1 Introduction

Let \(\Omega \subset \mathbb {R}^2\) be a bounded polygonal domain. The energy space for the quad-curl problem to be considered in this paper is

$$\begin{aligned} {\mathbb {E}}=\{\varvec{v}\in [L_2(\Omega )]^2: \mathrm {curl}\,\varvec{v}\in H^1_0(\Omega ), \mathrm {div}\,\varvec{v}=0 \;\text {and}\;\varvec{n}\times \varvec{v}=0\;\text {on}\;\partial \Omega \}, \end{aligned}$$
(1.1)

with the norm \(\Vert \cdot \Vert _{\mathbb {E}}\) given by

$$\begin{aligned} \Vert \varvec{v}\Vert _{\mathbb {E}}^2=\Vert \varvec{v}\Vert _{L_2(\Omega )}^2+|\mathrm {curl}\,\varvec{v}|_{H^1(\Omega )}^2. \end{aligned}$$
(1.2)

Here and below we will follow standard notation for differential operators, function spaces and norms that can be found for example in [1, 13, 16, 21, 25].

We will consider the following problem: Find \(\varvec{u}\in {\mathbb {E}}\) such that

$$\begin{aligned} \big (\mathrm {curl}\,(\mathrm {curl}\,\varvec{u}),\mathrm {curl}\,(\mathrm {curl}\,\varvec{v})\big )+\beta (\mathrm {curl}\,\varvec{u},\mathrm {curl}\,\varvec{v})+ \gamma (\varvec{u},\varvec{v}) =(\varvec{f},\varvec{v}) \qquad \forall \,\varvec{v}\in {\mathbb {E}}, \end{aligned}$$
(1.3)

where \((\cdot ,\cdot )\) denotes the inner product for \({L_2(\Omega )}\) (or \([L_2(\Omega )]^2\)), \(\beta \) and \(\gamma \) are nonnegative constants (\(\gamma >0\) if \(\Omega \) is not simply connected), and \(\varvec{f}\in [L_2(\Omega )]^2\). Since the divergence-free condition is included in the definition of \({\mathbb {E}}\), the problem (1.3) provides an elliptic formulation for the quad-curl problem.

Remark 1.1

In two dimensions the curl of the vector field \(\varvec{v}=[v_1, v_2]^t\) is the scalar function \(\mathrm {curl}\,\varvec{v}=(\partial v_2/\partial x_1)-(\partial v_1/\partial x_2)\), and the curl of a scalar function \(\phi \) is the vector field \(\mathrm {curl}\,\phi =[\partial \phi /\partial x_2 , -\partial \phi /\partial x_1]^t\). An alternative notation for \(\mathrm {curl}\,\phi \) is \(\mathrm {rot}\,\phi \).

The quad-curl problem is related to the Maxwell transmission eigenvalue problem (cf. [14, 26] ) and mathematical models for magnetohydrodynamics with hyperresistivity (cf. [7, 15]). Finite element methods for the quad-curl problem (based on a non-elliptic formulation) were recently developed in [23, 29, 31] using a nonconforming finite element method, a discontinuous Galerkin method and a mixed finite element method. In this paper we will use a Hodge decomposition approach to reduce (1.3) to second order elliptic boundary value problems that can be solved by simple \(H^1\) conforming finite element methods.

We note that the Hodge decomposition approach to time harmonic Maxwell equations on planar domains was investigated in [9] for the perfectly conducting boundary condition and extended to general boundary conditions in [11] with applications to metamaterials. Adaptive and multigrid methods for these problems based on the Hodge decomposition approach were developed in [10, 12, 17]. Applications of the Hodge decomposition to other electromagnetic problems can also be found in [2, 3, 5].

The rest of the paper is organized as follows. We recall the Hodge decomposition for divergence-free vector fields in Sect. 2, where the well-posedness of (1.3) is also addressed. The reduction of (1.3) to second order elliptic boundary value problems is established in Sect. 3. Based on this reduction, we develop \(P_k\) finite element methods for (1.3) in Sect. 4, followed by a convergence analysis in Sect. 5. Numerical results are presented in Sect. 6, and we end with some concluding remarks in Sect. 7.

2 Hodge Decomposition for \(H(\mathrm {div}\,^0;\Omega )\)

The space \(H(\mathrm {div}\,^0;\Omega )\) of divergence-free vector fields is the orthogonal complement of \(\,\mathrm {grad}\,H^1_0(\Omega )\), i.e.,

$$\begin{aligned} H(\mathrm {div}\,^0;\Omega )=\{\varvec{v}\in [L_2(\Omega )]^2:\,(\varvec{v},\,\mathrm {grad}\,\eta )=0 \quad \forall \,\eta \in H^1_0(\Omega )\}, \end{aligned}$$

and \(L_2^0(\Omega )=\{v\in {L_2(\Omega )}:\,(1,v)=0\}\) is the zero-mean subspace of \({L_2(\Omega )}\). Given any \(\varvec{v}\in H(\mathrm {div}\,^0;\Omega )\), we have a unique decomposition (cf. [9, 21]):

$$\begin{aligned} \varvec{v}=\mathrm {curl}\,\psi +\sum _{j=1}^m d_j\,\mathrm {grad}\,\varphi _j, \end{aligned}$$
(2.1)

where \(\psi \in H^1(\Omega )\cap L_2^0(\Omega )\), the non-negative integer m is the Betti number for \(\Omega \) (\(m=0\) if \(\Omega \) is simply connected, cf. Fig. 1), \(d_j\) (\(1\le j\le m\)) are real numbers, and the harmonic functions \(\varphi _1,\ldots ,\varphi _m\) are defined as follows.

Let the outer boundary of \(\Omega \) be denoted by \({\Gamma }_0\) and the m components of the inner boundary be denoted by \({\Gamma }_1,\ldots ,{\Gamma }_m\). Then the harmonic functions \(\varphi _j\) are determined by

$$\begin{aligned} (\,\mathrm {grad}\,\varphi _j,\,\mathrm {grad}\,v)= & {} 0 \quad \forall \,v\in H^1_0(\Omega ), \end{aligned}$$
(2.2a)
$$\begin{aligned} \varphi _j\big |_{{\Gamma }_0}= & {} 0,\end{aligned}$$
(2.2b)
$$\begin{aligned} \varphi _j\big |_{{\Gamma }_k}= & {} {\left\{ \begin{array}{ll} 1&{}\quad j=k\\ 0&{}\quad j\ne k \end{array}\right. } \quad \text {for}\quad 1\le k\le m. \end{aligned}$$
(2.2c)
Fig. 1
figure 1

Betti numbers

Remark 2.1

Note that \(\,\mathrm {grad}\,\varphi _j\) belongs to \({\mathbb {E}}\) for \(1\le j\le m\) (cf. [9, Corollary 2.5]) and

$$\begin{aligned} (\mathrm {curl}\,\psi ,\,\mathrm {grad}\,\varphi _j)=0 \qquad \forall \,\psi \in H^1(\Omega ) \end{aligned}$$
(2.3)

(cf. [9, Lemma 2.4]).

2.1 Properties of the Space \({\mathbb {E}}\)

Let \(\varvec{v}\in {\mathbb {E}}\) be represented by (2.1). We have

$$\begin{aligned} (\mathrm {curl}\,\psi ,\mathrm {curl}\,\rho )=(\varvec{v},\mathrm {curl}\,\rho ) \qquad \forall \,\rho \in H^1(\Omega ) \end{aligned}$$

by (2.3). Since \({\mathbb {E}}\) is a subspace of \(H_0(\mathrm {curl}\,;\Omega )\), we also have (cf. [21, Theorems 2.2.11 and 2.2.12])

$$\begin{aligned} (\varvec{v},\mathrm {curl}\,\rho )=(\mathrm {curl}\,\varvec{v},\rho ) \qquad \forall \,\rho \in H^1(\Omega ). \end{aligned}$$

It follows that the function \(\psi \in H^1(\Omega )\cap L_2^0(\Omega )\) satisfies

$$\begin{aligned} (\mathrm {curl}\,\psi ,\mathrm {curl}\,\rho )=(\mathrm {curl}\,\varvec{v},\rho ) \qquad \forall \,\rho \in H^1(\Omega ). \end{aligned}$$
(2.4)

Remark 2.2

Since \(\mathrm {curl}\,\phi =[{\partial \phi }/{\partial x_2}, -{\partial \phi }/{\partial x_1}]^t\), we have

$$\begin{aligned} (\mathrm {curl}\,\phi ,\mathrm {curl}\,\rho )=(\,\mathrm {grad}\,\phi ,\,\mathrm {grad}\,\rho ) \qquad \forall \,\phi ,\rho \in H^1(\Omega ) \end{aligned}$$

and \(\Vert \mathrm {curl}\,\rho \Vert _{L_2(\Omega )}=|\rho |_{H^1(\Omega )}\) for all \(\rho \in H^1(\Omega )\).

