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

Energy-preserving mixed finite element methods for the Hodge wave equation

Yongke Wu and Yanhong Bai School of Mathematical Sciences, University of Electronic Science and Technology of China, Chengdu 611731, China. wuyongke1982@sina.com School of Science, Xihua University, Chengdu 610039, China. baiyanhong1982@126.com
Abstract.

Energy-preserving numerical methods for solving the Hodge wave equation is developed in this paper. Based on the de Rham complex, the Hodge wave equation can be formulated as a first-order system and mixed finite element methods using finite element exterior calculus is used to discretize the space. A continuous time Galerkin method, which can be viewed as a modification of the Crank-Nicolson method, is used to discretize the time which results in a full discrete method preserving the energy exactly when the source term is vanished. A projection based operator is used to establish the optimal order convergence of the proposed methods. Numerical experiments are present to support the theoretical results.

Key words and phrases:
the Hodge wave equation, energy conservation, de Rham complex, optimal error estimates
2010 Mathematics Subject Classification:
65M60; 65M12; 65J08
Y. Wu was supported by the National Natural Science Foundation of China (11971094 and 11501088). Y. Bai was supported by the National Natural Science Foundation of China (11701481).

1. Introduction

We consider energy-preserving numerical methods for solving the Hodge wave equation, the hyperbolic equation in n\mathbb{R}^{n} associated to the Hodge Laplacian of differential kk-forms for 0kn0\leq k\leq n. The initial-boundary value problem we study is: Find u:(0,T]H0Λk(Ω)u:\ (0,T]\mapsto H_{0}\Lambda^{k}(\Omega) satisfying

(1) utt+(dδ+δd)u\displaystyle u_{tt}+(\,{\rm d}\delta+\delta\,{\rm d})u =fin Ω×(0,T],\displaystyle=f\qquad\text{in }\Omega\times(0,T],

with homogeneous boundary conditions

(2) tr(u)=0,tr(du)\displaystyle\operatorname*{tr}(u)=0,\quad\operatorname*{tr}(\star\,{\rm d}u) =0on Ω×(0,T],\displaystyle=0\qquad\text{on }\partial\Omega\times(0,T],

and initial conditions

(3) u(,0)=u0(),ut(,0)\displaystyle u(\cdot,0)=u_{0}(\cdot),\quad u_{t}(\cdot,0) =u1()in Ω.\displaystyle=u_{1}(\cdot)\qquad\text{in }\Omega.

Here Ωn\Omega\subset\mathbb{R}^{n} is a domain homomorphism to a ball with piecewise smooth and Lipschitz boundary. The unknown uu is a time dependent differential kk-form on Ω\Omega, utu_{t} and uttu_{tt} denote its partial derivatives with respect to time variable, and d\,{\rm d}, δ\delta, \star, and tr\operatorname*{tr} denote exterior derivative, co-derivative, Hodge star, and the trace operator, respectively; see Section 2 for precise definitions. We assume that TT is a finite positive real number denoting the ending time.

Many physical problems can be described by (1), such as the mathematical models of sound waves (n=3n=3 and k=0k=0), electromagnetic waves (n=3n=3 and k=1k=1), structural vibration (n=3n=3 and k=2k=2) and so on. There are many theoretical analyses of finite element method for (1) in the special case n=2n=2 or 33 and k=0k=0 or k=n1k=n-1; see [15, 4, 17, 11, 16, 28, 19, 21, 9, 22, 20] and the references therein. The pioneer work on mixed finite element methods [5] for the general form of the Hodge wave equation (1) can be found in Quenneville-Bélair’s Ph. D. thesis [26]; see also [3]. In this work, he has presented (1) the abstract Hodge wave equation in the mixed form, (2) the semi-discretization in space for solving the Hodge wave equation (3) the existence and uniqueness of the solution for the semi-discretization in space, (4) the error estimates in the L(L2)\|\cdot\|_{L^{\infty}(L^{2})} norm for the semi-discretization in space based on the elliptic projection operator.

In the present work, we shall give more thorough analysis of the mixed finite element method developed in [26, 3]. Introduce a (k1)(k-1)-form σ=δu\sigma=\delta u and a (k+1)(k+1)-form ω=du\omega=\,{\rm d}u with standard modification for k=0k=0 or k=nk=n, and a kk-form μ=ut\mu=u_{t}. The first order formulation of (1) reads as: find σH0Λ\sigma\in H_{0}\Lambda^{-}, μH0Λ\mu\in H_{0}\Lambda, and ωH0Λ+\omega\in H_{0}\Lambda^{+} such that

(4) σt,τdτ,μ\displaystyle\langle\sigma_{t},\tau\rangle-\langle\,{\rm d}^{-}\tau,\mu\rangle =0τH0Λ,\displaystyle=0\qquad\qquad\forall~{}~{}\tau\in H_{0}\Lambda^{-},
(5) μt,v+dσ,v+ω,dv\displaystyle\langle\mu_{t},v\rangle+\langle\,{\rm d}^{-}\sigma,v\rangle+\langle\omega,\,{\rm d}v\rangle =f,vvH0Λ,\displaystyle=\langle f,v\rangle\,\qquad\forall~{}~{}v\in H_{0}\Lambda,
(6) ωt,ϕdμ,ϕ\displaystyle\langle\omega_{t},\phi\rangle-\langle\,{\rm d}\mu,\phi\rangle =0ϕH0Λ+,\displaystyle=0\qquad\qquad\forall~{}~{}\phi\in H_{0}\Lambda^{+},

with initial conditions

σ0=δu0,μ0=u1,ω0=du0.\sigma_{0}=\delta u_{0},\quad\mu_{0}=u_{1},\quad\omega_{0}=\,{\rm d}u_{0}.

Comparing with [26], the main contributions of this paper are as follows. Firstly, we use the skew-symmetric property of the formulation (4)-(6) to get the following energy estimates

(7) sup0tTE(t)\displaystyle\sup\limits_{0\leq t\leq T}E(t) E(0)+20Tf(,s)ds,\displaystyle\leq E(0)+2\int_{0}^{T}\|f(\cdot,s)\|\,{\rm d}s,
(8) sup0tTH(t)\displaystyle\sup\limits_{0\leq t\leq T}H(t) H(0)+4fL(L2)+20Tft(,s)ds,\displaystyle\leq H(0)+4\|f\|_{L^{\infty}(L^{2})}+2\int_{0}^{T}\|f_{t}(\cdot,s)\|\,{\rm d}s,

with

E(t)\displaystyle E(t) =(σ(,t)2+μ(,t)2+ω(,t)2)1/2,\displaystyle=(\|\sigma(\cdot,t)\|^{2}+\|\mu(\cdot,t)\|^{2}+\|\omega(\cdot,t)\|^{2})^{1/2},
H(t)\displaystyle H(t) =(dσ(,t)2+dμ(,t)2+δμ(,t)2+δ+ω(,t)2)1/2.\displaystyle=(\|\,{\rm d}^{-}\sigma(\cdot,t)\|^{2}+\|\,{\rm d}\mu(\cdot,t)\|^{2}+\|\delta\mu(\cdot,t)\|^{2}+\|\delta^{+}\omega(\cdot,t)\|^{2})^{1/2}.

These energy estimates imply the existence and uniqueness of solution for (4)-(6); see Remark 2.4. When (4)-(6) is self-conserve, i.e., f=0f=0, the inequality (7)-(8) become equalities which implies the energies EE and HH are preserved exactly; see Remark 2.3. Due to the structure preserving properties of the finite element exterior calculus (FEEC) [1, 2, 3], the semi-discretization in space also inherit the skew-symmetric property of the spatial differential terms, and thus the energy conservation is preserved naturally. We then use the continuous time Galerkin method [16] to give unconditioned energy conservation schemes. Here we follow the approach in [17, 16, 22], where the energy estimates has been derived for scalar wave equations but not for Hodge wave equations. As we know, energy conservation numerical schemes can have a crucial influence on the quality of the numerical simulations. Especially, in long-time simulations, energy-preserving can have a dramatic effect on stability and global error growth.

