)\qquad\operatorname{\forall}u,v\in X^{k}(T).

We again use the notation uΓu_{\Gamma}, uu_{\mathcal{I}}, uIu_{\mathrm{I}}, and uIIu_{\mathrm{II}} to denote the decomposition of uXku\in X^{k} into its components in the decompositions B.1. The next result shows that these decompositions are energy stable.

Corollary B.3.

There exist constants C1,C2>0C_{1},C_{2}>0 depending only on shape regularity, dd, and kk such that for all uXpk(𝒯h)u\in X^{k}_{p}(\mathcal{T}_{h}) and T𝒯hT\in\mathcal{T}_{h}, there holds

(B.10) 12ua,T2\displaystyle\frac{1}{2}\|u\|_{a,T}^{2} uΓa,T2+ua,T2\displaystyle\leq\|u_{\Gamma}\|_{a,T}^{2}+\|u_{\mathcal{I}}\|_{a,T}^{2} C1max{1,βαhT2}ua,T2,\displaystyle\leq C_{1}\max\left\{1,\frac{\beta}{\alpha}h_{T}^{2}\right\}\|u\|_{a,T}^{2},
(B.11) 12ua,T2\displaystyle\frac{1}{2}\|u\|_{a,T}^{2} uIa,T2+uIIa,T2\displaystyle\leq\|u_{\mathrm{I}}\|_{a,T}^{2}+\|u_{\mathrm{II}}\|_{a,T}^{2} C2ω1,pk,I/IIua,T2,\displaystyle\leq\frac{C_{2}}{\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}}\|u\|_{a,T}^{2},

where ω1,pk,I/II\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}} is defined in B.8.

Proof.

Let T𝒯hT\in\mathcal{T}_{h} and u^:=(Tk)1u\hat{u}:=(\mathcal{F}_{T}^{k})^{-1}u so that dku=Tk+1(dku^)\operatorname{d}^{k}u=\mathcal{F}_{T}^{k+1}(\operatorname{d}^{k}\hat{u}).

Step 1: H(dk;T)H(\operatorname{d}^{k};T)-stability of uu_{\mathcal{I}}. Note that Tk\mathcal{F}^{k}_{T} is of the form Tkv^=GTkv^FT1\mathcal{F}^{k}_{T}\hat{v}=G_{T}^{k}\hat{v}\circ F_{T}^{-1} for some constant or matrix GTkG^{k}_{T}. Thus, there holds

uT2=|detJT|GTkΠT^,ku^T^2\displaystyle\|u_{\mathcal{I}}\|_{T}^{2}=|\det J_{T}|\cdot\|G^{k}_{T}\Pi_{\hat{T},\mathcal{I}}^{k}\hat{u}\|_{\hat{T}}^{2} |detJT||GTk|2ΠT^,ku^T^2\displaystyle\leq|\det J_{T}|\cdot|G^{k}_{T}|^{2}\|\Pi_{\hat{T},\mathcal{I}}^{k}\hat{u}\|_{\hat{T}}^{2}
|detJT||GTk|2(u^T^2+dku^T^2),\displaystyle\leq|\det J_{T}|\cdot|G^{k}_{T}|^{2}\left(\|\hat{u}\|_{\hat{T}}^{2}+\|\operatorname{d}^{k}\hat{u}\|_{\hat{T}}\|^{2}\right),

where |GTk||G^{k}_{T}| also denotes the spectral norm on matrices. Now,

u^T^2=|detJT1|(GTk)1uT2|detJT|1|(GTk)1|2uT2\displaystyle\|\hat{u}\|_{\hat{T}}^{2}=|\det J_{T}^{-1}|\cdot\|(G^{k}_{T})^{-1}u\|_{T}^{2}\leq|\det J_{T}|^{-1}|(G^{k}_{T})^{-1}|^{2}\|u\|_{T}^{2}

and

dku^T^2=|detJT1|(GTk+1)1dkuT2|detJT|1|(GTk+1)1|2dkuT2,\displaystyle\|\operatorname{d}^{k}\hat{u}\|_{\hat{T}}^{2}=|\det J_{T}^{-1}|\cdot\|(G^{k+1}_{T})^{-1}\operatorname{d}^{k}u\|_{T}^{2}\leq|\det J_{T}|^{-1}|(G^{k+1}_{T})^{-1}|^{2}\|\operatorname{d}^{k}u\|_{T}^{2},

and so

uT2|GTk|2(|(GTk)1|2uT2+|(GTk+1)1|2dkuT2).\displaystyle\|u_{\mathcal{I}}\|_{T}^{2}\leq|G^{k}_{T}|^{2}\left(|(G^{k}_{T})^{-1}|^{2}\|u\|_{T}^{2}+|(G^{k+1}_{T})^{-1}|^{2}\|\operatorname{d}^{k}u\|_{T}^{2}\right).

One may then verify that |GTk||(GTk)1|C|G^{k}_{T}|\cdot|(G^{k}_{T})^{-1}|\leq C and |GTk||(GTk+1)1|ChT|G^{k}_{T}|\cdot|(G^{k+1}_{T})^{-1}|\leq Ch_{T} for some C>0C>0 depending only on shape regularity, dd, and kk. Thus,

uT2C(uT2+hT2dkuT2).\displaystyle\|u_{\mathcal{I}}\|_{T}^{2}\leq C\left(\|u\|_{T}^{2}+h_{T}^{2}\|\operatorname{d}^{k}u\|_{T}^{2}\right).

Similar scaling arguments applied to B.5 show that

(B.12) dkuTCdkuT.\displaystyle\|\operatorname{d}^{k}u_{\mathcal{I}}\|_{T}\leq C\|\operatorname{d}^{k}u\|_{T}.