We can deduce properties of \(\varvec{v}\) from the following results (cf. [22, Sect. 5.1] and [18, Sect. 2.5]) for elliptic boundary value problems on polygonal domains, where \(\omega \) is the largest interior angle at the corners of \(\Omega \).

Lemma 2.3

Let \(\mu \in H^1_0(\Omega )\) satisfy

$$\begin{aligned} (\,\mathrm {grad}\,\mu ,\,\mathrm {grad}\,\eta )+\beta (\mu ,\eta )=(g,\eta ) \qquad \forall \,\eta \in H^1_0(\Omega ), \end{aligned}$$

where \(g\in H^1(\Omega )\). Then we have \(\mu \in H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for any \(\epsilon >0\) and

$$\begin{aligned} \Vert \mu \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}\le {C_\epsilon }\Vert g\Vert _{H^1(\Omega )}. \end{aligned}$$

Lemma 2.4

Let \(\lambda \in H^1(\Omega )\) satisfy

$$\begin{aligned} (\,\mathrm {grad}\,\lambda ,\,\mathrm {grad}\,\psi )+(\lambda ,1)(\psi ,1)=(g,\psi ) \qquad \forall \,\psi \in H^1(\Omega ), \end{aligned}$$

where \(g\in H^1(\Omega )\). Then we have \(\lambda \in H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for any \(\epsilon >0\) and

$$\begin{aligned} \Vert \lambda \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}\le {C_\epsilon }\Vert g\Vert _{H^1(\Omega )}. \end{aligned}$$

Since \(\psi \) belongs to \(H^1(\Omega )\cap L_2^0(\Omega )\) and \(\mathrm {curl}\,\varvec{v}\) belongs to \(H^1_0(\Omega )\), we can apply Lemma 2.4 to (2.4) and conclude that \(\psi \) belongs to \(H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for any \(\epsilon >0\). Note that (2.2a), (2.2b), (2.2c) can be transformed to a problem of the form in Lemma 2.3 where \(\beta =0\) and \(g\in C^\infty (\bar{\Omega })\). Hence we can apply Lemma 2.3 to conclude that \(\varphi _j\) belongs to \(H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for \(1\le j\le m\) and any \(\epsilon >0\). Then (2.1) implies that \(\varvec{v}\) belongs to \([H^{{(\pi /\omega )-\epsilon }}(\Omega )]^2\) and we have established the following result.

Theorem 2.5

The space \({\mathbb {E}}\) is a subspace of \([H^{{(\pi /\omega )-\epsilon }}(\Omega )]^2\) for any \(\epsilon >0\), where \(\omega \) is the largest angle at the corners of \(\Omega \).

Remark 2.6

If \(\Omega \) is a smooth domain, then we can apply the elliptic regularity theory for such domains [27] to conclude that \(\psi \) belongs to \(H^3(\Omega )\) and \(\varphi _j\) belongs to \(C^\infty (\bar{\Omega })\) for \(1\le j\le m\). It follows that \({\mathbb {E}}\) is a subspace of \([H^2(\Omega )]^2\).

2.2 Well-Posedness of (1.3)

Since \({\mathbb {E}}\) is compactly embedded in \([L_2(\Omega )]^2\) by Theorem 2.5 and the Rellich–Kondrachov Theorem [1], we can establish the well-posedness of (1.3) by the Fredholm theory [30]. It suffices to show that if \(\varvec{v}\in {\mathbb {E}}\) satisfies

$$\begin{aligned} \big (\mathrm {curl}\,(\mathrm {curl}\,\varvec{v}),\mathrm {curl}\,(\mathrm {curl}\,\varvec{w})\big )+\beta (\mathrm {curl}\,\varvec{v},\mathrm {curl}\,\varvec{w})+ \gamma (\varvec{v},\varvec{w}) =0\qquad \forall \,\varvec{w}\in {\mathbb {E}}, \end{aligned}$$
(2.5)

then \(\varvec{v}=0\).

This is obvious if \(\gamma >0\). In the case where \(\gamma =0\) and \(\Omega \) is simply connected, we deduce from (2.5) that

$$\begin{aligned} \mathrm {curl}\,(\mathrm {curl}\,\varvec{v})=0 \end{aligned}$$

and hence \(\mathrm {curl}\,\varvec{v}=0\) (because \(\mathrm {curl}\,\varvec{v}\in H^1_0(\Omega )\)). It then follows from (2.4) that \(\psi =0\) and hence \(\varvec{v}=0\) by (2.1).

3 Reduction to Second Order Elliptic Boundary Value Problems

According to (2.1), we can write

$$\begin{aligned} \varvec{u}=\mathrm {curl}\,\phi +\sum _{j=1}^m c_j\,\mathrm {grad}\,\varphi _j, \end{aligned}$$
(3.1)

where \(\phi \in H^1(\Omega )\cap L_2^0(\Omega )\) and \(c_1,\ldots ,c_m\) are real numbers. The idea of the Hodge decomposition approach is to find \(\phi \) and \(c_1,\ldots ,c_m\), and then recover \(\varvec{u}\) by (3.1).

Below we will find problems that determine the function \(\phi \) and the coefficients \(c_1,\ldots ,c_m\) in the decomposition (3.1).

3.1 A Problem for \(\phi \)

It follows from (2.4) and (3.1) that

$$\begin{aligned} (\mathrm {curl}\,\phi ,\mathrm {curl}\,\psi )=(\xi ,\psi ) \qquad \forall \,\psi \in H^1(\Omega ), \end{aligned}$$
(3.2)

where \(\xi =\mathrm {curl}\,\varvec{u}\in H^1_0(\Omega )\). Note that \(\varvec{n}\times \varvec{u}=0\) on \(\partial \Omega \) implies \((1,\xi )=0\) and hence the singular Neumann boundary value problem (3.2) has a unique solution \(\phi \in H^1(\Omega )\cap L_2^0(\Omega )\). An equivalent formulation that avoids the zero-mean constraint is to find \(\phi \in H^1(\Omega )\) such that

$$\begin{aligned} (\mathrm {curl}\,\phi ,\mathrm {curl}\,\psi )+(\phi ,1)(\psi ,1)=(\xi ,\psi ) \qquad \forall \,\psi \in H^1(\Omega ). \end{aligned}$$
(3.3)

It only remains to find a problem that determines \(\xi \).

3.2 A Problem for \(\xi \)

We begin with a lemma.

Lemma 3.1

The function \(\xi =\mathrm {curl}\,\varvec{u}\in H^1_0(\Omega )\cap L_2^0(\Omega )\) satisfies

$$\begin{aligned} \big (\mathrm {curl}\,\xi ,\mathrm {curl}\,(\mathrm {curl}\,\varvec{\zeta })\big )+\beta (\xi ,\mathrm {curl}\,\varvec{\zeta })+\gamma (\varvec{u},\varvec{\zeta })=(Q\varvec{f},\varvec{\zeta }) \qquad \forall \,\varvec{\zeta }\in [C^\infty _c(\Omega )]^2, \end{aligned}$$
(3.4)

where Q is the orthogonal projection from \([L_2(\Omega )]^2\) onto \(H(\mathrm {div}\,^0;\Omega )\).