Secondly, we obtain the optimal convergence order of the error estimates in both L2L^{2}-norm and 𝒜()L(L2)\|\mathcal{A}(\cdot)\|_{L^{\infty}(L^{2})}-norm for the semi- and full-discrete mixed finite element methods, where 𝒜\mathcal{A} is a skew-symmetric operator defined in Section 2. Such result has been derived for scalar wave equation [17, 16, 22] but generalization to general Hodge wave equation is non-trivial. Technically, the canonical interpolation operators πh\pi_{h} used in [17, 16, 22] cannot be commutated with the discrete co-derivative operator δh\delta_{h}, and the L2L^{2} projection operator QhQ_{h} cannot be commutated with the exterior derivative operators d\,{\rm d}. Using these standard operators in the convergence analysis will lead to the lost of the convergence order. To overcome this difficulty, we choose a projection based interpolation operator IhI_{h} briefly mentioned in [8, Proposition 5.44] and redefine it based on the Hodge decomposition. Such projection based operators has been introduced for H1,H(curl)H^{1},H({\rm curl\,}) and H(div)H(\operatorname{div}) spaces in [12, 13, 14, 25], where the authors have proved that these projection based operators made the de Rahm diagram commute and had the quasi-optimal interpolation error bound for hphp finite element spaces. Although this projection-based quasi-interpolation operator is not new, the properties we are going to prove are not fully explored in the literature. Specifically, we shall prove that (1) IhI_{h} is commuted with δh\delta_{h}, (2) IhI_{h} is stable in both d()\|\,{\rm d}(\cdot)\| and δh()\|\delta_{h}(\cdot)\| norms, (3) IhI_{h} is the L2L^{2} orthogonal projection to the space 0,h\mathfrak{Z}_{0,h}, (4) IhI_{h} is an orthogonal projection operator with respect to the inner-product d(),d()\langle\,{\rm d}(\cdot),\,{\rm d}(\cdot)\rangle, (5) IhI_{h} has the same approximation properties as the classical interpolation operators; see Lemma 3.3 - 3.5. By using these properties of the projection-based operator IhI_{h}, we get the optimal error estimates for both the semi- and full- discretization with respect to both L(L2)\|\cdot\|_{L^{\infty}(L^{2})} and 𝒜()L(L2)\|\mathcal{A}(\cdot)\|_{L^{\infty}(L^{2})} norms (the detail definition of these norms can be found in Section 3), while recall that [26] only give the error estimates of L(L2)\|\cdot\|_{L^{\infty}(L^{2})} for the semi-discretization in space and as the line of Quenneville’s proof, it seems difficulty to get the error estimate of the energy norm 𝒜()L(L2)\|\mathcal{A}(\cdot)\|_{L^{\infty}(L^{2})}. But the control of the energy norm is very important, since the L2L^{2}-norm is possible small but the energy norm is larger due to the small oscillation in the error. Furthermore our error estimate, comparing with [26] is robust to TT in the sense that the factor TT is absent on the error estimates; see Theorem 3.9, 3.13, 4.6 and 4.8. Such error estimates imply that our algorithms are robust for long time problems and the numerical experiment supports this result; see Table 3.

What remains of this paper is organized as follow. In Section 2 we introduce the required background on finite element exterior calculus (FEEC) and the Hodge wave equation. We obtain the mixed formulation of the Hodge wave equation and get the energy conservation estimates. Section 3, we briefly introduce the finite element spaces on kk-forms, give the semi-discrete form of the Hodge wave equation, introduce a projection-based quasi-interpolation operator and explore properties of this operator, obtain the energy estimates of the semi-discrete form, and get the optimal error estimates of the semi-discrete form. In section 4, the full-discrete form of the Hodge wave equation is obtained, the energy estimates and the optimal error estimates are obtained. Section 5 give some numerical experiments to confirm our theoretical results.

Throughout this paper, ii, hh and Δt\,\Delta t denote the time level, the mesh size and the time step size, respectively. The capital CC may be different in different places, denotes a positive constant which is independent on ii, hh and Δt\,\Delta t. We denote by m,p\|\cdot\|_{m,p} the norm of the classical Sobolev spaces Wm,pΛk(Ω)W^{m,p}\Lambda^{k}(\Omega), 1p1\leq p\leq\infty and 0kn0\leq k\leq n. If p=2p=2, we write m,p\|\cdot\|_{m,p} simply as m\|\cdot\|_{m} and denote by ||m|\cdot|_{m} the semi-norm in Wm,2Λk(Ω)W^{m,2}\Lambda^{k}(\Omega). In addition, for any Sobolev space YY, we define the space Lp([a,b],Y)L^{p}([a,b],Y) with norm fLp(Y)=(abf(,t)Ypdt)1/p\|f\|_{L^{p}(Y)}=\left(\int_{a}^{b}\|f(\cdot,t)\|_{Y}^{p}\,{\rm d}t\right)^{1/p}, and if p=p=\infty, the integral is replaced by the essential supremum.

2. Preliminaries

In this section, we follow the convention of [1, 2, 3] to introduce necessary background of finite element exterior calculus. Then, we introduce the Hodge wave equation and its mixed formulation. Finally, we get the energy conservation estimates for this mixed form.

2.1. de Rham complex

Let Ωn\Omega\subset\mathbb{R}^{n} (n2n\geq 2) be a bounded Lipschitz domain. For a given integer 0kn0\leq k\leq n, Λk(Ω)\Lambda^{k}(\Omega) represents the linear space of all smooth kk-forms on Ω\Omega. For any ωΛk(Ω)\omega\in\Lambda^{k}(\Omega), ω\omega can be written as

ω=1σ1<<σknaσdxσ1dxσk,\omega=\sum\limits_{1\leq\sigma_{1}<\cdots<\sigma_{k}\leq n}a_{\sigma}\,{\rm d}x^{\sigma_{1}}\wedge\cdots\wedge\,{\rm d}x^{\sigma_{k}},

with aσC(Ω)a_{\sigma}\in C^{\infty}(\Omega) and \wedge the wedge product. As Ω\Omega is a flat domain in n\mathbb{R}^{n}, we can identify each tangent space of Ω\Omega with n\mathbb{R}^{n}. Given an ωΛk(Ω)\omega\in\Lambda^{k}(\Omega) and vectors v1,v2,,vknv_{1},~{}v_{2},\cdots,~{}v_{k}\in\mathbb{R}^{n}, we have that the map 𝒙Ωω𝒙(v1,v2,,vk)\boldsymbol{x}\in\Omega\mapsto\omega_{\boldsymbol{x}}(v_{1},v_{2},\cdots,v_{k})\in\mathbb{R} is a smooth map (infinitely differentiable).

We define the exterior derivative dk:Λk(Ω)Λk+1(Ω)\,{\rm d}^{k}:~{}\Lambda^{k}(\Omega)\rightarrow\Lambda^{k+1}(\Omega) as

dkωx(v1,v2,,vk+1)=j=1k+1(1)j+1vjωx(v1,,v^j,,vk+1),\,{\rm d}^{k}\omega_{x}(v_{1},v_{2},\cdots,v_{k+1})=\sum\limits_{j=1}^{k+1}(-1)^{j+1}\partial_{v_{j}}\omega_{x}(v_{1},\cdots,\hat{v}_{j},\cdots,v_{k+1}),

where the hat is used to indicate a suppressed argument. By the definition of dk\,{\rm d}^{k}, it is easy to see that dk\,{\rm d}^{k} is a sequence of differential operators satisfying that the range of dk\,{\rm d}^{k} lies in the domain of dk+1\,{\rm d}^{k+1}, i.e., dk+1dk=0\,{\rm d}^{k+1}\circ\,{\rm d}^{k}=0 for k=0,1,,n1k=0,1,\cdots,n-1. For convenience of notation, we shall skip the superscript kk if there is no confusion.

Let vol be the unique volume form in Λk(Ω)\Lambda^{k}(\Omega), define the L2L^{2}-inner product of any two differential kk-forms on Ω\Omega as the integral of their pointwise inner product:

ω,μ=Ωωx,μxvol.\langle\omega,\mu\rangle=\int_{\Omega}\langle\omega_{x},\mu_{x}\rangle\text{vol}.

The completion of Λk(Ω)\Lambda^{k}(\Omega) under the corresponding norm defines the Hilbert space L2Λk(Ω)L^{2}\Lambda^{k}(\Omega). The domain of the exterior derivative dk\,{\rm d}^{k} can be enlarged to

HΛk(Ω)={ωL2Λk(Ω):dωL2Λk+1(Ω)}.H\Lambda^{k}(\Omega)=\{\omega\in L^{2}\Lambda^{k}(\Omega):~{}\,{\rm d}\omega\in L^{2}\Lambda^{k+1}(\Omega)\}.

HΛk(Ω)H\Lambda^{k}(\Omega) is a Hilbert space with inner product ω,μ+dω,dμ\langle\omega,\mu\rangle+\langle\,{\rm d}\omega,\,{\rm d}\mu\rangle and associated graph norm HΛ\|\cdot\|_{H\Lambda}. The de Rham complex

(9) HΛ0(Ω)dHΛ1(Ω)ddHΛn1(Ω)dHΛn(Ω)\begin{CD}H\Lambda^{0}(\Omega)@>{\,{\rm d}}>{}>H\Lambda^{1}(\Omega)@>{\,{\rm d}}>{}>\cdots @>{\,{\rm d}}>{}>H\Lambda^{n-1}(\Omega)@>{\,{\rm d}}>{}>H\Lambda^{n}(\Omega)\end{CD}

is then bounded in the sense that d:HΛk(Ω)HΛk+1(Ω)\,{\rm d}:~{}H\Lambda^{k}(\Omega)\rightarrow H\Lambda^{k+1}(\Omega) is a bounded operator.