Step 2: B.10. We then obtain

(u,βu)T\displaystyle(u_{\mathcal{I}},\beta u_{\mathcal{I}})_{T} Cβ{(u,u)T+hT2(dku,dku)T}\displaystyle\leq C\beta\left\{(u,u)_{T}+h_{T}^{2}(\operatorname{d}^{k}u,\operatorname{d}^{k}u)_{T}\right\}
C{(u,βu)T+βαhT2(dku,αdku)T}Cmax{1,βαhT2}aTk(u,u),\displaystyle\leq C\left\{(u,\beta u)_{T}+\frac{\beta}{\alpha}h_{T}^{2}(\operatorname{d}^{k}u,\alpha\operatorname{d}^{k}u)_{T}\right\}\leq C\max\left\{1,\frac{\beta}{\alpha}h_{T}^{2}\right\}a^{k}_{T}(u,u),

and so B.12 gives

aTk(u,u)Cmax{1,βαhT2}aTk(u,u).\displaystyle a^{k}_{T}(u_{\mathcal{I}},u_{\mathcal{I}})\leq C\max\left\{1,\frac{\beta}{\alpha}h_{T}^{2}\right\}a^{k}_{T}(u,u).

Inequality B.10 now follows from the bound uΓa,T22(ua,T2+ua,T2)\|u_{\Gamma}\|_{a,T}^{2}\leq 2\left(\|u\|_{a,T}^{2}+\|u_{\mathcal{I}}\|_{a,T}^{2}\right).

Step 3: B.11. Inequality B.11 follows from Lemma B.2 using analogous arguments from Steps 1 and 2 on noting that ω1,pk,I/II2\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}\leq 2 by B.9. ∎

We may further decompose the interior component into individual basis functions, and this decomposition is stable for any choice of α\alpha and β\beta.

Lemma B.4.

Let NK,:=dimX̊k(T)N_{K,\mathcal{I}}:=\dim\mathring{X}^{k}(T) and let {ϕT,ik:i1:NK,}\{\phi_{T,i}^{k}:i\in 1:N_{K,\mathcal{I}}\} denote interior basis functions supported on TT. Then, there exists a constant C>0C>0 depending only on shape regularity such that if u=i=1NK,u,iϕT,iku_{\mathcal{I}}=\sum_{i=1}^{N_{K,\mathcal{I}}}u_{\mathcal{I},i}\phi_{T,i}^{k}, then

(B.13) C1ua,T2i=1NK,|u,i|2ϕT,ika,T2Cua,T2.\displaystyle C^{-1}\|u_{\mathcal{I}}\|_{a,T}^{2}\leq\sum_{i=1}^{N_{K,\mathcal{I}}}|u_{\mathcal{I},i}|^{2}\|\phi_{T,i}^{k}\|_{a,T}^{2}\leq C\|u_{\mathcal{I}}\|_{a,T}^{2}.
Proof.

Thanks to Lemma 4.5, B.13 holds in the case that T=T^T=\hat{T} is the reference cell and (α,β){(0,1),(1,0)}(\alpha,\beta)\in\{(0,1),(1,0)\}. A standard scaling argument similar to the one in Step 1 of the proof of Corollary B.3 shows that B.13 holds for all T𝒯hT\in\mathcal{T}_{h} and (α,β){(0,1),(1,0)}(\alpha,\beta)\in\{(0,1),(1,0)\}, where CC only depend on shape regularity. The case of general α\alpha and β\beta follows from linearity. ∎

B.3. Proof of Lemma 5.3

The result follows from Theorem 6.1 applied on to a single element with X0k=XkX_{0}^{k}=X^{k}. ∎

Appendix C Proof of Theorem 6.1

Step 1: Stable decomposition 6.6. Let uXku\in X^{k} and let Πk:XkXk\Pi^{k}_{\mathcal{I}}:X^{k}\to X^{k}_{\mathcal{I}} be defined by the rule

Πkv:=vvXk.\displaystyle\Pi^{k}_{\mathcal{I}}v:=v_{\mathcal{I}}\qquad\forall v\in X^{k}.

Note that Πk\Pi^{k}_{\mathcal{I}} is a projection operator. Thanks to 6.5, there exist umXmu_{m}\in X_{m}, m1:Mm\in 1:M, satisfying

m=1Mum=uandm=1Muma2Cstable(α,β)ua2,\displaystyle\sum_{m=1}^{M}u_{m}=u\quad\text{and}\quad\sum_{m=1}^{M}\|u_{m}\|_{a}^{2}\leq C_{\mathrm{stable}}(\alpha,\beta)\|u\|_{a}^{2},

where let a\|\cdot\|_{a} denotes the energy norm: a2:=ak(,)\|\cdot\|_{a}^{2}:=a^{k}(\cdot,\cdot). Let um,Γ:=(IΠk)umu_{m,\Gamma}:=(I-\Pi_{\mathcal{I}}^{k})u_{m} and u:=Πkuu_{\mathcal{I}}:=\Pi_{\mathcal{I}}^{k}u. Since Πk\Pi^{k}_{\mathcal{I}} is a projection, there holds

i=1Mum,Γ+u=(IΠk)i=1Mum+Πku=(IΠk)u+Πku=u,\displaystyle\sum_{i=1}^{M}u_{m,\Gamma}+u_{\mathcal{I}}=(I-\Pi_{\mathcal{I}}^{k})\sum_{i=1}^{M}u_{m}+\Pi_{\mathcal{I}}^{k}u=(I-\Pi_{\mathcal{I}}^{k})u+\Pi_{\mathcal{I}}^{k}u=u,

and applying B.10 element-by-element gives

m=1Mum,Γa2Cmax{1,βαh2}m=1Muma2CCstable(α,β)max{1,βαh2}ua2.\displaystyle\sum_{m=1}^{M}\|u_{m,\Gamma}\|_{a}^{2}\leq C\max\left\{1,\frac{\beta}{\alpha}h^{2}\right\}\sum_{m=1}^{M}\|u_{m}\|_{a}^{2}\leq CC_{\mathrm{stable}}(\alpha,\beta)\max\left\{1,\frac{\beta}{\alpha}h^{2}\right\}\|u\|_{a}^{2}.

Consequently, there holds

ua2Cmax{1,βαh2}ua2.\displaystyle\|u_{\mathcal{I}}\|_{a}^{2}\leq C\max\left\{1,\frac{\beta}{\alpha}h^{2}\right\}\|u\|_{a}^{2}.