Proof

Since \(\varvec{\zeta }-Q\varvec{\zeta }\) belongs to \(\,\mathrm {grad}\,(H^1_0(\Omega ))\), we have \(\mathrm {curl}\,Q\varvec{\zeta }=\mathrm {curl}\,\varvec{\zeta }\in H^1_0(\Omega )\), \(\varvec{n}\times Q\varvec{\zeta }=\varvec{n}\times \varvec{\zeta }=0\) on \(\partial \Omega \) (cf. [9, Corollary 2.5]) and hence \(Q\varvec{\zeta }\in {\mathbb {E}}\). Therefore (3.4) follows from (1.3):

$$\begin{aligned}&\,\big (\mathrm {curl}\,\xi ,\mathrm {curl}\,(\mathrm {curl}\,\varvec{\zeta })\big )+\beta (\xi ,\mathrm {curl}\,\varvec{\zeta })+\gamma (\varvec{u},\varvec{\zeta })\\&\quad = \big (\mathrm {curl}\,(\mathrm {curl}\,\varvec{u}),\mathrm {curl}\,(\mathrm {curl}\,\varvec{\zeta })\big )+\beta (\mathrm {curl}\,\varvec{u},\mathrm {curl}\,\varvec{\zeta })+ \gamma (\varvec{u},\varvec{\zeta })\\&\quad = \big (\mathrm {curl}\,(\mathrm {curl}\,\varvec{u}),\mathrm {curl}\,(\mathrm {curl}\,Q\varvec{\zeta })\big )+\beta (\mathrm {curl}\,\varvec{u},\mathrm {curl}\,Q\varvec{\zeta })+ \gamma (\varvec{u},Q\varvec{\zeta })\\&\quad = (Q\varvec{f},\varvec{\zeta }) \end{aligned}$$

\(\square \)

It follows from (3.4) that

$$\begin{aligned} \mathrm {curl}\,(-\Delta \xi )=-\beta \,\mathrm {curl}\,\xi -\gamma \varvec{u}+Q\varvec{f}\end{aligned}$$
(3.5)

in the sense of distributions.

We will exploit (3.5) through the following lemma due to Nečas [8, 20, 28].

Lemma 3.2

If \(\tau \), \(\partial \tau /\partial x_1\) and \(\partial \tau /\partial x_2\) belong to \(H^{-1}(\Omega )\) , then \(\tau \) belongs to \({L_2(\Omega )}\).

Since \(-\Delta \xi \) belongs to \(H^{-1}(\Omega )=[H^1_0(\Omega )]'\) and the right-hand side of (3.5) belongs to \([L_2(\Omega )]^2\), we can apply Lemma 3.2 to conclude that \(-\Delta \xi \) belongs to \({L_2(\Omega )}\). Then (3.5) implies

$$\begin{aligned} -\Delta \xi \in H^1(\Omega ). \end{aligned}$$
(3.6)

Let \(\rho \in H^1(\Omega )\cap L_2^0(\Omega )\) be defined by the following consistent singular Neumann boundary value problem

$$\begin{aligned} (\mathrm {curl}\,\rho ,\mathrm {curl}\,\psi )= & {} -\gamma (\varvec{u},\mathrm {curl}\,\psi )+(Q\varvec{f},\mathrm {curl}\,\psi )\\= & {} -\gamma (\xi ,\psi )+(\varvec{f},\mathrm {curl}\,\psi ) \qquad \qquad \forall \,\psi \in H^1(\Omega ).\nonumber \end{aligned}$$
(3.7)

Since the relations (3.5)–(3.7) imply \(-\Delta \xi +\beta \xi \in H^1(\Omega )\) and

$$\begin{aligned} (\mathrm {curl}\,(-\Delta \xi +\beta \xi ),\mathrm {curl}\,\psi )=(\mathrm {curl}\,\rho ,\mathrm {curl}\,\psi ) \qquad \forall \,\psi \in H^1(\Omega ), \end{aligned}$$

we have

$$\begin{aligned} -\Delta \xi +\beta \xi =\rho +c \end{aligned}$$

for some constant c, and hence

$$\begin{aligned} (\mathrm {curl}\,\xi ,\mathrm {curl}\,\eta )+\beta (\xi ,\eta )=(\rho ,\eta ) \qquad \forall \,\eta \in H^1_0(\Omega )\cap L_2^0(\Omega ). \end{aligned}$$
(3.8)

Below we will show that the function \(\xi \in H^1(\Omega )\cap L_2^0(\Omega )\) is determined by (3.7) and (3.8).

3.2.1 The Case Where \(\gamma =0\)

When \(\gamma \) is 0 (and \(\Omega \) is simply connected), the two equations (3.7) and (3.8) are decoupled. We can first solve (3.7) for \(\rho \) and then solve (3.8) for \(\xi \).

In this case (3.7) becomes a consistent singular Neumann boundary value problem: Find \(\rho \in H^1(\Omega )\cap L_2^0(\Omega )\) such that

$$\begin{aligned} (\mathrm {curl}\,\rho ,\mathrm {curl}\,\psi )=(\varvec{f},\mathrm {curl}\,\psi ) \qquad \forall \,\psi \in H^1(\Omega ). \end{aligned}$$
(3.9)

An equivalent formulation without the zero-mean constraint is to find \(\rho \in H^1(\Omega )\) such that

$$\begin{aligned} (\mathrm {curl}\,\rho ,\mathrm {curl}\,\psi )+(\rho ,1)(\psi ,1)=(\varvec{f},\mathrm {curl}\,\psi ) \qquad \forall \,\psi \in H^1(\Omega ). \end{aligned}$$
(3.10)

Once we have found \(\rho \in H^1(\Omega )\cap L_2^0(\Omega )\), \(\xi \in H^1_0(\Omega )\cap L_2^0(\Omega )\) is determined by the well-posed (nonstandard) elliptic boundary value problem (3.8). We can also determine \(\xi \) through standard boundary value problems that do not involve the zero-mean constraint.

Lemma 3.3

The solution \(\xi \) of (3.8) is given by

$$\begin{aligned} \xi =\xi _0-\frac{(1,\xi _0)}{(1,\xi _1)}\xi _1, \end{aligned}$$
(3.11)

where \(\xi _0,\xi _1\in H^1_0(\Omega )\) satisfy

$$\begin{aligned} (\mathrm {curl}\,\xi _0,\mathrm {curl}\,\eta )+\beta (\xi _0,\eta )= & {} (\rho ,\eta ) \qquad \forall \,\eta \in H^1_0(\Omega ),\end{aligned}$$
(3.12)
$$\begin{aligned} (\mathrm {curl}\,\xi _1,\mathrm {curl}\,\eta )+\beta (\xi _1,\eta )= & {} (1,\eta ) \qquad \forall \,\eta \in H^1_0(\Omega ). \end{aligned}$$
(3.13)

Proof

First we note that (3.12) and (3.13) are standard elliptic boundary value problems and that (3.13) implies \((1,\xi _1)>0\).

By construction, the function \(\xi \) belongs to \(H^1_0(\Omega )\cap L_2^0(\Omega )\) and

$$\begin{aligned} (\mathrm {curl}\,\xi ,\mathrm {curl}\,\eta )+\beta (\xi ,\eta )=(\rho ,\eta ) -\frac{(1,\xi _0)}{(1,\xi _1)}(1,\eta ) \qquad \forall \,\eta \in H^1_0(\Omega ), \end{aligned}$$

which implies (3.8). \(\square \)

3.2.2 The Case Where \(\gamma >0\)

When \(\gamma \) is positive, the problems (3.7) and (3.8) are coupled and we can reformulate them as the following problem:

Find \((\zeta ,\xi )\in [H^1(\Omega )\cap L_2^0(\Omega )]\times [H^1_0(\Omega )\cap L_2^0(\Omega )]\) such that

$$\begin{aligned} (\mathrm {curl}\,\zeta ,\mathrm {curl}\,\psi )+\gamma ^\frac{1}{2}(\psi ,\xi )= & {} \gamma ^{-\frac{1}{2}}(\varvec{f},\mathrm {curl}\,\psi ) \quad \forall \,\psi \in H^1(\Omega )\cap L_2^0(\Omega ), \nonumber \\\end{aligned}$$
(3.14a)
$$\begin{aligned} -\gamma ^\frac{1}{2}(\zeta ,\eta )+ (\mathrm {curl}\,\xi ,\mathrm {curl}\,\eta )+\beta (\xi ,\eta )= & {} 0\quad \forall \,\eta \in H^1_0(\Omega )\cap L_2^0(\Omega ), \end{aligned}$$
(3.14b)

where \(\zeta =\gamma ^{-\frac{1}{2}}\rho \).