For any smooth manifold MM and any xMx\in M, we use TxMT_{x}M to denote the tangential space of MM at xx. For any smooth kk-form ωΛk(Ω)\omega\in\Lambda^{k}(\Omega), we define trωΛk(M)\operatorname*{tr}\omega\in\Lambda^{k}(\partial M) as

trω(v1,v2,,vk)=ω(v1,v2,,vk)\operatorname*{tr}\omega(v_{1},v_{2},\cdots,v_{k})=\omega(v_{1},v_{2},\cdots,v_{k})

for tangential vectors viTxMTxMv_{i}\in T_{x}\partial M\subset T_{x}M (i=1,2,,ki=1,2,\cdots,k). This operator can be extended continuous to Lipschitz domain Ω\Omega, also denote by tr:H1Λk(Ω)H1/2Λk(Ω)\operatorname*{tr}:~{}H^{1}\Lambda^{k}(\Omega)\rightarrow H^{1/2}\Lambda^{k}(\partial\Omega) and tr:HΛk(Ω)H1/2Λk(Ω)\operatorname*{tr}:~{}H\Lambda^{k}(\Omega)\rightarrow H^{-1/2}\Lambda^{k}(\partial\Omega). Define

H0Λk(Ω)\displaystyle H_{0}\Lambda^{k}(\Omega) ={ωHΛk(Ω):trω=0 on Ω},\displaystyle=\{\omega\in H\Lambda^{k}(\Omega):~{}\operatorname*{tr}\omega=0\text{ on }\partial\Omega\},
H01Λk(Ω)\displaystyle H_{0}^{1}\Lambda^{k}(\Omega) ={ωH1Λk(Ω):trω=0 on Ω}.\displaystyle=\{\omega\in H^{1}\Lambda^{k}(\Omega):~{}\operatorname*{tr}\omega=0\text{ on }\partial\Omega\}.

In the following sections, we will focus on the de Rham complex with homogeneous trace

(10) H0Λ0(Ω)dH0Λ1(Ω)ddH0Λn1(Ω)dH0Λn(Ω)\begin{CD}H_{0}\Lambda^{0}(\Omega)@>{\,{\rm d}}>{}>H_{0}\Lambda^{1}(\Omega)@>{\,{\rm d}}>{}>\cdots @>{\,{\rm d}}>{}>H_{0}\Lambda^{n-1}(\Omega)@>{\,{\rm d}}>{}>H_{0}\Lambda^{n}(\Omega)\end{CD}

In order to define the dual complex, we start with the Hodge star operator :Λk(Ω)Λnk(Ω)\star:~{}\Lambda^{k}(\Omega)\rightarrow\Lambda^{n-k}(\Omega),

Ωωμ=ω,μ,ωΛk(Ω),μΛnk(Ω).\int_{\Omega}\omega\wedge\mu=\langle\star\omega,\mu\rangle,\qquad\forall~{}~{}\omega\in\Lambda^{k}(\Omega),~{}~{}\mu\in\Lambda^{n-k}(\Omega).

The coderivative operator δk:Λk(Ω)Λk1(Ω)\delta^{k}:~{}\Lambda^{k}(\Omega)\rightarrow\Lambda^{k-1}(\Omega) is defined as

δkω=(1)k(nk+1)dnkω.\delta^{k}\omega=(-1)^{k(n-k+1)}\star\,{\rm d}^{n-k}\star\omega.

dk1\,{\rm d}^{k-1} and δk\delta^{k} are related by the Stokes theorem

dω,μ=ω,δμ+Ωtrωtr(μ),ωΛk1(Ω),μΛk(Ω).\langle\,{\rm d}\omega,\mu\rangle=\langle\omega,\delta\mu\rangle+\int_{\partial\Omega}\operatorname*{tr}\omega\wedge\operatorname*{tr}(\star\mu),\qquad\omega\in\Lambda^{k-1}(\Omega),~{}~{}\mu\in\Lambda^{k}(\Omega).

We define the spaces

HΛk(Ω)\displaystyle H^{*}\Lambda^{k}(\Omega) ={ωL2Λk(Ω):δωL2Λk1(Ω)},\displaystyle=\{\omega\in L^{2}\Lambda^{k}(\Omega):~{}\delta\omega\in L^{2}\Lambda^{k-1}(\Omega)\},
H0Λk(Ω)\displaystyle H_{0}^{*}\Lambda^{k}(\Omega) ={ωHΛk(Ω):trω=0 on Ω}.\displaystyle=\{\omega\in H^{*}\Lambda^{k}(\Omega):~{}\operatorname*{tr}\star\omega=0\text{ on }\partial\Omega\}.

Treat d:H0Λk(Ω)L2Λk(Ω)L2Λk+1(Ω)\,{\rm d}:~{}H_{0}\Lambda^{k}(\Omega)\subset L^{2}\Lambda^{k}(\Omega)\rightarrow L^{2}\Lambda^{k+1}(\Omega) as an unbounded and densely defined operator. Then Stokes theorem implies that δ:HΛk+1(Ω)L2Λk+1(Ω)L2Λk(Ω)\delta:~{}H^{*}\Lambda^{k+1}(\Omega)\subset L^{2}\Lambda^{k+1}(\Omega)\rightarrow L^{2}\Lambda^{k}(\Omega) is the adjoint of d\,{\rm d} as

(11) dω,μ=ω,δμ,ωH0Λk(Ω),μHΛk+1(Ω).\langle\,{\rm d}\omega,\mu\rangle=\langle\omega,\delta\mu\rangle,\qquad\forall~{}~{}\omega\in H_{0}\Lambda^{k}(\Omega),~{}~{}\mu\in H^{*}\Lambda^{k+1}(\Omega).

We have a dual sequence of (10)

(12) HΛ0(Ω)δHΛ1(Ω)δδHΛn1(Ω)δHΛn(Ω).\begin{CD}H^{*}\Lambda^{0}(\Omega)@<{\delta}<{}<H^{*}\Lambda^{1}(\Omega)@<{\delta}<{}<\cdots @<{{\delta}}<{}<H^{*}\Lambda^{n-1}(\Omega)@<{\delta}<{}<H^{*}\Lambda^{n}(\Omega).\end{CD}

Let 0k\mathfrak{Z}_{0}^{k} be the kernel of d\,{\rm d} in the space H0Λk(Ω)H_{0}\Lambda^{k}(\Omega), then 0k\mathfrak{Z}_{0}^{k} can be decomposed as 0k=𝔅0kL20k\mathfrak{Z}_{0}^{k}=\mathfrak{B}_{0}^{k}\oplus^{\bot_{L^{2}}}\mathfrak{H}_{0}^{k}, where 𝔅0k\mathfrak{B}_{0}^{k} is the range of dk1\,{\rm d}^{k-1}, i.e., 𝔅0k=d(H0Λk1(Ω))\mathfrak{B}_{0}^{k}=\,{\rm d}(H_{0}\Lambda^{k-1}(\Omega)) and 0k\mathfrak{H}_{0}^{k} is the space of harmonic forms, i.e., 0k={ωH0Λk(Ω)HΛk(Ω):dω=0 and δω=0}\mathfrak{H}_{0}^{k}=\{\omega\in H_{0}\Lambda^{k}(\Omega)\cap H^{*}\Lambda^{k}(\Omega):~{}~{}\,{\rm d}\omega=0\text{ and }\delta\omega=0\}, L2\oplus^{\bot_{L^{2}}} means that the decomposition is orthogonal in the sense of the L2L^{2}-inner product. The following Hodge decomposition has been established in [1, page 22]:

L2Λk(Ω)=𝔅0kL20kL2δHΛk+1(Ω).L^{2}\Lambda^{k}(\Omega)=\mathfrak{B}_{0}^{k}\oplus^{\bot_{L^{2}}}\mathfrak{H}_{0}^{k}\oplus^{\bot_{L^{2}}}\delta H^{*}\Lambda^{k+1}(\Omega).

Denote 𝔎k\mathfrak{K}^{k} as the L2L^{2} orthogonal complement of 0k\mathfrak{Z}_{0}^{k} in H0Λk(Ω)H_{0}\Lambda^{k}(\Omega), i.e., 𝔎k=H0Λk(Ω)δHΛk+1(Ω)\mathfrak{K}^{k}=H_{0}\Lambda^{k}(\Omega)\cap\delta H^{*}\Lambda^{k+1}(\Omega). Then we have the Hodge decomposition of H0Λk(Ω)H_{0}\Lambda^{k}(\Omega):

(13) H0Λk(Ω)=0kL2𝔎k=𝔅0kL20kL2𝔎k.H_{0}\Lambda^{k}(\Omega)=\mathfrak{Z}_{0}^{k}\oplus^{\bot_{L^{2}}}\mathfrak{K}^{k}=\mathfrak{B}_{0}^{k}\oplus^{\bot_{L^{2}}}\mathfrak{H}_{0}^{k}\oplus^{\bot_{L^{2}}}\mathfrak{K}^{k}.

It should be point out that when k=0k=0, we have 00={0}\mathfrak{Z}_{0}^{0}=\{0\} and 𝔎1={0}\mathfrak{K}^{-1}=\{0\}. When k=nk=n, we have 𝔎n={0}\mathfrak{K}^{n}=\{0\}.