Summing B.13 over the elements then gives

ak(u,u)+m=0Mum,Γa2CCstable(α,β)max{1,βαh2}ua2,\displaystyle a_{\mathcal{I}}^{k}(u_{\mathcal{I}},u_{\mathcal{I}})+\sum_{m=0}^{M}\|u_{m,\Gamma}\|_{a}^{2}\leq CC_{\mathrm{stable}}(\alpha,\beta)\max\left\{1,\frac{\beta}{\alpha}h^{2}\right\}\|u\|_{a}^{2},

which completes the proof of 6.6.

Step 2: Local Stability. Inequality 6.7 follows immediately from B.13. ∎

Appendix D Further discussion of HT(Xk1,I,Xk,I)\mathrm{HT}({X^{k-1,\mathrm{I}}},{X^{k,\mathrm{I}}})

We first prove Lemma 6.3 and then discuss the pp-dependence of the constant CI/IIC_{\mathrm{I}/\mathrm{II}}.

D.1. Proof of Lemma 6.3

Let uXku\in X^{k} and let u0Wku_{0}\in W^{k}, uJdk1Xk1|Ju_{J}\in\left.\operatorname{d}^{k-1}X^{k-1}\right|_{\star J}, JΔk1(𝒯h)J\in\Delta_{k-1}(\mathcal{T}_{h}), and uLXk|Lu_{L}\in\left.X^{k}\right|_{\star L}, LΔk(𝒯h)L\in\Delta_{k}(\mathcal{T}_{h}), be such that

u0+JΔk1(𝒯h)uJ+LΔk(𝒯h)uL\displaystyle u_{0}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}u_{J}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}u_{L} =u,\displaystyle=u,
u0a2+JΔk1(𝒯h)uJa2+LΔk(𝒯h)uLa2\displaystyle\|u_{0}\|_{a}^{2}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}\|u_{J}\|_{a}^{2}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}\|u_{L}\|_{a}^{2} CHTua2,\displaystyle\leq C_{\mathrm{HT}}\|u\|_{a}^{2},

where a\|\cdot\|_{a} again denotes the energy norm. Thanks to 5.7, dk1Xk1|J=dk1Xk1,I|J\left.\operatorname{d}^{k-1}X^{k-1}\right|_{\star J}=\left.\operatorname{d}^{k-1}X^{k-1,\mathrm{I}}\right|_{\star J}, and so uJdk1Xk1,I|Ju_{J}\in\left.\operatorname{d}^{k-1}X^{k-1,\mathrm{I}}\right|_{\star J} for all JΔk1(𝒯h)J\in\Delta_{k-1}(\mathcal{T}_{h}).

For each LΔk(𝒯h)L\in\Delta_{k}(\mathcal{T}_{h}), there exist unique uL,IXk,I|Lu_{L,\mathrm{I}}\in\left.X^{k,\mathrm{I}}\right|_{\star L} and uL,IIXk,II|Lu_{L,\mathrm{II}}\in\left.X^{k,\mathrm{II}}\right|_{\star L} such that uL=uL,I+uL,IIu_{L}=u_{L,\mathrm{I}}+u_{L,\mathrm{II}} thanks to 5.3 restricted to L\star L. Moreover, summing B.11 over the cells in L\star L, we obtain

(D.1) uL,Ia,L2+uL,IIa,L2\displaystyle\|u_{L,\mathrm{I}}\|_{a,\star L}^{2}+\|u_{L,\mathrm{II}}\|_{a,\star L}^{2} Cω1,pk,I/IIuLa,L2.\displaystyle\leq\frac{C}{\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}}\|u_{L}\|_{a,\star L}^{2}.

Thanks to [35, Corollary 2], the complex

Xk1|L{\left.X^{k-1}\right|_{\star L}}Xk|L{\left.X^{k}\right|_{\star L}}dkXk|L{\operatorname{d}^{k}\left.X^{k}\right|_{\star L}}0{0}dk1\scriptstyle{\operatorname{d}^{k-1}}dk\scriptstyle{\operatorname{d}^{k}}

is exact since L\star L is contractible, and so uL,IIdk1Xk1|L=dk1Xk1,I|Lu_{L,\mathrm{II}}\in\operatorname{d}^{k-1}\left.X^{k-1}\right|_{\star L}=\operatorname{d}^{k-1}\left.X^{k-1,\mathrm{I}}\right|_{\star L}.

For JΔk1(𝒯h)J\in\Delta_{k-1}(\mathcal{T}_{h}), let Δk(J)\Delta_{k}(J) denote that subsimplices of JJ of dimension kk and define

u~J=uJ+1|Δk(J)|LΔk(J)uL,II.\displaystyle\tilde{u}_{J}=u_{J}+\frac{1}{|\Delta_{k}(J)|}\sum_{L\in\Delta_{k}(J)}u_{L,\mathrm{II}}.

Then, u~Jdk1Xk1,I|L\tilde{u}_{J}\in\operatorname{d}^{k-1}\left.X^{k-1,\mathrm{I}}\right|_{\star L} and

u~Ja22(uJa2+1|Δk(J)|LΔk(J)uL,IIa2).\displaystyle\|\tilde{u}_{J}\|_{a}^{2}\leq 2\left(\|u_{J}\|_{a}^{2}+\frac{1}{|\Delta_{k}(J)|}\sum_{L\in\Delta_{k}(J)}\|u_{L,\mathrm{II}}\|_{a}^{2}\right).