We can also write (3.14a), (3.14b) concisely as

$$\begin{aligned} \mathcal {A} \big ((\zeta ,\xi ),(\psi ,\eta )\big )=\gamma ^{-\frac{1}{2}}(\varvec{f},\mathrm {curl}\,\psi )\qquad \forall \,\psi \in H^1(\Omega )\cap L_2^0(\Omega ), \eta \in H^1_0(\Omega )\cap L_2^0(\Omega ),\nonumber \\ \end{aligned}$$
(3.15)

where the bilinear form \(\mathcal {A}(\cdot ,\cdot )\) on \(H^1(\Omega )\times H^1_0(\Omega )\) is defined by

$$\begin{aligned} \mathcal {A}\big ((\zeta ,\xi ),(\psi ,\eta )\big )= (\mathrm {curl}\,\zeta ,\mathrm {curl}\,\psi )+ \gamma ^\frac{1}{2}(\psi ,\xi ) -\gamma ^\frac{1}{2}(\zeta ,\eta )+ (\mathrm {curl}\,\xi ,\mathrm {curl}\,\eta )+\beta (\xi ,\eta ).\nonumber \\ \end{aligned}$$
(3.16)

The bilinear form \(\mathcal {A}(\cdot ,\cdot )\) is clearly bounded on \(H^1(\Omega )\times H^1_0(\Omega )\), and it follows from the identity

$$\begin{aligned} \mathcal {A}\big ((\psi ,\eta ),(\psi ,\eta )\big )=(\mathrm {curl}\,\psi ,\mathrm {curl}\,\psi )+(\mathrm {curl}\,\eta ,\mathrm {curl}\,\eta )+\beta (\eta ,\eta ) \end{aligned}$$
(3.17)

and standard Poincaré-Friedrichs inequalities [27] that \(\mathcal {A}(\cdot ,\cdot )\) is coercive on \([H^1(\Omega )\cap L_2^0(\Omega )]\times H^1_0(\Omega )\). Therefore the problem (3.14a), (3.14b) is well-posed by the Lax–Milgram theorem [24].

We can also determine \((\zeta ,\xi )\) through problems that do not involve the zero-mean constraint.

Lemma 3.4

The solution \((\zeta ,\xi )\) of (3.14a), (3.14b) is given by

$$\begin{aligned} (\zeta ,\xi )=(\zeta _0,\xi _0)-\frac{(1,\xi _0)}{(1,\xi _1)}(\zeta _1,\xi _1), \end{aligned}$$
(3.18)

where \((\zeta _0,\xi _0), (\zeta _1,\xi _1)\in H^1(\Omega )\times H^1_0(\Omega )\) are defined by

$$\begin{aligned} \mathcal {A}\big ((\zeta _0,\xi _0),(\psi ,\eta )\big )+(\zeta _0,1)(\psi ,1)= & {} \gamma ^{-\frac{1}{2}}(\varvec{f},\mathrm {curl}\,\psi )\qquad \forall \, (\psi ,\eta )\in H^1(\Omega )\times H^1_0(\Omega ),\nonumber \\\end{aligned}$$
(3.19)
$$\begin{aligned} \mathcal {A}\big ((\zeta _1,\xi _1),(\psi ,\eta )\big )+(\zeta _1,1)(\psi ,1)= & {} (1,\eta )\qquad \forall \, (\psi ,\eta )\in H^1(\Omega )\times H^1_0(\Omega ). \end{aligned}$$
(3.20)

Proof

First we note that (3.17) implies

$$\begin{aligned}&\mathcal {A}\big (\psi ,\eta ),(\psi ,\eta )\big )+(\psi ,1)(\psi ,1)\nonumber \\&\quad =(\mathrm {curl}\,\psi ,\mathrm {curl}\,\psi )+(\psi ,1)(\psi ,1)+(\mathrm {curl}\,\eta ,\mathrm {curl}\,\eta )+\beta (\eta ,\eta ) \end{aligned}$$
(3.21)

and hence, by standard Poincaré-Friedrichs inequalities, the problems (3.19) and (3.20) are well-posed. Moreover (3.20) and (3.21) imply that \((1,\xi _1)>0\).

By construction, we have \((1,\xi )=0\) and

$$\begin{aligned} \mathcal {A}\big ((\zeta ,\xi ),(\psi ,\eta )\big )+(\zeta ,1)(\psi ,1)=\gamma ^{-\frac{1}{2}}(\varvec{f},\mathrm {curl}\,\psi ) -\frac{(1,\xi _0)}{(1,\xi _1)}(1,\eta ) \end{aligned}$$
(3.22)

for all \((\psi ,\eta )\in H^1(\Omega )\times H^1_0(\Omega )\), which implies (3.15). We can also recover the zero-mean condition \((1,\zeta )=0\) by taking \((\psi ,\eta )=(1,0)\) in (3.22). \(\square \)

3.3 A Problem for \(\varvec{c}_1,\ldots , \varvec{c}_m\)

If \(\Omega \) is not simply connected, then \(\gamma \) is positive. According to Remark 2.1, we can take \(\varvec{v}=\,\mathrm {grad}\,\varphi _i\) in (1.3) to obtain

$$\begin{aligned} \gamma (\varvec{u},\,\mathrm {grad}\,\varphi _i)=(\varvec{f},\,\mathrm {grad}\,\varphi _i) \qquad \text {for}\quad 1\le i\le m. \end{aligned}$$

It then follows from (2.3) and (3.1) that the coefficients \(c_1,\ldots ,c_m\) in the Hodge decomposition (3.1) are determined by the \(m\times m\) system

$$\begin{aligned} \sum _{j=1}^m (\,\mathrm {grad}\,\varphi _i,\,\mathrm {grad}\,\varphi _j)c_j=\frac{1}{\gamma }(\varvec{f},\,\mathrm {grad}\,\varphi _i) \qquad \text {for}\quad 1\le i\le m. \end{aligned}$$
(3.23)

Note that (3.23) is symmetric positive definite because of (2.2b).

3.4 Regularity of \(\varvec{u}\)

First we observe that \(\rho \) belongs to \(H^1(\Omega )\) by construction. Then (3.8) and Lemma 2.3 imply that \(\xi \in H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for any \(\epsilon >0\), and (3.2) and Lemma 2.4 imply that \(\phi \) belongs to \(H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for any \(\epsilon >0\). Moreover the harmonic functions \(\varphi _j\) (\(1\le j\le m)\) also belong to \(H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) by Lemma 2.3. It follows that \(\varvec{u}\) belongs to \([H^{{(\pi /\omega )-\epsilon }}(\Omega )]^2\) for any \(\epsilon >0\). Thus the regularity of \(\varvec{u}\) is better than \(H^1\) for a convex polygon and worse than \(H^1\) for a nonconvex polygon.

Remark 3.5

Note that the regularity of \(\xi =\mathrm {curl}\,\varvec{u}\) is better than \(H^1(\Omega )\). But the regularity of \(\varvec{u}\) is the same as the one in Theorem 2.5 for \({\mathbb {E}}\). This is due to the presence of singularities at the corners of \(\Omega \) that prevents full elliptic regularity for \(\varvec{u}\). In contrast, for a smooth \(\Omega \), \(\rho \in H^1(\Omega )\) implies \(\xi \in H^3(\Omega )\) by (3.8), which in turn implies \(\phi \in H^5(\Omega )\) by (3.2). Since the harmonic functions \(\varphi _j\) for \(1\le j\le m\) belong to \(C^\infty (\bar{\Omega })\), \(\varvec{u}\) belongs to \([H^4(\Omega )]^2\) by (3.1), which is two orders higher than the regularity of the vector fields in \({\mathbb {E}}\) (cf. Remark 2.6).

4 \({P_k}\) Finite Element Methods

The reduction in Sect. 3 leads to the following numerical procedure for (1.3).

  1. (1)

    Find an approximation \(\tilde{\xi }\in H^1_0(\Omega )\cap L_2^0(\Omega )\) for \(\xi \) numerically. In the case where \(\gamma =0\) (and \(\Omega \) is simply connected), one can first solve (3.9) numerically to find an approximation \(\tilde{\rho }\in H^1_0(\Omega )\cap L_2^0(\Omega )\) for \(\rho \), and then solve (3.8) (with \(\rho \) replaced by \(\tilde{\rho }\)) numerically to find an approximation \(\tilde{\xi }\) for \(\xi \) (cf. Sect. 3.2.1). In the case where \(\gamma >0\), we can obtain \(\tilde{\xi }\) by solving the coupled problem (3.14a), (3.14b) numerically (cf. Sect. 3.2.2).

  2. (2)

    Solve (3.2) (with \(\xi \) replaced by \(\tilde{\xi }\)) numerically to find an approximation \(\tilde{\phi }\in H^1(\Omega )\cap L_2^0(\Omega )\) for \(\phi \).