In the following sections, when spaces of the consecutive differential forms are involved, we use the short sequences

(14) H0Λ(Ω)dH0Λ(Ω)dH0Λ+(Ω)\begin{CD}H_{0}\Lambda^{-}(\Omega)@>{{\,{\rm d}^{-}}}>{}>H_{0}\Lambda(\Omega)@>{\,{\rm d}}>{}>H_{0}\Lambda^{+}(\Omega)\end{CD}

or the one with the Hodge decomposition

(15) 𝔅0L20L2𝔎d𝔅0L20L2𝔎d𝔅0+L20+L2𝔎+.\begin{CD}\mathfrak{B}_{0}^{-}\oplus^{\bot_{L^{2}}}\mathfrak{H}_{0}^{-}\oplus^{\bot_{L^{2}}}\mathfrak{K}^{-}@>{{\,{\rm d}^{-}}}>{}>\mathfrak{B}_{0}\oplus^{\bot_{L^{2}}}\mathfrak{H}_{0}\oplus^{\bot_{L^{2}}}\mathfrak{K}@>{{\,{\rm d}}}>{}>\mathfrak{B}_{0}^{+}\oplus^{\bot_{L^{2}}}\mathfrak{H}_{0}^{+}\oplus^{\bot_{L^{2}}}\mathfrak{K}^{+}.\end{CD}

In this paper, we consider the domain Ω\Omega with zero Betti numbers, namely, we impose the following assumption on the domain Ω\Omega:

(A):

We assume that Ω\Omega is simple in the sense that dim0k=0\dim\mathfrak{H}_{0}^{k}=0 for all 1kn11\leq k\leq n-1.

2.2. The Hodge wave equation

The Hodge wave equation reads as given f:(0,T)L2Λkf:~{}(0,T)\mapsto L^{2}\Lambda^{k}, find uH2((0,T),H0ΛkHΛk)u\in H^{2}((0,T),H_{0}\Lambda^{k}\cap H^{*}\Lambda^{k}) such that

(16) utt+u=fin Ω,u_{tt}+\mathcal{L}u=f\qquad\text{in }\Omega,

where =dδ+δ+d\mathcal{L}=\,{\rm d}^{-}\delta+\delta^{+}\,{\rm d} is called the Hodge Laplacian operator [2], with the initial conditions

(17) u(,0)=u0(),ut(,0)=u1().u(\cdot,0)=u_{0}(\cdot),\qquad u_{t}(\cdot,0)=u_{1}(\cdot).

For easy to preserve the energy exactly, we will use mixed method to discrete (16). Introduce a (k1)(k-1)-form σ=δu\sigma=\delta u and a (k+1)(k+1)-form ω=du\omega=\,{\rm d}u with standard modification for k=0k=0 or k=nk=n, and a kk-form μ=ut\mu=u_{t}. The mixed formulation [26] of the Hodge wave equation (16) is: given fL2((0,T),L2Λ)f\in L^{2}((0,T),L^{2}\Lambda), find (σ,μ,ω):(0,T]H0Λ×H0Λ×H0Λ+:=𝑾(\sigma,\mu,\omega):~{}(0,T]\mapsto H_{0}\Lambda^{-}\times H_{0}\Lambda\times H_{0}\Lambda^{+}:=\boldsymbol{W} such that

(18) σt,τdτ,μ\displaystyle\langle\sigma_{t},\tau\rangle-\langle\,{\rm d}^{-}\tau,\mu\rangle =0τH0Λ,\displaystyle=0\qquad\qquad\forall~{}~{}\tau\in H_{0}\Lambda^{-},
(19) μt,v+dσ,v+ω,dv\displaystyle\langle\mu_{t},v\rangle+\langle\,{\rm d}^{-}\sigma,v\rangle+\langle\omega,\,{\rm d}v\rangle =f,vvH0Λk,\displaystyle=\langle f,v\rangle\,\qquad\forall~{}~{}v\in H_{0}\Lambda^{k},
(20) ωt,ϕdμ,ϕ\displaystyle\langle\omega_{t},\phi\rangle-\langle\,{\rm d}\mu,\phi\rangle =0ϕH0Λ+,\displaystyle=0\qquad\qquad\forall~{}~{}\phi\in H_{0}\Lambda^{+},

with initial conditions

σ(,0)=δu0,μ(,0)=u1(),ω(,0)=du0().\sigma(\cdot,0)=\delta u_{0},\quad\mu(\cdot,0)=u_{1}(\cdot),\quad\omega(\cdot,0)=\,{\rm d}u_{0}(\cdot).

Denoted by

𝒜=(0δ0d0δ+0d0).\mathcal{A}=\begin{pmatrix}0&\delta&0\\ -\,{\rm d}^{-}&0&-\delta^{+}\\ 0&\,{\rm d}&0\end{pmatrix}.

The existence of solutions for the mixed formulation (18)-(20) can be found in [26] and it can also be obtained by Picard Theorem since the operator

𝒜:H0Λ×(H0ΛHΛ)×(H0Λ+HΛ+)L2Λ×L2Λ×L2Λ+\mathcal{A}:~{}H_{0}\Lambda^{-}\times(H_{0}\Lambda\cap H^{*}\Lambda)\times(H_{0}\Lambda^{+}\cap H^{*}\Lambda^{+})\rightarrow L^{2}\Lambda^{-}\times L^{2}\Lambda\times L^{2}\Lambda^{+}

is bounded. To prove the uniqueness of the solution, we need the energy estimates. We introduce a basic inequality.

Lemma 2.1.

([22, Lemma 1]) Suppose that a real number xx satisfies the quadratic inequality

x2γ2+βxx^{2}\leq\gamma^{2}+\beta x

for β,γ0\beta,~{}\gamma\geq 0 and β2+γ2>0\beta^{2}+\gamma^{2}>0. Then

xβ+γ.x\leq\beta+\gamma.

We define two energies of the mixed formulation (18)-(20) as

E(t)=(σ(,t)2+μ(,t)2+ω(,t)2)1/2E(t)=\left(\|\sigma(\cdot,t)\|^{2}+\|\mu(\cdot,t)\|^{2}+\|\omega(\cdot,t)\|^{2}\right)^{1/2}

and

H(t)=(dσ(,t)2+dμ(,t)2+δμ(,t)2+δ+ω(,t)2)1/2.H(t)=\left(\|\,{\rm d}^{-}\sigma(\cdot,t)\|^{2}+\|\,{\rm d}\mu(\cdot,t)\|^{2}+\|\delta\mu(\cdot,t)\|^{2}+\|\delta^{+}\omega(\cdot,t)\|^{2}\right)^{1/2}.

We have the following energy estimates.

Theorem 2.2.

Let 𝐮=(σ,μ,ω)𝐖\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal}\in\boldsymbol{W} be the solution of the mixed formulation (18)-(20). Provided fL1((0,T),L2Λ)f\in L^{1}((0,T),L^{2}\Lambda), we have the energy bound

(21) sup0sTE(t)E(0)+20Tf(,s)ds.\sup\limits_{0\leq s\leq T}E(t)\leq E(0)+2\int_{0}^{T}\|f(\cdot,s)\|\,{\rm d}s.

Furthermore, if fW1,1((0,T),L2Λ)f\in W^{1,1}((0,T),L^{2}\Lambda), we have the bound

(22) sup0tTH(t)H(0)+4fL(L2)+20Tft(,s)ds.\sup\limits_{0\leq t\leq T}H(t)\leq H(0)+4\|f\|_{L^{\infty}(L^{2})}+2\int_{0}^{T}\|f_{t}(\cdot,s)\|\,{\rm d}s.

When f=0f=0, the inequalities become equalities and thus we have the energy conservation

E(t)=E(0),H(t)=H(0),t>0.E(t)=E(0),H(t)=H(0),\quad\forall t>0.
Proof.

Taking τ=σ\tau=\sigma, v=μv=\mu and ϕ=ω\phi=\omega in (18) - (20) and adding them together, we obtain

12ddtE2(t)=f,μ.\frac{1}{2}\frac{\,{\rm d}}{\,{\rm d}t}E^{2}(t)=\langle f,\mu\rangle.

Integrate the above equation on the interval (0,s)(0,s), for any s(0,T]s\in(0,T], we have

E2(s)\displaystyle E^{2}(s) =E2(0)+20sf,μdt\displaystyle=E^{2}(0)+2\int_{0}^{s}\langle f,\mu\rangle\,{\rm d}t
E2(0)+2sup0tTE(t)0Tfdt.\displaystyle\leq E^{2}(0)+2\sup\limits_{0\leq t\leq T}E(t)\int_{0}^{T}\|f\|\,{\rm d}t.

Then (21) follows by Lemma 2.1.

Taking τ=δμt\tau=\delta\mu_{t}, v=dσtδ+ωtv=-\,{\rm d}^{-}\sigma_{t}-\delta^{+}\omega_{t} and ϕ=dμt\phi=\,{\rm d}\mu_{t} in (18)-(20), we have

dσt,μt12ddtδμ2\displaystyle\langle\,{\rm d}^{-}\sigma_{t},\mu_{t}\rangle-\frac{1}{2}\frac{\,{\rm d}}{\,{\rm d}t}\|\delta\mu\|^{2} =0,\displaystyle=0,
μt,dσtdμt,ωt12ddtdσ212ddtδ+ω2\displaystyle-\langle\mu_{t},\,{\rm d}^{-}\sigma_{t}\rangle-\langle\,{\rm d}\mu_{t},\omega_{t}\rangle-\frac{1}{2}\frac{\,{\rm d}}{\,{\rm d}t}\|\,{\rm d}^{-}\sigma\|^{2}-\frac{1}{2}\frac{\,{\rm d}}{\,{\rm d}t}\|\delta^{+}\omega\|^{2} =f,dσt+δ+ωt,\displaystyle=-\langle f,\,{\rm d}^{-}\sigma_{t}+\delta^{+}\omega_{t}\rangle,
ωt,dμt12ddtdμ2\displaystyle\langle\omega_{t},\,{\rm d}\mu_{t}\rangle-\frac{1}{2}\frac{\,{\rm d}}{\,{\rm d}t}\|\,{\rm d}\mu\|^{2} =0.\displaystyle=0.