Consequently, there holds

u0+JΔk1(𝒯h)u~J+LΔk(𝒯h)uL,I=u0+JΔk1(𝒯h)uJ+LΔk(𝒯h)uL=u,\displaystyle u_{0}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}\tilde{u}_{J}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}u_{L,\mathrm{I}}=u_{0}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}u_{J}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}u_{L}=u,

and applying D.1 gives

u0a2+JΔk1(𝒯h)u~Ja2+LΔk(𝒯h)uL,Ia2\displaystyle\|u_{0}\|_{a}^{2}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}\|\tilde{u}_{J}\|_{a}^{2}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}\|u_{L,\mathrm{I}}\|_{a}^{2}
2(u0a2+JΔk1(𝒯h)u~Ja2+LΔk(𝒯h)(uL,Ia2+uL,IIa2))\displaystyle\qquad\leq 2\left(\|u_{0}\|_{a}^{2}+\sum_{J\in\Delta_{k-1}(\mathcal{T}_{h})}\|\tilde{u}_{J}\|_{a}^{2}+\sum_{L\in\Delta_{k}(\mathcal{T}_{h})}\left(\|u_{L,\mathrm{I}}\|_{a}^{2}+\|u_{L,\mathrm{II}}\|_{a}^{2}\right)\right)
CHTCω1,pk,I/IIua2,\displaystyle\qquad\leq C_{\mathrm{HT}}\frac{C}{\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}}\|u\|_{a}^{2},

which completes the proof with CI/II=C/ω1,pk,I/IIC_{\mathrm{I}/\mathrm{II}}=C/\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}}. ∎

D.2. The pp-dependence of CI/IIC_{\mathrm{I}/\mathrm{II}}

The previous section shows that we may take CI/II=C/ω1,pk,I/IIC_{\mathrm{I}/\mathrm{II}}=C/\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}} in Lemma 6.3. We show the values of ω1,pk,I/II\omega_{1,p}^{k,\mathrm{I}/\mathrm{II}} for Nedp1\mathrm{Ned}^{1}_{p} and RTp\mathrm{RT}_{p} with p1:10p\in 1:10 in Figure 8. It appears that ω1,p1,I/II𝒪(p3)\omega_{1,p}^{1,\mathrm{I}/\mathrm{II}}\sim\mathcal{O}(p^{-3}), while ω1,p2,I/II𝒪(p2)\omega_{1,p}^{2,\mathrm{I}/\mathrm{II}}\sim\mathcal{O}(p^{-2}). Thus, the type-I/type-II splitting approach may incur an algebraic factor of pp in the condition number estimate for the Riesz map problems. However, we only observed such a drastic growth in terms of iteration counts for the solvers HT(X0,X1,I)\mathrm{HT}({X^{0}},{X^{1,\mathrm{I}}}) and HT(XΓ0,XΓ1,I)+J(X1)\mathrm{HT}({X_{\Gamma}^{0}},{X_{\Gamma}^{1,\mathrm{I}}})+\mathrm{J}({X^{1}_{\mathcal{I}}}) when α=103\alpha=10^{-3} and β=1\beta=1, as shown in section 7.1.2.

11 22 44 66 88 1010 10310^{-3}10210^{-2}10110^{-1}10010^{0}p2p^{-2}p3p^{-3}Degree, ppω1,pk,I/II\omega^{k,\mathrm{I}/\mathrm{II}}_{1,p}Nedp1\mathrm{Ned}^{1}_{p}RTp\mathrm{RT}_{p}
Figure 8. Type based decoupling constant ω1,pk,I/II\omega^{k,\mathrm{I}/\mathrm{II}}_{1,p} on an equilateral reference tetrahedron.

Appendix E Augmented Lagrangian preconditioning for Hodge Laplacians

We first recall a few properties of Hilbert complexes from [5]. Let ({Wk},{dk})(\{W^{k}\},\{\operatorname{d}^{k}\}) be a Hilbert complex with inner-product ,\langle\cdot,\cdot\rangle inducing the norm \|\cdot\|, and let ({Vk},{dk})(\{V^{k}\},\{\operatorname{d}^{k}\}) be the associated domain complex:

(E.1) {\cdots}Vk1{V^{k-1}}Vk{V^{k}}Vk+1{V^{k+1}},{\cdots,}dk2\scriptstyle{\operatorname{d}^{k-2}}dk1\scriptstyle{\operatorname{d}^{k-1}}dk\scriptstyle{\operatorname{d}^{k}}dk+1\scriptstyle{\operatorname{d}^{k+1}}

where the inner-product on VkV^{k} is ,+dk,dk\langle\cdot,\cdot\rangle+\langle\operatorname{d}^{k}\cdot,\operatorname{d}^{k}\cdot\rangle. The complex induces the Hodge decomposition

(E.2) Vk=𝔅kkk,,\displaystyle V^{k}=\mathfrak{B}^{k}\oplus\mathfrak{H}^{k}\oplus\mathfrak{Z}^{k,\perp},

where 𝔅k\mathfrak{B}^{k} is the image of dk1\operatorname{d}^{k-1}, k\mathfrak{H}^{k} is the space of harmonic forms, and k,\mathfrak{Z}^{k,\perp} is the orthogonal complement of the kernel of dk\operatorname{d}^{k}:

𝔅k\displaystyle\mathfrak{B}^{k} :=dk1Vk1=dk1k1,,\displaystyle:=\operatorname{d}^{k-1}V^{k-1}=\operatorname{d}^{k-1}\mathfrak{Z}^{k-1,\perp},
k\displaystyle\mathfrak{H}^{k} :={𝔥Vk:dk𝔥=0 and 𝔥,dk1σ=0σVk1},\displaystyle:=\{\mathfrak{h}\in V^{k}:\operatorname{d}^{k}\mathfrak{h}=0\text{ and }\langle\mathfrak{h},\operatorname{d}^{k-1}\sigma\rangle=0\ \forall\sigma\in V^{k-1}\},
k,\displaystyle\mathfrak{Z}^{k,\perp} :={vVk:v,z=0zVk with dkz=0}.\displaystyle:=\{v\in V^{k}:\langle v,z\rangle=0\ \forall z\in V^{k}\text{ with }\operatorname{d}^{k}z=0\}.

Note that the Hodge decomposition is orthogonal with respect to the inner-product on WkW^{k} and on VkV^{k}. Moreover, given vVkv\in V^{k}, we write v=v𝔅+v+vv=v_{\mathfrak{B}}+v_{\mathfrak{H}}+v_{\perp} to denote the Hodge decomposition of vv.