  3. (3)

    In the case where \(\Omega \) is not simply connected (and \(\gamma >0\)), solve numerically the boundary value problems in (2.2a), (2.2b), (2.2c) to find approximations \(\tilde{\varphi }_j \in H^1(\Omega )\) for \(\varphi _j\) (\(1\le j\le m\)) and then solve (3.23) (with \(\varphi _j\) replaced by \(\tilde{\varphi }_j\)) numerically to find approximations \(\tilde{c}_1,\ldots ,\tilde{c}_m\) for \(c_1,\ldots ,c_m\). Note that the computation of \(\tilde{\varphi }_j\) (\(1\le j\le m\)) only involves \(\Omega \) and hence can be carried out in advance.

  4. (4)

    The approximation \(\tilde{\varvec{u}}\) of \(\varvec{u}\) is given by

    $$\begin{aligned} \tilde{\varvec{u}}=\mathrm {curl}\,\tilde{\phi }+\sum _{j=1}^m\tilde{c}_j\,\mathrm {grad}\,\tilde{\varphi }_j. \end{aligned}$$

Therefore any numerical method that works for second order elliptic boundary value problems can also be applied to (1.3). Here we will consider \(P_k\) Lagrange finite element methods.

Let \(\mathcal {T}_h\) be a quasi-uniform simplicial triangulation of \(\Omega \). We denote by \(V_h(\subset H^1(\Omega ))\) the \(P_k\) \((k\ge 1)\) Lagrange finite element space [13, 16] associated with \(\mathcal {T}_h\) and by \(\mathring{V}_h(\subset H^1_0(\Omega ))\) the subspace of \(V_h\) whose members vanish on \(\partial \Omega \).

4.1 The \(\varvec{P}_k\) Finite Element Method for the Approximation of \(\varvec{\xi }\)

We consider two separate cases depending on whether \(\gamma \) is 0 or positive.

4.1.1 The Case Where \(\gamma =0\)

Following the discussion in Sect. 3.2.1, we can first compute \(\rho \) and then \(\xi \).

The \(P_k\) finite element method for (3.10) is to find \(\rho _h\in V_h\) such that

$$\begin{aligned} (\mathrm {curl}\,\rho _h,\mathrm {curl}\,\psi )+(\rho ,1)(\psi ,1)=(\varvec{f},\mathrm {curl}\,\psi ) \qquad \forall \,\psi \in V_h. \end{aligned}$$
(4.1)

The \(P_k\) finite element method for (3.8) [cf. (3.11)–(3.13)] is to find

$$\begin{aligned} \xi _h=\xi _{0,h}-\frac{(1,\xi _{0,h})}{(1,\xi _{1,h})}\xi _{1,h}, \end{aligned}$$
(4.2)

where \(\xi _{0,h},\xi _{1,h}\in \mathring{V}_h\) satisfy

$$\begin{aligned} (\mathrm {curl}\,\xi _{0,h},\mathrm {curl}\,\eta )+\beta (\xi _{0,h},\eta )= & {} (\rho _h,\eta ) \qquad \forall \,\eta \in \mathring{V}_h,\end{aligned}$$
(4.3)
$$\begin{aligned} (\mathrm {curl}\,\xi _{1,h},\mathrm {curl}\,\eta )+\beta (\xi _{1,h},\eta )= & {} (1,\eta ) \qquad \forall \,\eta \in \mathring{V}_h. \end{aligned}$$
(4.4)

4.1.2 The Case Where \(\gamma >0\)

The \(P_k\) finite element method for (3.14a), (3.14b) [cf. (3.18)–(3.20)] is to find

$$\begin{aligned} (\zeta _h,\xi _h)=(\zeta _{0,h},\xi _{0,h})-\frac{(1,\xi _{0,h})}{(1,\xi _{1,h})}(\zeta _{1,h},\xi _{1,h}), \end{aligned}$$
(4.5)

where \((\zeta _{0,h},\xi _{0,h}), (\zeta _{1,h},\xi _{1,h})\in V_h\times \mathring{V}_h\) satisfy

$$\begin{aligned} \mathcal {A}\big ((\zeta _{0,h},\xi _{0,h}),(\psi ,\eta )\big )+(\zeta _{0,h},1)(\psi ,1)= & {} \gamma ^{-\frac{1}{2}}(\varvec{f},\mathrm {curl}\,\psi )\qquad \forall \, (\psi ,\eta )\in V_h\times \mathring{V}_h,\nonumber \\\end{aligned}$$
(4.6)
$$\begin{aligned} \mathcal {A}\big ((\zeta _{1,h},\xi _{1,h}),(\psi ,\eta )\big )+(\zeta _{1,h},1)(\psi ,1)= & {} (1,\eta )\qquad \forall \, (\psi ,\eta )\in V_h\times \mathring{V}_h. \end{aligned}$$
(4.7)

Remark 4.1

The function \(\xi _h\) is an approximation of \(\mathrm {curl}\,\varvec{u}\).

4.2 The \(\varvec{P}_k\) Finite Element Method for the Approximation of \(\varvec{\phi }\)

The finite element method for (3.3) is to find \(\phi _h\in V_h\cap L_2^0(\Omega )\) such that

$$\begin{aligned} (\mathrm {curl}\,\phi ,\mathrm {curl}\,\psi )+(\phi ,1)(\psi ,1)=(\xi _h,\psi ) \qquad \forall \,\psi \in V_h. \end{aligned}$$
(4.8)

4.3 The Approximation of \(\varvec{u}\)

We take

$$\begin{aligned} \varvec{u}_h=\mathrm {curl}\,\phi _h+\sum _{j=1}^m c_{j,h}\varphi _{j,h} \end{aligned}$$
(4.9)

to be the approximation of \(\varvec{u}\), where \(c_{1,h},\ldots ,c_{m,h}\) are determined by

$$\begin{aligned} \sum _{j=1}^m (\,\mathrm {grad}\,\varphi _{i,h},\,\mathrm {grad}\,\varphi _{j,h})c_{j,h}=\gamma ^{-1}(\varvec{f},\,\mathrm {grad}\,\varphi _{i,h}) \qquad \text {for}\quad 1\le i\le m, \end{aligned}$$
(4.10)

and the discrete harmonic functions \(\varphi _{1,h},\ldots ,\varphi _{m,h}\) are determined by (cf. (2.2a), (2.2b), (2.2c))

$$\begin{aligned} (\,\mathrm {grad}\,\varphi _{j,h},\,\mathrm {grad}\,v)= & {} 0\qquad \forall \,v\in \mathring{V}_h, \end{aligned}$$
(4.11a)
$$\begin{aligned} \varphi _{j,h}\big |_{{\Gamma }_0}= & {} 0,\end{aligned}$$
(4.11b)
$$\begin{aligned} \varphi _{j,h}\big |_{{\Gamma }_k}= & {} {\left\{ \begin{array}{ll} 1\quad j=k\\ 0\quad j\ne k \end{array}\right. }\qquad \text {for}\quad 1\le k\le m. \end{aligned}$$
(4.11c)

For a simply connected \(\Omega \), the approximation for \(\varvec{u}\) is simplified to \(\varvec{u}_h=\mathrm {curl}\,\phi _h\).

5 Convergence Analysis

Since the error analysis for the discrete harmonic functions \(\varphi _{1,h},\ldots ,\varphi _{m,h}\) and the coefficients \(c_{1,h},\ldots ,c_{m,h}\) has already been carried out in [9], we only need to focus on the error analysis for \(\xi _h\) and \(\phi _h\).

We will use the following standard polynomial approximation result [13, 16, 19].

Lemma 5.1

Given any \(\delta >0\), there exists a positive constant C independent of h such that

$$\begin{aligned}&\inf _{\psi \in V_h}\Vert \lambda -\psi \Vert _{H^1(\Omega )}\le Ch^{\min (\delta ,k)}\Vert \lambda \Vert _{H^{1+\delta }(\Omega )} \qquad \forall \,\lambda \in H^{1+\delta }(\Omega ),\\&\inf _{\eta \in \mathring{V}_h}\Vert \mu -\psi \Vert _{H^1(\Omega )}\le Ch^{\min (\delta ,k)}\Vert \mu \Vert _{H^{1+\delta }(\Omega )} \qquad \forall \,\mu \in H^{1+\delta }(\Omega )\cap H^1_0(\Omega ). \end{aligned}$$

From here on we will use C (with or without subscript) to denote a generic positive constant that is independent of h. The error estimates below will depend on \(\omega \), the largest interior angle at the corners of \(\Omega \).