Add the above equations together, we obtain

12ddtH2(t)=f,dσt+δ+ωt.\frac{1}{2}\frac{\,{\rm d}}{\,{\rm d}t}H^{2}(t)=\langle f,\,{\rm d}^{-}\sigma_{t}+\delta^{+}\omega_{t}\rangle.

Pick any 0sT0\leq s\leq T and integrate from 0 to ss to obtain

H2(s)\displaystyle H^{2}(s) =H2(0)+20sf,dσt+δ+ωtdt\displaystyle=H^{2}(0)+2\int_{0}^{s}\langle f,\,{\rm d}^{-}\sigma_{t}+\delta^{+}\omega_{t}\rangle\,{\rm d}t
=H2(0)+2f(,s),dσ(,s)+δ+ω(,s)2f(,0),dσ(,0)+δ+ω(,0)\displaystyle=H^{2}(0)+2\langle f(\cdot,s),\,{\rm d}^{-}\sigma(\cdot,s)+\delta^{+}\omega(\cdot,s)\rangle-2\langle f(\cdot,0),\,{\rm d}^{-}\sigma(\cdot,0)+\delta^{+}\omega(\cdot,0)\rangle
20sft,dσ+δ+ωdt\displaystyle\quad-2\int_{0}^{s}\langle f_{t},\,{\rm d}^{-}\sigma+\delta^{+}\omega\rangle\,{\rm d}t
H2(0)+sup0tTH(t)(4fL(L2)+20Tftdt).\displaystyle\leq H^{2}(0)+\sup\limits_{0\leq t\leq T}H(t)\left(4\|f\|_{L^{\infty}(L^{2})}+2\int_{0}^{T}\|f_{t}\|\,{\rm d}t\right).

Then the desired inequality (22) follows from Lemma 2.1. ∎

Remark 2.3.

When the source term ff of (16) equal 0, i.e., (16) is a self-conserve system, Theorem 2.2 implies that the mixed form (18)-(20) preserves the energies EE and HH exactly.

Remark 2.4.

Theorem 2.2 implies the uniqueness of solution of (18)-(20) in the space 𝐖\boldsymbol{W}. Together with the existence of solutions in the space H0Λ×(H0ΛHΛ)×(H0Λ+HΛ+)H_{0}\Lambda^{-}\times(H_{0}\Lambda\cap H^{*}\Lambda)\times(H_{0}\Lambda^{+}\cap H^{*}\Lambda^{+}), we obtain that (18)-(20) have a unique solution 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} in the space H0Λ×(H0ΛHΛ)×(H0Λ+HΛ+)H_{0}\Lambda^{-}\times(H_{0}\Lambda\cap H^{*}\Lambda)\times(H_{0}\Lambda^{+}\cap H^{*}\Lambda^{+}) and for any t(0,T]t\in(0,T] satisfying

(23) 𝒖t,𝒗+𝒜𝒖,𝒗=𝑭,𝒗𝒗𝑾,\langle\boldsymbol{u}_{t},\boldsymbol{v}\rangle+\langle\mathcal{A}\boldsymbol{u},\boldsymbol{v}\rangle=\langle\boldsymbol{F},\boldsymbol{v}\rangle\qquad\forall~{}~{}\boldsymbol{v}\in\boldsymbol{W},

with 𝐅=(0,f,0)\boldsymbol{F}=(0,~{}f,~{}0)^{\intercal}.

3. Semi-discretization of the Hodge wave equation

In this section, we will introduce mixed finite element methods developed in [26, 3] for the spatial discretization of the Hodge wave equation (16), and give the energy estimates and optimal error estimates.

3.1. Finite element spaces

Let 𝒯h\mathcal{T}_{h} be a shape regular triangulation of Ω\Omega. For each nn-simplex K𝒯hK\in\mathcal{T}_{h}, we define hK=|K|1/nh_{K}=|K|^{1/n} and h=maxK𝒯hhKh=\max\limits_{K\in\mathcal{T}_{h}}h_{K}. For completeness, we briefly introduce the construction of finite element spaces following [1, 3].

Denote 𝒫r(n)\mathcal{P}_{r}(\mathbb{R}^{n}) as the space of polynomials in nn variables of degree at most rr and r(n)\mathcal{H}_{r}(\mathbb{R}^{n}) as the space of homogeneous polynomial functions of degree rr. Spaces of polynomial differential forms 𝒫rΛk(n)\mathcal{P}_{r}\Lambda^{k}(\mathbb{R}^{n}) and r(n)\mathcal{H}_{r}(\mathbb{R}^{n}) can be defined by using the corresponding polynomial as the coefficients. We will suppress n\mathbb{R}^{n} from the notation for simplicity. For each integer rnr\geq n, we have the polynomial subcomplex of the de Rham complex

0𝒫rΛ0d𝒫r1Λ1dd𝒫rnΛn0.\setcounter{MaxMatrixCols}{11}\begin{CD}0@>{}>{}>\mathcal{P}_{r}\Lambda^{0}@>{\,{\rm d}}>{}>\mathcal{P}_{r-1}\Lambda^{1}@>{\,{\rm d}}>{}>\cdots @>{\,{\rm d}}>{}>\mathcal{P}_{r-n}\Lambda^{n}@>{}>{}>0.\end{CD}

Given a point xnx\in\mathbb{R}^{n}, treat xx as a vector in the tangential space TxnT_{x}\mathbb{R}^{n} and define the Koszul operator κ:Λk(n)Λk1(n)\kappa:~{}\Lambda^{k}(\mathbb{R}^{n})\rightarrow\Lambda^{k-1}(\mathbb{R}^{n}) as

(κω)x(v1,v2,,vk1)=ωx(x,v1,v2,,vk1).(\kappa\omega)_{x}(v_{1},v_{2},\cdots,v_{k-1})=\omega_{x}(x,v_{1},v_{2},\cdots,v_{k-1}).

This κ\kappa satisfying the identity κd+dκ=(k+r)id\kappa\,{\rm d}+\,{\rm d}\kappa=(k+r){\rm id} [1, Theorem 3.1] on the space rΛk\mathcal{H}_{r}\Lambda^{k} and there is a direct sum

rΛk=κr1Λk+1dr+1Λk1.\mathcal{H}_{r}\Lambda^{k}=\kappa\mathcal{H}_{r-1}\Lambda^{k+1}\oplus\,{\rm d}\mathcal{H}_{r+1}\Lambda^{k-1}.

Based on the decomposition, the incomplete polynomial differential form can be introduced as

𝒫rΛk=𝒫r1Λk+κr1Λk+1\mathcal{P}_{r}^{-}\Lambda^{k}=\mathcal{P}_{r-1}\Lambda^{k}+\kappa\mathcal{H}_{r-1}\Lambda^{k+1}

and, for r1r\geq 1, have the following subcomplex of the de Rham complex

0𝒫rΛ0d𝒫rΛ1dd𝒫rΛn0.\setcounter{MaxMatrixCols}{11}\begin{CD}0@>{}>{}>\mathcal{P}_{r}^{-}\Lambda^{0}@>{\,{\rm d}}>{}>\mathcal{P}_{r}^{-}\Lambda^{1}@>{\,{\rm d}}>{}>\cdots @>{\,{\rm d}}>{}>\mathcal{P}_{r}^{-}\Lambda^{n}@>{}>{}>0.\end{CD}

For each simplex K𝒯hK\in\mathcal{T}_{h}, denote 𝒫rΛk(K)\mathcal{P}_{r}\Lambda^{k}(K) or 𝒫rΛk(K)\mathcal{P}_{r}^{-}\Lambda^{k}(K) as the spaces of kk forms obtained by restricting the forms 𝒫rΛk(n)\mathcal{P}_{r}\Lambda^{k}(\mathbb{R}^{n}) or 𝒫rΛk(n)\mathcal{P}_{r}^{-}\Lambda^{k}(\mathbb{R}^{n}), respectively, to KK. We then obtain the finite element spaces

𝒫rΛk(𝒯h)\displaystyle\mathcal{P}_{r}\Lambda^{k}(\mathcal{T}_{h}) ={ωHΛk(Ω):ω|K𝒫rΛk(K),K𝒯h},\displaystyle=\{\omega\in H\Lambda^{k}(\Omega):~{}~{}\omega|_{K}\in\mathcal{P}_{r}\Lambda^{k}(K),~{}~{}\forall~{}~{}K\in\mathcal{T}_{h}\},
𝒫rΛk(𝒯h)\displaystyle\mathcal{P}_{r}^{-}\Lambda^{k}(\mathcal{T}_{h}) ={ωHΛk(Ω):ω|K𝒫rΛk(K),K𝒯h}.\displaystyle=\{\omega\in H\Lambda^{k}(\Omega):~{}~{}\omega|_{K}\in\mathcal{P}_{r}^{-}\Lambda^{k}(K),~{}~{}\forall~{}~{}K\in\mathcal{T}_{h}\}.