The mixed formulation of the Hodge Laplace problem associated to E.1 reads as follows: Find (σ,u,p)Vk1×Vk×k(\sigma,u,p)\in V^{k-1}\times V^{k}\times\mathfrak{H}^{k} such that

(E.3a) σ,τ+u,dk1τ\displaystyle-\langle\sigma,\tau\rangle+\langle u,\operatorname{d}^{k-1}\tau\rangle =0\displaystyle=0\qquad τVk1,\displaystyle\forall\tau\in V^{k-1},
(E.3b) dk1σ,v+p,v+dku,dkv\displaystyle\langle\operatorname{d}^{k-1}\sigma,v\rangle+\langle p,v\rangle+\langle\operatorname{d}^{k}u,\operatorname{d}^{k}v\rangle =f,v\displaystyle=\langle f,v\rangle\qquad vVk,\displaystyle\forall v\in V^{k},
(E.3c) u,q\displaystyle\langle u,q\rangle =0\displaystyle=0\qquad qk.\displaystyle\forall q\in\mathfrak{H}^{k}.

We may combine the bilinear forms in E.3 into one large symmetric bilinear form

A(σ,u,p;τ,v,q):=dku,dkv+dk1σ,v+u,dk1τσ,τ+p,v+u,q.\displaystyle A(\sigma,u,p;\tau,v,q):=\langle\operatorname{d}^{k}u,\operatorname{d}^{k}v\rangle+\langle\operatorname{d}^{k-1}\sigma,v\rangle+\langle u,\operatorname{d}^{k-1}\tau\rangle-\langle\sigma,\tau\rangle+\langle p,v\rangle+\langle u,q\rangle.

Given γ>0\gamma>0, we also introduce the following weighted inner-product on Vk1×Vk×kV^{k-1}\times V^{k}\times\mathfrak{H}^{k}:

Bγ(σ,u,p;τ,v,q)\displaystyle B_{\gamma}(\sigma,u,p;\tau,v,q) :=γ1uu,vv+u,v+dku,dkv\displaystyle:=\gamma^{-1}\langle u-u_{\mathfrak{H}},v-v_{\mathfrak{H}}\rangle+\langle u_{\mathfrak{H}},v_{\mathfrak{H}}\rangle+\langle\operatorname{d}^{k}u,\operatorname{d}^{k}v\rangle
+σ,τ+γdk1σ,dk1τ+p,q.\displaystyle\qquad+\langle\sigma,\tau\rangle+\gamma\langle\operatorname{d}^{k-1}\sigma,\operatorname{d}^{k-1}\tau\rangle+\langle p,q\rangle.
Theorem E.1.

Let γ>0\gamma>0 and suppose that Vk1V^{k-1} and VkV^{k} are finite dimensional. Consider the following symmetric eigenvalue problem: Find (σ,u,p)Vk1×Vk×k(\sigma,u,p)\in V^{k-1}\times V^{k}\times\mathfrak{H}^{k} and λ\lambda\in\mathbb{R} such that

(E.4) A(σ,u,p;τ,v,q)=λBγ(σ,u,p;τ,v,q)(τ,v,q)Vk1×Vk×k.\displaystyle A(\sigma,u,p;\tau,v,q)=\lambda B_{\gamma}(\sigma,u,p;\tau,v,q)\qquad\forall(\tau,v,q)\in V^{k-1}\times V^{k}\times\mathfrak{H}^{k}.

All of the eigenpairs of E.4 are classified as follows.

  1. (i)

    For any σVk1\sigma\in V^{k-1}, (σ,γdk1σ,0)(\sigma,-\gamma\operatorname{d}^{k-1}\sigma,0) and λ=1\lambda=-1 is an eigenpair of E.4.

  2. (ii)

    For any pkp\in\mathfrak{H}^{k}, (0,λp,p)(0,\lambda p,p) and λ=±1\lambda=\pm 1 is an eigenpair of E.4.

  3. (iii)

    Let uk,u_{\perp}\in\mathfrak{Z}^{k,\perp} and μ>0\mu>0 be an eigenpair of the following eigenvalue problem:

    (E.5) dku,dkv=μu,vvk,.\displaystyle\langle\operatorname{d}^{k}u_{\perp},\operatorname{d}^{k}v_{\perp}\rangle=\mu\langle u_{\perp},v_{\perp}\rangle\qquad\forall v_{\perp}\in\mathfrak{Z}^{k,\perp}.

    Then, (0,u,0)(0,u_{\perp},0) and λ=μγμγ+1\lambda=\frac{\mu\gamma}{\mu\gamma+1} is an eignpair of E.4.

  4. (iv)

    Let σk1,\sigma_{\perp}\in\mathfrak{Z}^{k-1,\perp} and ν>0\nu>0 be an eigenpair of the following eigenvalue problem:

    (E.6) dk1σ,dk1τ=νσ,ττk1,.\displaystyle\langle\operatorname{d}^{k-1}\sigma_{\perp},\operatorname{d}^{k-1}\tau_{\perp}\rangle=\nu\langle\sigma_{\perp},\tau_{\perp}\rangle\qquad\forall\tau_{\perp}\in\mathfrak{Z}^{k-1,\perp}.

    Then, the following is an eigenpair of E.4:

    (σ,γλ1dk1σ,0)andλ=νγνγ+1.\displaystyle(\sigma_{\perp},\gamma\lambda^{-1}\operatorname{d}^{k-1}\sigma_{\perp},0)\quad\text{and}\quad\lambda=\frac{\nu\gamma}{\nu\gamma+1}.
Proof.