5.1 Error Analysis for \(\varvec{\xi }_{\varvec{h}}\)

Our goal is to establish the following result.

Lemma 5.2

For any \(\epsilon >0\), there exists a positive constant \({C_\epsilon }\) independent of h such that

$$\begin{aligned} |\xi -\xi _h|_{H^1(\Omega )}\le C_\epsilon h^{{\min ({(\pi /\omega )-\epsilon },k)}}. \end{aligned}$$
(5.1)

We will consider the two cases \(\gamma =0\) and \(\gamma >0\) separately.

5.1.1 The Case Where \({\gamma }=0\)

We first estimate the error for \(\rho _h\) in the norm of \([H^1(\Omega )]' \) by a duality argument.

Lemma 5.3

For any \(\epsilon >0\), there exists a positive constant \({C_\epsilon }\) independent of h such that

$$\begin{aligned} |(\rho -\rho _h,\chi )|\le {C_\epsilon }h^{\min ((\pi /\omega )-\epsilon ,k)}\Vert \chi \Vert _{H^1(\Omega )} \qquad \forall \,\chi \in H^1(\Omega ). \end{aligned}$$
(5.2)

Proof

In view of (3.10), (4.1) and the fact that \(\rho ,\rho _h\in L_2^0(\Omega )\), we have

$$\begin{aligned} \Vert \mathrm {curl}\,\rho \Vert _{L_2(\Omega )}\le \Vert \varvec{f}\Vert _{L_2(\Omega )}, \quad \Vert \mathrm {curl}\,\rho _h\Vert _{L_2(\Omega )}\le \Vert \varvec{f}\Vert _{L_2(\Omega )}, \end{aligned}$$
(5.3)

and a Galerkin orthogonality relation

$$\begin{aligned} \big (\mathrm {curl}\,(\rho -\rho _h),\mathrm {curl}\,\psi )=0 \qquad \forall \,\psi \in V_h. \end{aligned}$$
(5.4)

Let \(\chi \in H^1(\Omega )\) be arbitrary and \(\lambda \in H^1(\Omega )\) be defined by

$$\begin{aligned} (\mathrm {curl}\,\psi ,\mathrm {curl}\,\lambda )+(\psi ,1)(\lambda ,1)=(\psi ,\chi ) \qquad \forall \,\psi \in H^1(\Omega ). \end{aligned}$$
(5.5)

Then we have

$$\begin{aligned} (\rho -\rho _h,\chi )=\big (\mathrm {curl}\,(\rho -\rho _h),\mathrm {curl}\,\lambda \big )= \big (\mathrm {curl}\,(\rho -\rho _h),\mathrm {curl}\,(\lambda -\psi )\big ) \qquad \forall \,\psi \in V_h \end{aligned}$$

by (5.4), (5.5) and the fact that \(\rho ,\rho _h\in L_2^0(\Omega )\), which implies

$$\begin{aligned} |(\rho -\rho _h,\chi )|\le \Vert \mathrm {curl}\,(\rho -\rho _h)\Vert _{L_2(\Omega )}\inf _{\psi \in V_h}|\lambda -\psi |_{H^1(\Omega )}. \end{aligned}$$
(5.6)

According to Lemma 2.4 and (5.5), we have \(\lambda \in H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for any \(\epsilon >0\) and also \( \Vert \lambda \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}\le {C_\epsilon }\Vert \chi \Vert _{H^1(\Omega )}\). Lemma 5.1 then implies

$$\begin{aligned} \inf _{\psi \in V_h}|\lambda -\psi |_{H^1(\Omega )}\le Ch^{\min ({(\pi /\omega )-\epsilon },k)}\Vert \lambda \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )} \le {C_\epsilon }h^{\min ({(\pi /\omega )-\epsilon },k)}\Vert \chi \Vert _{H^1(\Omega )}.\nonumber \\ \end{aligned}$$
(5.7)

The estimate (5.2) follows from (5.3), (5.6) and (5.7). \(\square \)

Next we estimate \(|\xi -\xi _h|_{H^1(\Omega )}\).

Let \(\tilde{\xi }_{0,h}\in \mathring{V}_h\) be defined by

$$\begin{aligned} (\mathrm {curl}\,\tilde{\xi }_{0,h},\mathrm {curl}\,\eta )+\beta (\tilde{\xi }_{0,h},\eta )=(\rho ,\eta ) \qquad \forall \,\eta \,\in \mathring{V}_h. \end{aligned}$$
(5.8)

On one hand we have

$$\begin{aligned} \big (\mathrm {curl}\,(\tilde{\xi }_{0,h}-\xi _{0,h}),\mathrm {curl}\,\eta \big )+\beta \big (\tilde{\xi }_{0,h}-\xi _{0,h},\eta \big )=(\rho -\rho _h,\eta )\qquad \forall \,\eta \in \mathring{V}_h \end{aligned}$$

by comparing (4.3) and (5.8). It follows that

$$\begin{aligned} |\tilde{\xi }_{0,h}-\xi _{0,h}|_{H^1(\Omega )}^2\le (\rho -\rho _h,\tilde{\xi }_{0,h}-\xi _{0,h}), \end{aligned}$$

which together with (5.2) and a standard Poncaré-Friedrichs inequality implies

$$\begin{aligned} |\tilde{\xi }_{0,h}-\xi _{0,h}|_{H^1(\Omega )}\le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}}. \end{aligned}$$
(5.9)

On the other hand \(\tilde{\xi }_{0,h}\) is the Galerkin finite element approximation of \(\xi _0\) (cf. (3.12) and (5.8)). Therefore we have

$$\begin{aligned} |\xi _0-\tilde{\xi }_{0,h}|_{H^1(\Omega )}\le C\inf _{\eta \in \mathring{V}_h}|\xi _0-\eta |_{H^1(\Omega )} \end{aligned}$$
(5.10)

by Céa’s lemma [13, 16].

According to Lemma 2.3 and (3.12), we have \(\xi _0\in H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) and \( \Vert \xi _0\Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}\le {C_\epsilon }\Vert \rho \Vert _{H^1(\Omega )}\). It then follows from Lemma 5.1 that

$$\begin{aligned} \inf _{\eta \in \mathring{V}_h}|\xi _0-\eta |_{H^1(\Omega )}\le C h^{{\min ({(\pi /\omega )-\epsilon },k)}}\Vert \xi _0\Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}\le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}}.\quad \end{aligned}$$
(5.11)

Putting (5.9)–(5.11) together we obtain

$$\begin{aligned} |\xi _0-\xi _{0,h}|_{H^1(\Omega )}\le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}}. \end{aligned}$$
(5.12)

Similarly, since \(\xi _{1,h}\) is the Galerkin finite element approximation of \(\xi _1\) (cf. (3.13) and (4.4)), we have

$$\begin{aligned} |\xi _1-\xi _{1,h}|_{H^1(\Omega )}\le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}}. \end{aligned}$$
(5.13)

The estimate (5.1) follows from (3.11), (4.2), (5.12) and (5.13).

5.1.2 The Case Where \({\gamma }>0\)

The error analysis follows the ideas in Sect. 5.1.1 within the setting of the coupled problem (3.14a), (3.14b).