We choose Vhk=𝒫rΛk(𝒯h)H0ΛkV_{h}^{k}=\mathcal{P}_{r}\Lambda^{k}(\mathcal{T}_{h})\cap H_{0}\Lambda^{k} or Vhk=𝒫rΛk(𝒯h)H0ΛkV_{h}^{k}=\mathcal{P}_{r}^{-}\Lambda^{k}(\mathcal{T}_{h})\cap H_{0}\Lambda^{k} so that (Vhk,d)(V_{h}^{k},\,{\rm d}) forms a subcomplex of (H0Λk(Ω),d)(H_{0}\Lambda^{k}(\Omega),\,{\rm d}). For the consecutive spaces, we shall use short sequence

VhdVhdVh+.\begin{CD}V_{h}^{-}@>{\,{\rm d}^{-}}>{}>V_{h}@>{\,{\rm d}}>{}>V_{h}^{+}.\end{CD}

The discrete coderivative δh:VhVh\delta_{h}:~{}V_{h}\rightarrow V_{h}^{-} is defined as the L2L^{2}-adjoint of d:VhVh\,{\rm d}^{-}:~{}V_{h}^{-}\rightarrow V_{h}, i.e., for any given ωhVh\omega_{h}\in V_{h}, δhωh\delta_{h}\omega_{h} is the unique element in VhV_{h}^{-} such that

(24) δhωh,vh=ωh,dvhvhVh.\langle\delta_{h}\omega_{h},v_{h}\rangle=\langle\omega_{h},\,{\rm d}^{-}v_{h}\rangle\qquad\forall~{}~{}v_{h}\in V_{h}^{-}.

The discrete Hodge decomposition of VhkV_{h}^{k} is

(25) Vhk=0,hL2𝔎h,V_{h}^{k}=\mathfrak{Z}_{0,h}\oplus^{\bot_{L^{2}}}\mathfrak{K}_{h},

where 0,h=ker(d)Vh=d𝑽h0\mathfrak{Z}_{0,h}=\ker(\,{\rm d})\cap V_{h}=\,{\rm d}^{-}\boldsymbol{V}_{h}^{-}\subset\mathfrak{Z}_{0} and 𝔎h=δh+Vh+\mathfrak{K}_{h}=\delta_{h}^{+}V_{h}^{+} is the L2L^{2} orthogonal complement of h,0\mathfrak{Z}_{h,0} in VhV_{h}. Generally 𝔎h𝔎\mathfrak{K}_{h}\not\subset\mathfrak{K}, since δh\delta_{h} is not a conforming discretization of δ\delta. It should be point out that when k=0k=0, we have 0,h={0}\mathfrak{Z}_{0,h}=\{0\} and 𝔎h={0}\mathfrak{K}_{h}^{-}=\{0\}. When k=nk=n, we have 𝔎h={0}\mathfrak{K}_{h}=\{0\}.

We have the following discrete Poincaré inequality; cf. [1, Theorem 5.11]

Lemma 3.1 (discrete Poincaré inequality for d\,{\rm d}).

There is a positive constant CpC_{p}, independent of hh, such that

(26) ωhCpdωhωh𝔎h.\|\omega_{h}\|\leq C_{p}\|\,{\rm d}\omega_{h}\|\qquad\forall~{}~{}\omega_{h}\in\mathfrak{K}_{h}.

Since δh\delta_{h} is the adjoint operator of d:VhVh\,{\rm d}^{-}:~{}V_{h}^{-}\rightarrow V_{h}, we have the following discrete Poincaré inequality for δh\delta_{h} as well; cf. [7] and [6, Lemma 4].

Lemma 3.2 (discrete Poincaré inequality for δh\delta_{h}).

Let CpC_{p} be the constant in (26). Then we have

ωhCpδhωhωh0,h.\|\omega_{h}\|\leq C_{p}\|\delta_{h}\omega_{h}\|\qquad\forall~{}~{}\omega_{h}\in\mathfrak{Z}_{0,h}.

3.2. A projection-based quasi-interpolation operator

In this section, we introduce a projection-based quasi-interpolation operator briefly mentioned in [8, Proposition 5.44] which is a generalization of projection based operators introduced for H1,H(curl)H^{1},H({\rm curl\,}) and H(div)H(\operatorname{div}) spaces in [12, 13, 14, 25]. We redefine this operator based on the Hodge decomposition and prove more properties of this operator: it is commuted with δh\delta_{h}, stable in both d()\|\,{\rm d}(\cdot)\| and δh()\|\delta_{h}(\cdot)\| norms, a L2L^{2} orthogonal projection to the space 0,h\mathfrak{Z}_{0,h}, an orthogonal projection operator in the inner product d(),d()\langle\,{\rm d}(\cdot),\,{\rm d}(\cdot)\rangle on the subspace 𝔎h\mathfrak{K}_{h}, and has the same approximation properties as the classical interpolation operators.

For any given vH0Λ(Ω)v\in H_{0}\Lambda(\Omega), define Phv𝔎hP_{h}v\in\mathfrak{K}_{h} such that

(27) dPhv,dϕh=dv,dϕhϕh𝔎h.\langle\,{\rm d}P_{h}v,\,{\rm d}\phi_{h}\rangle=\langle\,{\rm d}v,\,{\rm d}\phi_{h}\rangle\qquad\forall~{}~{}\phi_{h}\in\mathfrak{K}_{h}.

Equation (27) determines Phv𝔎hP_{h}v\in\mathfrak{K}_{h} uniquely since the Poincaré inequality (26) implies d(),d()\langle\,{\rm d}(\cdot),\,{\rm d}(\cdot)\rangle is an inner product on the subspace 𝔎h\mathfrak{K}_{h}.

For any vH0Λ(Ω)v\in H_{0}\Lambda(\Omega), the Hodge decomposition (13) implies that there exist v1𝔎v_{1}\in\mathfrak{K}^{-} and v2𝔎v_{2}\in\mathfrak{K} such that

v=dv1L2v2v=\,{\rm d}^{-}v_{1}\oplus^{\bot_{L^{2}}}v_{2}

The projection-based quasi-interpolation operator Ih:H0Λ(Ω)VhI_{h}:~{}H_{0}\Lambda(\Omega)\rightarrow V_{h} is defined as:

(28) Ihv=dPhv1L2Phv2.I_{h}v=\,{\rm d}^{-}P_{h}^{-}v_{1}\oplus^{\bot_{L^{2}}}P_{h}v_{2}.

We have the following properties.

Lemma 3.3.

For any vH0Λ(Ω)v\in H_{0}\Lambda(\Omega), there hold

Ihv,dϕh=v,dϕhϕhVh,\langle I_{h}v,\,{\rm d}^{-}\phi_{h}\rangle=\langle v,\,{\rm d}^{-}\phi_{h}\rangle\qquad\forall~{}~{}\phi_{h}\in V_{h}^{-},

and

dIhv,dψh=dv,dψhψhVh.\langle\,{\rm d}I_{h}v,\,{\rm d}\psi_{h}\rangle=\langle\,{\rm d}v,\,{\rm d}\psi_{h}\rangle\qquad\forall~{}~{}\psi_{h}\in V_{h}.

Here we denote Vh1={0}V_{h}^{-1}=\{0\}.

Proof.

For any ϕhVh\phi_{h}\in V_{h}^{-}, the discrete Hodge decomposition (25) implies that there exists ϕh,1𝔎h\phi_{h,1}\in\mathfrak{K}_{h}^{-} such that dϕh=dϕh,1\,{\rm d}^{-}\phi_{h}=\,{\rm d}^{-}\phi_{h,1}, therefore

Ihv,dϕh=dPhv1,dϕh,1=dv1,dϕh,1=v,dϕ1,h=v,dϕh.\langle I_{h}v,\,{\rm d}^{-}\phi_{h}\rangle=\langle\,{\rm d}^{-}P_{h}^{-}v_{1},\,{\rm d}^{-}\phi_{h,1}\rangle=\langle\,{\rm d}^{-}v_{1},\,{\rm d}^{-}\phi_{h,1}\rangle=\langle v,\,{\rm d}^{-}\phi_{1,h}\rangle=\langle v,\,{\rm d}^{-}\phi_{h}\rangle.

For any ψhVh\psi_{h}\in V_{h}, the discrete Hodge decomposition (25) implies that there exists ψh,1𝔎h\psi_{h,1}\in\mathfrak{K}_{h} such that dψh=dψh,1\,{\rm d}\psi_{h}=\,{\rm d}\psi_{h,1},

dIhv,dψh=dPhv2,dψh,1=dv2,dψh,1=dv,dψh.\langle\,{\rm d}I_{h}v,\,{\rm d}\psi_{h}\rangle=\langle\,{\rm d}P_{h}v_{2},\,{\rm d}\psi_{h,1}\rangle=\langle\,{\rm d}v_{2},\,{\rm d}\psi_{h,1}\rangle=\langle\,{\rm d}v,\,{\rm d}\psi_{h}\rangle.

Then, the desired results are obtained. ∎

We have the following stability results of IhI_{h}.

Lemma 3.4.