(i-iv) may be verified by direct computation. We only show the details for (iii) and (iv). Let uu_{\perp}, μ\mu, and λ\lambda be as in (iii) and (τ,v,q)Vk1×Vk×k(\tau,v,q)\in V^{k-1}\times V^{k}\times\mathfrak{H}^{k}. Then,

A(0,u,0;τ,v,p)\displaystyle A(0,u_{\perp},0;\tau,v,p) =dku,dkv+u,dk1τ+u,q,=dku,dkv\displaystyle=\langle\operatorname{d}^{k}u_{\perp},\operatorname{d}^{k}v\rangle+\langle u_{\perp},\operatorname{d}^{k-1}\tau\rangle+\langle u_{\perp},q\rangle,=\langle\operatorname{d}^{k}u_{\perp},\operatorname{d}^{k}v_{\perp}\rangle
Bγ(0,u,0;τ,v,p)\displaystyle B_{\gamma}(0,u_{\perp},0;\tau,v,p) =1γu,vv+dku,dkv=1γu,v+dku,dkv.\displaystyle=\frac{1}{\gamma}\langle u_{\perp},v-v_{\mathfrak{H}}\rangle+\langle\operatorname{d}^{k}u_{\perp},\operatorname{d}^{k}v\rangle=\frac{1}{\gamma}\langle u_{\perp},v_{\perp}\rangle+\langle\operatorname{d}^{k}u_{\perp},\operatorname{d}^{k}v_{\perp}\rangle.

Consequently, there holds

λBγ(0,u,0;τ,v,p)=μγμγ+1(1γ+μ)u,v=μu,v=A(0,u,0;τ,v,p).\displaystyle\lambda B_{\gamma}(0,u_{\perp},0;\tau,v,p)=\frac{\mu\gamma}{\mu\gamma+1}\left(\frac{1}{\gamma}+\mu\right)\langle u_{\perp},v_{\perp}\rangle=\mu\langle u_{\perp},v_{\perp}\rangle=A(0,u_{\perp},0;\tau,v,p).

Now let σ\sigma_{\perp}, ν\nu, and λ\lambda be as in (iv). Then,

A(σ,γλ1dk1σ,0;τ,v,q)\displaystyle A(\sigma_{\perp},\gamma\lambda^{-1}\operatorname{d}^{k-1}\sigma_{\perp},0;\tau,v,q) =dk1σ,v+γλ1dk1σ,dk1τσ,τ\displaystyle=\langle\operatorname{d}^{k-1}\sigma_{\perp},v\rangle+\gamma\lambda^{-1}\langle\operatorname{d}^{k-1}\sigma_{\perp},\operatorname{d}^{k-1}\tau\rangle-\langle\sigma_{\perp},\tau\rangle
=dk1σ,v+γλ1dk1σ,dk1τσ,τ,\displaystyle=\langle\operatorname{d}^{k-1}\sigma_{\perp},v_{\perp}\rangle+\gamma\lambda^{-1}\langle\operatorname{d}^{k-1}\sigma_{\perp},\operatorname{d}^{k-1}\tau\rangle-\langle\sigma_{\perp},\tau\rangle,

and

Bγ(σ,γλ1dk1σ,0;τ,v,q)\displaystyle B_{\gamma}(\sigma_{\perp},\gamma\lambda^{-1}\operatorname{d}^{k-1}\sigma_{\perp},0;\tau,v,q)
=λ1dk1σ,vv+σ,τ+γdk1σ,dk1τ\displaystyle\qquad=\lambda^{-1}\langle\operatorname{d}^{k-1}\sigma_{\perp},v-v_{\mathfrak{H}}\rangle+\langle\sigma_{\perp},\tau\rangle+\gamma\langle\operatorname{d}^{k-1}\sigma_{\perp},\operatorname{d}^{k-1}\tau\rangle
=λ1dk1σ,v+σ,τ+γdk1σ,dk1τ\displaystyle\qquad=\lambda^{-1}\langle\operatorname{d}^{k-1}\sigma_{\perp},v_{\perp}\rangle+\langle\sigma_{\perp},\tau\rangle+\gamma\langle\operatorname{d}^{k-1}\sigma_{\perp},\operatorname{d}^{k-1}\tau\rangle
=λ1dk1σ,v+(1+νγ)σ,τ.\displaystyle\qquad=\lambda^{-1}\langle\operatorname{d}^{k-1}\sigma_{\perp},v_{\perp}\rangle+(1+\nu\gamma)\langle\sigma_{\perp},\tau\rangle.

Consequently, there holds

λBγ(σ,γλ1dk1σ,0;τ,v,q)\displaystyle\lambda B_{\gamma}(\sigma_{\perp},\gamma\lambda^{-1}\operatorname{d}^{k-1}\sigma_{\perp},0;\tau,v,q) =dk1σ,v+λ(1+νγ)σ,τ\displaystyle=\langle\operatorname{d}^{k-1}\sigma_{\perp},v_{\perp}\rangle+\lambda\left(1+\nu\gamma\right)\langle\sigma_{\perp},\tau\rangle
=dk1σ,v+νγσ,τ\displaystyle=\langle\operatorname{d}^{k-1}\sigma_{\perp},v_{\perp}\rangle+\nu\gamma\langle\sigma_{\perp},\tau\rangle
=dk1σ,v+(νγλ11)σ,τ\displaystyle=\langle\operatorname{d}^{k-1}\sigma_{\perp},v_{\perp}\rangle+(\nu\gamma\lambda^{-1}-1)\langle\sigma_{\perp},\tau\rangle
=A(σ,γλ1dk1σ,0;τ,v,q).\displaystyle=A(\sigma,\gamma\lambda^{-1}\operatorname{d}^{k-1}\sigma_{\perp},0;\tau,v,q).