First we use a duality argument to estimate the error for \(\zeta _{0,h}\) in the norm of \([H^1(\Omega )]'\). In view of (3.19), (3.21) and (4.6), we have

$$\begin{aligned} \Vert \zeta _0\Vert _{H^1(\Omega )}+ |\xi _0|_{H^1(\Omega )}\le C\Vert \varvec{f}\Vert _{L_2(\Omega )}, \quad \Vert \zeta _{0,h}\Vert _{H^1(\Omega )}+|\xi _{0,h}|_{H^1(\Omega )}\le C\Vert \varvec{f}\Vert _{L_2(\Omega )},\nonumber \\ \end{aligned}$$
(5.14)

and the Galerkin orthogonality relation

$$\begin{aligned} \mathcal {A}\big ((\zeta _0-\zeta _{0,h},\xi _0-\xi _{0,h}),(\psi ,\eta )\big )+(\zeta _0-\zeta _{0,h},1)(\psi ,1) =0 \qquad \forall \,(\psi ,\eta )\in V_h\times \mathring{V}_h.\nonumber \\ \end{aligned}$$
(5.15)

Let \(\chi \in H^1(\Omega )\) be arbitrary and \((\lambda ,\mu )\in H^1(\Omega )\times H^1_0(\Omega )\) be defined by

$$\begin{aligned} \mathcal {A}\big ((\psi ,\eta ),(\lambda ,\mu )\big ) +(\psi ,1)(\lambda ,1)=(\psi ,\chi ) \qquad \forall \,(\psi ,\eta )\in H^1(\Omega )\times H^1_0(\Omega ). \end{aligned}$$
(5.16)

Then the function \(\zeta _0-\zeta _{0,h}\in H^1(\Omega )\) satisfies, by (5.15) and (5.16),

$$\begin{aligned} (\zeta _0-\zeta _{0,h},\chi )&=\mathcal {A}\big ((\zeta _0-\zeta _{0,h},\xi _0-\xi _{0,h}),(\lambda ,\mu ) \Big ) +(\zeta -\zeta _{0,h},1)(\lambda ,1)\\&=\mathcal {A}\big ((\zeta _0-\zeta _{0,h},\xi _0-\xi _{0,h}),(\lambda -\psi ,\mu -\eta ) \Big )+(\zeta -\zeta _{0,h},1)(\lambda -\psi ,1) \end{aligned}$$

for all \((\psi ,\eta )\in V_h\times \mathring{V}_h\), and hence

$$\begin{aligned} |(\zeta _0-\zeta _{0,h},\chi )|&\le C\big (\Vert \zeta _0-\zeta _{0,h}\Vert _{H^1(\Omega )}+|\xi _0-\xi _{0,h}|_{H^1(\Omega )}\big )\\&\times \inf _{(\psi ,\eta )\in V_h\times \mathring{V}_h}\big (\Vert \lambda -\psi \Vert _{H^1(\Omega )}+|\mu -\eta |_{H^1(\Omega )}\big ). \nonumber \end{aligned}$$
(5.17)

Observe that the well-posedness of (5.16) implies

$$\begin{aligned} \Vert \lambda \Vert _{H^1(\Omega )}+\Vert \mu \Vert _{H^1(\Omega )}\le C\Vert \chi \Vert _{L_2(\Omega )}. \end{aligned}$$
(5.18)

It then follows from Lemmas 2.3, 2.4 and the relations (cf. (3.16) and (5.16))

$$\begin{aligned} (\mathrm {curl}\,\psi ,\mathrm {curl}\,\lambda )+(\psi ,1)(\lambda ,1)= & {} \gamma ^\frac{1}{2}(\psi ,\mu )+(\psi ,\chi ) \qquad \forall \,\psi \in H^1(\Omega ),\\ (\mathrm {curl}\,\eta ,\mathrm {curl}\,\mu )+\beta (\eta ,\mu )= & {} -\gamma ^\frac{1}{2}(\lambda ,\eta )\qquad \forall \,\eta \in H^1_0(\Omega ), \end{aligned}$$

that \((\lambda ,\mu )\in H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\times H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) and, because of (5.18),

$$\begin{aligned} \Vert \lambda \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}+\Vert \mu \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}\le {C_\epsilon }\Vert \chi \Vert _{H^1(\Omega )}. \end{aligned}$$

Hence Lemma 5.1 implies

$$\begin{aligned} \inf _{(\psi ,\eta )\in V_h\times \mathring{V}_h}\big (\Vert \lambda -\psi \Vert _{H^1(\Omega )}+|\mu -\eta |_{H^1(\Omega )}\big ) \le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}} \Vert \chi \Vert _{H^1(\Omega )}. \end{aligned}$$
(5.19)

Putting (5.14), (5.17) and (5.19) together, we see that

$$\begin{aligned} |(\zeta _0-\zeta _{0,h},\chi )|\le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}}\Vert \chi \Vert _{H^1(\Omega )} \qquad \forall \,\chi \in H^1(\Omega ). \end{aligned}$$
(5.20)

Next we compare the equation

$$\begin{aligned} (\mathrm {curl}\,\xi _0,\mathrm {curl}\,\eta )+\beta (\xi _0,\eta )=\gamma ^\frac{1}{2}(\zeta ,\eta )\qquad \forall \,\eta \in H^1_0(\Omega ), \end{aligned}$$

that is a part of (3.19) with the equation

$$\begin{aligned} (\mathrm {curl}\,\xi _{0,h},\mathrm {curl}\,\eta )+\beta (\xi _{0,h},\eta )=\gamma ^\frac{1}{2}(\zeta _{0,h},\eta )\qquad \forall \,\eta \in \mathring{V}_h \end{aligned}$$

that is a part of (4.6). Using (5.20) and the arguments in the derivation of(5.12), we find

$$\begin{aligned} |\xi _h-\xi _{0,h}|_{H^1(\Omega )}\le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}}. \end{aligned}$$
(5.21)

Similarly, we have

$$\begin{aligned} |\xi _1-\xi _{1,h}|_{H^1(\Omega )}\le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}} \end{aligned}$$
(5.22)

by comparing (3.20) and (4.7).

The estimate (5.1) follows from (3.18), (4.5), (5.21) and (5.22).

5.2 Error Analysis for \(\varvec{\phi }_{\varvec{h}}\)

The error analysis for \(\phi _h\) is similar to the error analysis for \(\xi _h\) in Sect. 5.1.1.

Lemma 5.4

For any \(\epsilon >0\), there exists a positive constant \({C_\epsilon }\) independent of h such that

$$\begin{aligned} |\phi -\phi _h|_{H^1(\Omega )}\le {C_\epsilon }h^{\min ({(\pi /\omega )-\epsilon },k)}. \end{aligned}$$
(5.23)

Proof

It follows from Lemma 2.4 and (3.3) that \(\phi \in H^{1+{(\pi /\omega )-\epsilon }}(\Omega )\) for any \(\epsilon >0\) and \(\Vert \phi \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )}\le {C_\epsilon }\Vert \xi \Vert _{H^1(\Omega )}\).

Let the function \(\tilde{\phi }_h\in V_h\) be defined by

$$\begin{aligned} (\mathrm {curl}\,\tilde{\phi }_h,\mathrm {curl}\,\psi )+(\tilde{\phi }_h,1)(\psi ,1)=(\xi ,\psi )\qquad \forall \,\psi \in V_h. \end{aligned}$$
(5.24)

On one hand we have, by comparing (3.3) and (5.24),

$$\begin{aligned} |\tilde{\phi }_h-\phi _h|_{H^1(\Omega )}\le C\Vert \xi -\xi _h\Vert _{L_2(\Omega )}. \end{aligned}$$
(5.25)

On the other hand \(\tilde{\phi }_h\in V_h\) is the Galerkin \(P_k\) finite element approximation of \(\phi \), and the orthogonality relation

$$\begin{aligned} \big (\mathrm {curl}\,(\phi -\tilde{\phi }_h),\mathrm {curl}\,\psi \big )=0 \qquad \forall \,\psi \in V_h \end{aligned}$$

together with Lemma 5.1 implies that

$$\begin{aligned} |\phi -\tilde{\phi }_h|_{H^1(\Omega )}&\le \inf _{\psi \in V_h}|\phi -\psi |_{H^1(\Omega )}\nonumber \\&\le Ch^{{\min ({(\pi /\omega )-\epsilon },k)}}\Vert \phi \Vert _{H^{1+{(\pi /\omega )-\epsilon }}(\Omega )} \le {C_\epsilon }h^{{\min ({(\pi /\omega )-\epsilon },k)}}\Vert \xi \Vert _{H^1(\Omega )}. \end{aligned}$$
(5.26)

The estimate (5.23) follows from (5.25), (5.26) and the estimate in Lemma 5.2 for \(\xi -\xi _h\in H^1_0(\Omega )\). \(\square \)

5.3 Error Analysis for \(\varvec{\varphi }_{\varvec{h}, j}\) and \(\varvec{c}_{\varvec{h},j}\)

The following result can be found in [9, Lemmas 4.6 and 4.7].

Lemma 5.5

In the case where \(\Omega \) is not simply connected, we have

$$\begin{aligned} |\varphi _j-\varphi _{j,h}|_{H^1(\Omega )}+|c_j-c_{j,h}|\le Ch^{\pi /\omega }\qquad \text {for}\;1\le j\le m. \end{aligned}$$

5.4 Convergence Results

In view of Lemmas 5.2, 5.4, 5.5, (3.1) and (4.9), we immediately have the following result.