We have the following stability results of IhI_{h}:

  1. (1)

    For any vH0Λ(Ω)v\in H_{0}\Lambda(\Omega), there holds

    dIhvdv.\|\,{\rm d}I_{h}v\|\leq\|\,{\rm d}v\|.
  2. (2)

    For any vH0Λ(Ω)HΛ(Ω)v\in H_{0}\Lambda(\Omega)\cap H^{*}\Lambda(\Omega), it holds

    δhIhv=Qhδv,\delta_{h}I_{h}v=Q_{h}^{-}\delta v,

    where Qh:L2Λ(Ω)VhQ_{h}:~{}L^{2}\Lambda(\Omega)\rightarrow V_{h} is the L2L^{2} projection operator. Therefore

    δhIhvδv.\|\delta_{h}I_{h}v\|\leq\|\delta v\|.
Proof.

(1) For any vH0Λ(Ω)v\in H_{0}\Lambda(\Omega), there exist v1𝔎v_{1}\in\mathfrak{K}^{-} and v2𝔎v_{2}\in\mathfrak{K} such that

v=dv1L2v2.v=\,{\rm d}^{-}v_{1}\oplus^{\bot_{L^{2}}}v_{2}.

We have

dIhv=dPhv2dv2=dv.\displaystyle\|\,{\rm d}I_{h}v\|=\|\,{\rm d}P_{h}v_{2}\|\leq\|\,{\rm d}v_{2}\|=\|\,{\rm d}v\|.

(2) For any vH0Λ(Ω)HΛ(Ω)v\in H_{0}\Lambda(\Omega)\cap H^{*}\Lambda(\Omega). Using the facts that Phv2δh+Vh+P_{h}v_{2}\in\delta_{h}^{+}V_{h}^{+}, δhδh+=0\delta_{h}\delta_{h}^{+}=0 and δhIhv𝔎h\delta_{h}I_{h}v\in\mathfrak{K}_{h}^{-}, for any ϕh𝔎h\phi_{h}\in\mathfrak{K}_{h}^{-}, we have

δhIhv,ϕh\displaystyle\langle\delta_{h}I_{h}v,\phi_{h}\rangle =δhdPhv1,ϕh=dPhv1,dϕh\displaystyle=\langle\delta_{h}\,{\rm d}^{-}P_{h}^{-}v_{1},\phi_{h}\rangle=\langle\,{\rm d}^{-}P_{h}^{-}v_{1},\,{\rm d}^{-}\phi_{h}\rangle
=dv1,dϕh=v,dϕh\displaystyle=\langle\,{\rm d}^{-}v_{1},\,{\rm d}^{-}\phi_{h}\rangle=\langle v,\,{\rm d}^{-}\phi_{h}\rangle
=δv,ϕh=Qhδv,ϕh.\displaystyle=\langle\delta v,\phi_{h}\rangle=\langle Q_{h}^{-}\delta v,\phi_{h}\rangle.

Using the orthogonality result 0,h𝔎h\mathfrak{Z}_{0,h}^{-}\bot\mathfrak{K}_{h}^{-}, we get the desired result. ∎

To get approximation properties of the projection-based quasi-interpolation operator IhI_{h}, we need the de Rham complexes for smooth differential forms established in [10] and the following Sobolev embedding result

(29) H0ΛHΛH1Λ,H_{0}\Lambda\cap H^{*}\Lambda\hookrightarrow H^{1}\Lambda,

which holds when Ω\Omega is convex Lipschitz domain.

Lemma 3.5.

Assume that Ω\Omega is smooth enough such that (29) holds, then for any vH0Λ(Ω)Hr+1Λ(Ω)v\in H_{0}\Lambda(\Omega)\cap H^{r+1}\Lambda(\Omega) with r1r\geq 1, we have

(30) vIhv\displaystyle\|v-I_{h}v\| hlvlfor1lr,\displaystyle\lesssim h^{l}\|v\|_{l}\qquad\text{for}\quad 1\leq l\leq r,
(31) d(vIhv)\displaystyle\|\,{\rm d}(v-I_{h}v)\| hldvlfor1lr.\displaystyle\lesssim h^{l}\|\,{\rm d}v\|_{l}\qquad\text{for}\quad 1\leq l\leq r.
Proof.

We shall use the de Rham sequence in [10]. Then for any vH0Λ(Ω)HrΛ(Ω)v\in H_{0}\Lambda(\Omega)\cap H^{r}\Lambda(\Omega), there exist v1𝔎HrΛ(Ω)v_{1}\in\mathfrak{K}^{-}\cap H^{r}\Lambda^{-}(\Omega) and v2𝔎HrΛ(Ω)v_{2}\in\mathfrak{K}\cap H^{r}\Lambda(\Omega) such that

v=dv1L2v2,v=\,{\rm d}^{-}v_{1}\oplus^{\bot_{L^{2}}}v_{2},

and

(32) v2lvl.\|v_{2}\|_{l}\lesssim\|v\|_{l}.

Therefore,

Ihv=dPhv1L2Phv2.I_{h}v=\,{\rm d}^{-}P_{h}^{-}v_{1}\oplus^{\bot_{L^{2}}}P_{h}v_{2}.

Note that v2v_{2} is the solution of the problem

(33) v2=δ+s,dv2=qinΩ,trs=0onΩ.v_{2}=\delta^{+}s,\quad\,{\rm d}v_{2}=q\quad\text{in}\quad\Omega,\qquad\operatorname*{tr}s=0\quad\text{on}\quad\partial\Omega.

with q=dvq=\,{\rm d}v. Then Phv2P_{h}v_{2} is the mixed finite element approximation of v2v_{2} in VhV_{h}, the standard error estimates of the mixed finite element method [5, 18] implies that

v2Phv2hlv2lhlvl,\|v_{2}-P_{h}v_{2}\|\lesssim h^{l}\|v_{2}\|_{l}\lesssim h^{l}\|v\|_{l},

where in the second inequality, we have used (32).

Also note that v1v_{1} is the solution of the problem

(34) δdv1=g,δv1=0,trv1=0onΩ,\delta\,{\rm d}^{-}v_{1}=g,\quad\delta^{-}v_{1}=0,\quad\operatorname*{tr}v_{1}=0\quad\text{on}\quad\partial\Omega,

with g=δvg=\delta v. The definition of PhP_{h}^{-} implies that Phv1P_{h}^{-}v_{1} is the mixed finite element approximation of v1v_{1} in VhV_{h}^{-}, then the standard error estimates for the mixed finite element methods [5, 18] imply

d(v1Phv1)hldv1lhlgl1=hlδvl1vl.\|\,{\rm d}^{-}(v_{1}-P_{h}^{-}v_{1})\|\lesssim h^{l}\|\,{\rm d}^{-}v_{1}\|_{l}\lesssim h^{l}\|g\|_{l-1}=h^{l}\|\delta v\|_{l-1}\lesssim\|v\|_{l}.

Therefore,

vIhvd(v1Phv1)+v2Phv2hlvl.\|v-I_{h}v\|\leq\|\,{\rm d}^{-}(v_{1}-P_{h}^{-}v_{1})\|+\|v_{2}-P_{h}v_{2}\|\lesssim h^{l}\|v\|_{l}.

We turn to the estimates of (31). Since for any ϕh𝔎h\phi_{h}\in\mathfrak{K}_{h}, it holds

dPhv2,dϕh\displaystyle\langle\,{\rm d}P_{h}v_{2},\,{\rm d}\phi_{h}\rangle =dv2,dϕh=dπhv2+d(Iπh)v2,dϕh,\displaystyle=\langle\,{\rm d}v_{2},\,{\rm d}\phi_{h}\rangle=\langle\,{\rm d}\pi_{h}v_{2}+\,{\rm d}(I-\pi_{h})v_{2},\,{\rm d}\phi_{h}\rangle,

where πh:H0Λ(Ω)HrΛ(Ω)Vh\pi_{h}:~{}H_{0}\Lambda(\Omega)\cap H^{r}\Lambda(\Omega)\rightarrow V_{h} is the classical interpolation operator [1]. Therefore,

dPhv2=dπhv2+Q0,h+d(Iπh)v2,\,{\rm d}P_{h}v_{2}=\,{\rm d}\pi_{h}v_{2}+Q_{\mathfrak{Z}_{0,h}^{+}}\,{\rm d}(I-\pi_{h})v_{2},

where Q0,h+:L2Λ+0,h+Q_{\mathfrak{Z}_{0,h}^{+}}:~{}L^{2}\Lambda^{+}\rightarrow\mathfrak{Z}_{0,h}^{+} is the L2L^{2} orthogonal projection operator. Then, we have

d(vIhv)=d(v2Phv2)(Iπh+)dv2hldvl.\displaystyle\|\,{\rm d}(v-I_{h}v)\|=\|\,{\rm d}(v_{2}-P_{h}v_{2})\|\leq\|(I-\pi_{h}^{+})\,{\rm d}v_{2}\|\lesssim h^{l}\|\,{\rm d}v\|_{l}.