We now verify that (i-iv) contain dimVk1+dimVk+dimk\dim V^{k-1}+\dim V^{k}+\dim\mathfrak{H}^{k} linearly independent (L.I.) eigenvectors, and so (i-iv) classify all eigenpairs of E.4. It is immediate that (i) contains dimVk1\dim V^{k-1} L.I. eigenfunctions and (ii) contains dimk\dim\mathfrak{H}^{k} L.I. eigenfunctions for each value of λ{1,1}\lambda\in\{-1,1\}. Since E.5 and E.6 are well-defined, SPD eigenvalue problems, there are dimk,\dim\mathfrak{Z}^{k,\perp} L.I. eigenfunctions in (iii) and dimk1,\dim\mathfrak{Z}^{k-1,\perp} L.I. eigenfunctions in (iv). Thus, there are

dimVk1+dimk+dimk,+dimk1,+dimk=dimVk1+dimVk+dimk\dim V^{k-1}+\dim\mathfrak{H}^{k}+\dim\mathfrak{Z}^{k,\perp}+\dim\mathfrak{Z}^{k-1,\perp}+\dim\mathfrak{H}^{k}\\ =\dim V^{k-1}+\dim V^{k}+\dim\mathfrak{H}^{k}

L.I. eigenfunctions, which completes the proof. ∎

References

  • [1] Mark Ainsworth, Gaelle Andriamaro, and Oleg Davydov, Bernstein–Bézier finite elements of arbitrary order and optimal assembly procedures, SIAM J. Sci. Comput. 33 (2011), no. 6, 3087–3109.
  • [2] Mark Ainsworth and Guosheng Fu, Bernstein–Bézier bases for tetrahedral finite elements, Comput. Methods Appl. Mech. Engrg. 340 (2018), 178–201.
  • [3] Mark Ainsworth, Shuai Jiang, and Manuel A Sanchéz, An 𝒪(p3)\mathcal{O}(p^{3}) hphp-version FEM in two dimensions: Preconditioning and post-processing, Comput. Methods Appl. Mech. Engrg. 350 (2019), 766–802.
  • [4] D. N. Arnold, R. S. Falk, and R. Winther, Multigrid in H(div)H(\mathrm{div}) and H(curl)H(\mathrm{curl}), Numer. Math. 85 (2000), no. 2, 197–217.
  • [5] Douglas Arnold, Richard Falk, and Ragnar Winther, Finite element exterior calculus: From Hodge theory to numerical stability, Bull. Amer. Math. Soc. 47 (2010), no. 2, 281–354.
  • [6] Douglas N Arnold, Richard S Falk, and Ragnar Winther, Finite element exterior calculus, homological techniques, and applications, Acta Numer. 15 (2006), 1–155.
  • [7] S. Beuchler and V. Pillwein, Sparse shape functions for tetrahedral pp-FEM using integrated Jacobi polynomials, Computing 80 (2007), no. 4, 345–375.
  • [8] Sven Beuchler, Veronika Pillwein, Joachim Schöberl, and Sabine Zaglmayr, Sparsity optimized high order finite element functions on simplices, Numerical and Symbolic Scientific Computing (U. Langer and P. Paule, eds.), Texts Monogr. Symbol. Comput., Springer, Vienna, 2012, pp. 21–44. MR 3060506
  • [9] Sven Beuchler, Veronika Pillwein, and Sabine Zaglmayr, Sparsity optimized high order finite element functions for H(div)H({\rm div}) on simplices, Numer. Math. 122 (2012), no. 2, 197–225. MR 2969267
  • [10] by same author, Sparsity optimized high order finite element functions for H(curl)H(\mathrm{curl}) on tetrahedra, Adv. in Appl. Math. 50 (2013), no. 5, 749–769. MR 3044570
  • [11] Jürgen Bey, Simplicial grid refinement: On Freudenthal’s algorithm and the optimal number of congruence classes, Numer. Math. 85 (2000), no. 1, 1–29.
  • [12] D. Boffi, F. Brezzi, and M. Fortin, Mixed finite element methods and applications, Springer Ser. Comput. Math., Springer, Berlin, 2013.
  • [13] Franco Brezzi, Jim Douglas, and L. Donatella Marini, Two families of mixed finite elements for second order elliptic problems, Numerische Mathematik 47 (1985), 217–235.
  • [14] P. D. Brubeck and P. E. Farrell, A scalable and robust vertex-star relaxation for high-order FEM, SIAM J. Sci. Comput. 44 (2022), no. 5, A2991–A3017.
  • [15] Pablo D. Brubeck and Patrick E. Farrell, Multigrid solvers for the de Rham complex with optimal complexity in polynomial degree, SIAM J. Sci. Comput. 46 (2024), no. 3, A1549–A1573.
  • [16] N. Chalmers and T. Warburton, Low-order preconditioning of high-order triangular finite elements, SIAM J. Sci. Comput. 40 (2018), no. 6, A4040–A4059.
  • [17] Philippe G Ciarlet, The finite element method for elliptic problems, Stud. Math. Appl. 4, North Holland, Amsterdam, 1978, Reprinted by SIAM in 2002.
  • [18] L. Demkowicz, P. Monk, L. Vardapetyan, and W. Rachowicz, De Rham diagram for hphp finite element spaces, Comput. Math. Appl. 39 (2000), no. 7, 29–38.
  • [19] O. Egorova, M. Savchenko, V. Savchenko, and I. Hagiwara, Topology and geometry of hexahedral complex: Combined approach for hexahedral meshing, J. Comput. Sci. Technol. 3 (2009), no. 1, 171–182.
  • [20] Alexandre Ern, Johnny Guzmán, Pratyush Potu, and Martin Vohralík, Discrete Poincaré inequalities: A review on proofs, equivalent formulations, and behavior of constants, dec 2024.
  • [21] Richard Falk and Ragnar Winther, Local space-preserving decompositions for the bubble transform, Found. Comput. Math. (2025), 1–49.
  • [22] P. E. Farrell, M. G. Knepley, L. Mitchell, and F. Wechsung, PCPATCH: Software for the topological construction of multigrid relaxation methods, ACM Trans. Math. Softw. 47 (2021), no. 3.
  • [23] M. Fortin and R. Glowinski, Augmented lagrangian methods: Applications to the numerical solution of boundary-value problems, Stud. Math. Appl. 15, North-Holland, Amsertdam, 1983.