Theorem 5.6

The approximations \(\xi _h\) and \(\varvec{u}_h\) obtained by the \(P_k\) finite element method satisfy

$$\begin{aligned} \Vert \varvec{u}-\varvec{u}_h\Vert _{L_2(\Omega )}+ |\mathrm {curl}\,\varvec{u}-\xi _h|_{H^1(\Omega )}\le {C_\epsilon }h^{\min ({(\pi /\omega )-\epsilon },k)}, \end{aligned}$$
(5.27)

for any \(\epsilon >0\), where \(\omega \) is the largest angle at the corners of \(\Omega \).

Remark 5.7

In the case where \((\pi /\omega )\) is not an integer, a more detailed analysis that takes into account the nature of the singularities at the corners of \(\Omega \) (cf. [6]) shows that the \(\epsilon \) in (5.27) can be removed.

6 Numerical Results

In this section we report the results of numerical experiments for three different domains: the unit square, a nonconvex but simply connected domain and a domain whose Betti number is 1. We use quasi-uniform meshes in all the experiments.

Experiment 6.1

In the first experiment the domain \(\Omega \) is the unit square \((0,1)\times (0,1)\). We take \(\beta =\gamma =0\) and the exact solution to be \(\varvec{u}=\mathrm {curl}\,\phi \) where

$$\begin{aligned} \phi (x)=\sin ^3(\pi x_1)\sin ^3(\pi x_2). \end{aligned}$$

We solve (1.3) by the \(P_1\) and \(P_2\) finite element methods. The results are presented in Tables 1 and 2. They agree with Theorem 5.6 with \(\omega =\pi /2\).

Table 1 Results for the \(P_1\) finite element method for Experiment 6.1
Table 2 Results for the \(P_2\) finite element method for Experiment 6.1

Experiment 6.2

The domain \(\Omega \) for the second experiment is also the unit square \((0,1)\times (0,1)\). We take \(\beta =\gamma =0\) and

$$\begin{aligned} \varvec{f}= \begin{bmatrix} (x_1^2+1)\sin x_1+x_1 x_2^3 + 2\\ (x_2^2+1)\cos x_1+x_1^3x_2^2 - 1 \end{bmatrix}. \end{aligned}$$

We solve (1.3) by the \(P_1\) and \(P_2\) finite element methods and report the results in Tables 3 and 4. Since the exact solution is not known, the relative errors are estimated by comparing the numerical solutions on consecutive refinement levels. The results agree with Theorem 5.6 with \(\omega =\pi /2\).

Table 3 Results for the \(P_1\) finite element method for Experiment 6.2
Table 4 Results for the \(P_2\) finite element method for Experiment 6.2

Experiment 6.3

In the third experiment we solve (1.3) on the nonconvex domain (cf. Fig. 2) whose vertices are \((0, 0), (.5, 0), (.5, .7), (1, .7), (1, 1), (0, 1), (1, .75), (.25, .75), (.25, .625) \;\text {and}\; (0, .625).\)

Fig. 2
figure 2

Domain for Experiment 6.3

We take \(\beta =\gamma =0\) and use a piecewise constant vector field \(\varvec{f}\) defined by

$$\begin{aligned} {\varvec{f}}(x)= {\left\{ \begin{array}{ll} \left[ \frac{1}{4}, \frac{5}{4}\right] ^\mathrm{t} &{}\qquad |x|<2^{-1/2}\\ \left[ \frac{1}{2} , \frac{3}{2}\right] ^\mathrm{t} &{}\qquad 2^{-1/2}\le |x|<1\\ \left[ 1,2\right] ^\mathrm{t}&{}\qquad |x|\ge 1 \end{array}\right. }. \end{aligned}$$

The estimated relative errors for the \(P_1\) and \(P_2\) finite element methods are displayed in Tables 5 and 6.

Table 5 Results for the \(P_1\) finite element method for Experiment 6.3
Table 6 Results for the \(P_2\) finite element method for Experiment 6.3

For this problem the order of convergence predicted by Theorem 5.6 is 2 / 3 (since \(\omega =3\pi /2\)). This is observed in Table 6 for the \(P_2\) finite element method, and also in Table 5 for the \(P_1\) finite element method with respect to the convergence of \(\varvec{u}_h\). On the other hand the convergence observed in Table 5 for \(\xi _h\) is pre-asymptotic.

Experiment 6.4

The domain for the fourth experiment is

$$\begin{aligned} \Omega = (0, 1) \times (0, 1) \setminus [1/4, 3/4] \times [1/4, 3/4] \end{aligned}$$

whose Betti number is 1. We take \(\beta =\gamma =1\) and use the same \(\varvec{f}\) in Experiment 6.2.

Since the domain is not simply connected, the solution \(\varvec{u}\) of (1.3) is given by

$$\begin{aligned} \varvec{u}=\mathrm {curl}\,\phi +c\,\mathrm {grad}\,\varphi , \end{aligned}$$
(6.1)

where \(\varphi \) is the harmonic function that vanishes on the outer boundary of \(\Omega \) and equals 1 on the inner boundary of \(\Omega \). The approximation for \(\varvec{u}\) is given by

$$\begin{aligned} \varvec{u}_h=\mathrm {curl}\,\phi _h+c_h\,\mathrm {grad}\,\varphi _h, \end{aligned}$$
(6.2)

where \(\varphi _h\) is the discrete analog of \(\varphi \).

We solve (1.3) by the \(P_1\) finite element method and report the results in Table 7. The order of convergence for \(\varvec{u}_h\) is observed to be 2 / 3, which agrees with Theorem 5.6 with \(\omega =3\pi /2\). The convergence of \(\xi _h\) is pre-asymptotic.

Table 7 Results for the \(P_1\) finite element method for Experiment 6.4

Note that the order of convergence for \(c_h\) is better than 2 / 3, which is due to the fact that \(\varvec{f}\) is smooth (cf. [9, Remark 4.8 and Table 4.3]).

7 Concluding Remarks

We have designed and analyzed \(P_k\) Lagrange finite element methods for a quad-curl problem that are based on the Hodge decomposition approach. For simplicity we only considered quasi-uniform meshes and the performance of the methods suffer from the existence of reentrant corners. But optimal convergence rates can be recovered if we would use properly graded meshes (cf. [17] for the case of the Maxwell equations).

Our convergence analysis does not require \(H^2\) (or higher) regularity for the exact solution that is assumed in [23, 29, 31]. The computational cost of our approach also compares favorably to those for the methods in [23, 29, 31].

Below are some related topics that can also benefit from the Hodge decomposition approach.

  • As in the case of the Maxwell equations, the Hodge decomposition approach lends itself naturally to the development of fast solvers for the quad-curl problem.

  • The Hodge decomposition approach can also be applied to the following eigenvalue problems on two dimensional domains. The first eigenvalue problem is to find \((\varvec{u},\lambda )\in {\mathbb {E}}\times \mathbb {R}\) such that

    $$\begin{aligned} \big (\mathrm {curl}\,(\mathrm {curl}\,\varvec{u}),\mathrm {curl}\,(\mathrm {curl}\,\varvec{v})\big )=\lambda (\varvec{u},\varvec{v}) \qquad \forall \,\varvec{v}\in {\mathbb {E}}\quad \text {and}\quad \varvec{u}\ne \mathbf{0}. \end{aligned}$$

    The second eigenvalue problem is to find \((\varvec{u},\lambda )\in {\mathbb {E}}\times \mathbb {R}\) such that

    $$\begin{aligned} \big (\mathrm {curl}\,(\mathrm {curl}\,\varvec{u}),\mathrm {curl}\,(\mathrm {curl}\,\varvec{v})\big )=\lambda (\mathrm {curl}\,\varvec{u},\mathrm {curl}\,\varvec{v}) \qquad \forall \,\varvec{v}\in {\mathbb {E}}\quad \text {and}\quad \varvec{u}\ne \mathbf{0}. \end{aligned}$$

    For both problems the Hodge decomposition approach reduces an elliptic eigenvalue problem for vector fields to an elliptic eigenvalue problem for scalar functions that can be solved by standard \(H^1\) conforming finite elements.

  • For three dimensional domains, the Hodge decomposition approach reduces the quad-curl problem to problems that can be solved numerically by standard \(H(\mathrm {curl})\), \(H(\mathrm {div})\) and \(H^1\) conforming finite elements. Moreover these problems have already been analyzed in [4].

These topics are being investigated in our ongoing projects.