  • [24] David A. Ham, Paul H. J. Kelly, Lawrence Mitchell, Colin J. Cotter, Robert C. Kirby, Koki Sagiyama, Nacime Bouziani, Sophia Vorderwuelbecke, Thomas J. Gregory, Jack Betteridge, Daniel R. Shapero, Reuben W. Nixon-Hill, Connor J. Ward, Patrick E. Farrell, Pablo D. Brubeck, India Marsden, Thomas H. Gibson, Miklós Homolya, Tianjiao Sun, Andrew T. T. McRae, Fabio Luporini, Alastair Gregory, Michael Lange, Simon W. Funke, Florian Rathgeber, Gheorghe-Teodor Bercea, and Graham R. Markall, Firedrake user manual, Imperial College London and University of Oxford and Baylor University and University of Washington, first edition ed., 5 2023.
  • [25] R. Hiptmair, Multigrid method for Maxwell’s equations, SIAM J. Numer. Anal. 36 (1998), no. 1, 204–225.
  • [26] R. Hiptmair, Operator preconditioning, Comput. Math. Appl. 52 (2006), no. 5, 699–706.
  • [27] R. Hiptmair, T. Schiekofer, and B. Wohlmuth, Multilevel preconditioned augmented Lagrangian techniques for 2nd order mixed problems, Computing 57 (1996), no. 1, 25–48. MR 1398269
  • [28] Ralf Hiptmair, Multigrid method for H(div)H(\mathrm{div}) in three dimensions, Electron. Trans. Numer. Anal 6 (1997), no. 1, 133–152.
  • [29] Ralf Hiptmair and Andrea Toselli, Overlapping and multilevel Schwarz methods for vector valued elliptic problems in three dimensions, Parallel solution of partial differential equations (Minneapolis, MN, 1997), IMA Vol. Math. Appl., vol. 120, Springer, New York, 2000, pp. 181–208. MR 1838270
  • [30] George Em Karniadakis and Spencer J. Sherwin, Spectral/hphp element methods for computational fluid dynamics, second ed., Numerical Mathematics and Scientific Computation, Oxford University Press, New York, 2005. MR 2165335
  • [31] R. C. Kirby, From functional analysis to iterative methods, SIAM Rev. 52 (2010), no. 2, 269–293.
  • [32] Robert C. Kirby, Fast simplicial finite element algorithms using Bernstein polynomials, Numer. Math. 117 (2011), no. 4, 631–652.
  • [33] by same author, Low-complexity finite element algorithms for the de Rham complex on simplices, SIAM J. Sci. Comput. 36 (2014), no. 2, A846–A868.
  • [34] Cornelius Lanczos, An iteration method for the solution of the eigenvalue problem of linear differential and integral operators, J. Research Nat. Bur. Standards 45 (1950), 255–282. MR 42791
  • [35] Martin Werner Licht, Complexes of discrete distributional differential forms and their homology theory, Found. Comput. Math. 17 (2017), no. 4, 1085–1122.
  • [36] R. E. Lynch, J. R. Rice, and D. H. Thomas, Direct solution of partial difference equations by tensor product methods, Numer. Math. 6 (1964), 185–199.
  • [37] J. Málek and Z. Strakoš, Preconditioning and the conjugate gradient method in the context of solving pdes, SIAM Spotlights, vol. 1, SIAM, 2014.
  • [38] K.-A. Mardal and R. Winther, Preconditioning discretizations of systems of partial differential equations, Numer. Linear Algebra Appl. 18 (2011), no. 1, 1–40.
  • [39] James R. Munkres, Elements of algebraic topology, Addison-Wesley Publishing Company, Menlo Park, CA, 1984. MR 755006
  • [40] Jean-Claude Nédélec, Mixed finite elements in 3\mathbb{R}^{3}, Numer. Math. 35 (1980), no. 3, 315–341.
  • [41] by same author, A new family of mixed finite elements in 3\mathbb{R}^{3}, Numer. Math. 50 (1986), no. 1, 57–81.
  • [42] S. A. Orszag, Spectral methods for problems in complex geometries, J. Comput. Phys. 37 (1980), no. 1, 70–92.
  • [43] L. F. Pavarino, Additive Schwarz methods for the pp-version finite element method, Numer. Math. 66 (1993), no. 1, 493–515.
  • [44] Will Pazner, Tzanio Kolev, and Clark R. Dohrmann, Low-order preconditioning for the high-order finite element de Rham complex, SIAM J. Sci. Comput. 45 (2023), no. 2, A675–A702.
  • [45] Pierre-Arnaud Raviart and Jean-Marie Thomas, A mixed finite element method for 2nd order elliptic problems, Mathematical Aspects of Finite Element Methods (I. Galligani and E. Magenes, eds.), Lecture Notes in Mathematics, vol. 606, Springer, 1977, pp. 292–315.
  • [46] Marie E Rognes, Robert C Kirby, and Anders Logg, Efficient assembly of H(div)H(\mathrm{div}) and H(curl)H(\mathrm{curl}) conforming finite elements, SIAM J. Sci. Comput. 31 (2010), no. 6, 4130–4151.
  • [47] J. Schöberl, J. M. Melenk, C. Pechstein, and S. Zaglmayr, Additive Schwarz preconditioning for pp-version triangular and tetrahedral finite elements, IMA J. Numer. Anal. 28 (2008), no. 1, 1–24.
  • [48] S. J. Sherwin and G. E. Karniadakis, A new triangular and tetrahedral basis for high‐order (hphp) finite element methods, Inter. J. Numer. Methods Engrg. 38 (1995), no. 22, 3775–3802.
  • [49] Spencer J Sherwin and Mario Casarin, Low-energy basis preconditioning for elliptic substructured solvers based on unstructured spectral/hp element discretization, J. Comput. Phys. 171 (2001), no. 1, 394–417.
  • [50] Pavel Šolín and Tomáš Vejchodský, Higher-order finite elements based on generalized eigenfunctions of the Laplacian, Internat. J. Numer. Methods Engrg. 73 (2008), no. 10, 1374–1394. MR 2391331
  • [51] A. Toselli and O. Widlund, Domain decomposition methods–Algorithms and theory, Springer-Verlag, Berlin, 2005.
  • [52] J. Xu, Iterative methods by space decomposition and subspace correction, SIAM Rev. 34 (1992), no. 4, 581–613.
  • [53] S. Zaglmayr, High order finite element methods for electromagnetic field computation, Ph.D. thesis, Johannes Kepler University Linz, 2006.