3.3. Semi-discretization and error analysis

The semi-discrete formulation [26, 3] of (18)-(20) is: Given fL2((0,T),L2Λ)f\in L^{2}((0,T),L^{2}\Lambda), find 𝒖h=(σh,μh,ωh):(0,T]Vh×Vh×Vh+:=𝑾h\boldsymbol{u}_{h}=(\sigma_{h},\mu_{h},\omega_{h})^{\intercal}:~{}(0,T]\mapsto V_{h}^{-}\times V_{h}\times V_{h}^{+}:=\boldsymbol{W}_{h} such that

(35) σh,t,τhdτh,μh\displaystyle\langle\sigma_{h,t},\tau_{h}\rangle-\langle\,{\rm d}^{-}\tau_{h},\mu_{h}\rangle =0τhVh,\displaystyle=0\,\qquad\qquad\forall~{}~{}\tau_{h}\in V_{h}^{-},
(36) μh,t,vh+dσh,vh+ωh,dvh\displaystyle\langle\mu_{h,t},v_{h}\rangle+\langle\,{\rm d}^{-}\sigma_{h},v_{h}\rangle+\langle\omega_{h},\,{\rm d}v_{h}\rangle =f,vhvhVh,\displaystyle=\langle f,v_{h}\rangle\qquad\forall~{}~{}v_{h}\in V_{h},
(37) ωh,t,ϕhdμh,ϕh\displaystyle\langle\omega_{h,t},\phi_{h}\rangle-\langle\,{\rm d}\mu_{h},\phi_{h}\rangle =0ϕhVh+,\displaystyle=0\,\qquad\qquad\forall~{}~{}\phi_{h}\in V_{h}^{+},

with initial values

σh(,0)=Ihδu0,μh(,0)=Ihu1,ωh(,0)=Ih+du0.\sigma_{h}(\cdot,0)=I_{h}^{-}\delta u_{0},\quad\mu_{h}(\cdot,0)=I_{h}u_{1},\quad\omega_{h}(\cdot,0)=I_{h}^{+}\,{\rm d}u_{0}.

Introduce

𝒜h=(0δh0d0δh+0d0)\mathcal{A}_{h}=\begin{pmatrix}0&\delta_{h}&0\\ -\,{\rm d}^{-}&0&-\delta_{h}^{+}\\ 0&\,{\rm d}&0\end{pmatrix}

(35) - (37) can be rewritten as

(38) 𝒖h,t,𝒗h+𝒜h𝒖h,𝒗h=𝑭,𝒗h𝒗h𝑾h.\langle\boldsymbol{u}_{h,t},\boldsymbol{v}_{h}\rangle+\langle\mathcal{A}_{h}\boldsymbol{u}_{h},\boldsymbol{v}_{h}\rangle=\langle\boldsymbol{F},\boldsymbol{v}_{h}\rangle\qquad\forall~{}~{}\boldsymbol{v}_{h}\in\boldsymbol{W}_{h}.

Following the same line as the proof of Theorem 2.2, we have the energy estimates.

Theorem 3.6.

Let 𝐮h=(σh,μh,ωh)𝐖h\boldsymbol{u}_{h}=(\sigma_{h},\mu_{h},\omega_{h})^{\intercal}\in\boldsymbol{W}_{h} be the solution of the mixed formulation (35)-(37) or (38). Provided fL1((0,T),L2Λ)f\in L^{1}((0,T),L^{2}\Lambda), we have the energy bound

(39) sup0sT𝒖h(,t)𝒖h(,0)+20Tf(,s)ds.\sup\limits_{0\leq s\leq T}\|\boldsymbol{u}_{h}(\cdot,t)\|\leq\|\boldsymbol{u}_{h}(\cdot,0)\|+2\int_{0}^{T}\|f(\cdot,s)\|\,{\rm d}s.

Furthermore, if fW1,1((0,T),L2Λ)f\in W^{1,1}((0,T),L^{2}\Lambda), we have the bound

(40) sup0tT𝒜h𝒖h(,t)𝒜h𝒖h(,0)+4fL(L2)+20Tftdt.\sup\limits_{0\leq t\leq T}\|\mathcal{A}_{h}\boldsymbol{u}_{h}(\cdot,t)\|\leq\|\mathcal{A}_{h}\boldsymbol{u}_{h}(\cdot,0)\|+4\|f\|_{L^{\infty}(L^{2})}+2\int_{0}^{T}\|f_{t}\|\,{\rm d}t.

When f=0f=0, these inequalities become equalities and we have the energy conservation

𝒖h(,t)=𝒖h(,0),and𝒜h𝒖h(,t)=𝒜h𝒖h(,0)t>0.\|\boldsymbol{u}_{h}(\cdot,t)\|=\|\boldsymbol{u}_{h}(\cdot,0)\|,and\ \|\mathcal{A}_{h}\boldsymbol{u}_{h}(\cdot,t)\|=\|\mathcal{A}_{h}\boldsymbol{u}_{h}(\cdot,0)\|\quad\forall t>0.
Remark 3.7.

Since (35)-(37) (or (38)) is a linear system, Theorem 3.6 implies the existence and uniqueness of the solution at any time level t(0,T]t\in(0,T]. Also, when the source term f=0f=0, Theorem 3.6 implies that the energies 𝐮h(,t)\|\boldsymbol{u}_{h}(\cdot,t)\| and 𝒜h𝐮h(,t)\|\mathcal{A}_{h}\boldsymbol{u}_{h}(\cdot,t)\| are preserved exactly.

The rest of this section will focus on the error estimates of the semi-discretization (35)-(37) (or its simplified form (38)). We denote

h=(Ih000Ih000Ih+).\mathcal{I}_{h}=\begin{pmatrix}I_{h}^{-}&0&0\\ 0&I_{h}&0\\ 0&0&I_{h}^{+}\end{pmatrix}.

Then for any 𝒗h=(τh,vh,ϕh)𝑾h\boldsymbol{v}_{h}=(\tau_{h},v_{h},\phi_{h})^{\intercal}\in\boldsymbol{W}_{h}, (23) is equivalent to

(41) h𝒖t,𝒗h+𝒜hh𝒖,𝒗h=𝑭,𝒗h+Θh,t,𝒗h+𝒜hh𝒖𝒜𝒖,𝒗h\langle\mathcal{I}_{h}\boldsymbol{u}_{t},\boldsymbol{v}_{h}\rangle+\langle\mathcal{A}_{h}\mathcal{I}_{h}\boldsymbol{u},\boldsymbol{v}_{h}\rangle=\langle\boldsymbol{F},\boldsymbol{v}_{h}\rangle+\langle\Theta_{h,t},\boldsymbol{v}_{h}\rangle+\langle\mathcal{A}_{h}\mathcal{I}_{h}\boldsymbol{u}-\mathcal{A}\boldsymbol{u},\boldsymbol{v}_{h}\rangle

with

Θh=h𝒖𝒖.\Theta_{h}=\mathcal{I}_{h}\boldsymbol{u}-\boldsymbol{u}.

Using the properties of the projection-based quasi-interpolation operator IhI_{h}, we obtain

𝒜hh𝒖𝒜𝒖,𝒗h\displaystyle\langle\mathcal{A}_{h}\mathcal{I}_{h}\boldsymbol{u}-\mathcal{A}\boldsymbol{u},\boldsymbol{v}_{h}\rangle =Ihμμ,dτhd(IhI)σ,vh\displaystyle=\langle I_{h}\mu-\mu,\,{\rm d}^{-}\tau_{h}\rangle-\langle\,{\rm d}^{-}(I_{h}^{-}-I)\sigma,v_{h}\rangle
Ih+ωω,dvh+d(IhI)μ,ϕh\displaystyle\quad-\langle I_{h}^{+}\omega-\omega,\,{\rm d}v_{h}\rangle+\langle\,{\rm d}(I_{h}-I)\mu,\phi_{h}\rangle
=d(IhI)σ,vh+d(IhI)μ,ϕh\displaystyle=-\langle\,{\rm d}^{-}(I_{h}^{-}-I)\sigma,v_{h}\rangle+\langle\,{\rm d}(I_{h}-I)\mu,\phi_{h}\rangle
=𝑮,𝑽h,\displaystyle=\langle\boldsymbol{G},\boldsymbol{V}_{h}\rangle,

with 𝑮=(0,d(πhI)σ,d(IhI)μ)\boldsymbol{G}=(0,-\,{\rm d}^{-}(\pi_{h}^{-}-I)\sigma,\,{\rm d}(I_{h}-I)\mu)^{\intercal}. Denote

h=h𝒖𝒖h\mathcal{E}_{h}=\mathcal{I}_{h}\boldsymbol{u}-\boldsymbol{u}_{h}

and subtracting the semi-discrete form (38) from (41), we get

(42) h,t,𝒗h+𝒜hh,𝒗h=Θh,t+𝑮,𝒗h𝒗h𝑾h.\displaystyle\langle\mathcal{E}_{h,t},\boldsymbol{v}_{h}\rangle+\langle\mathcal{A}_{h}\mathcal{E}_{h},\boldsymbol{v}_{h}\rangle=\langle\Theta_{h,t}+\boldsymbol{G},\boldsymbol{v}_{h}\rangle\qquad\forall~{}~{}\boldsymbol{v}_{h}\in\boldsymbol{W}_{h}.

We have the following estimate of h\mathcal{E}_{h}.

Lemma 3.8.

Suppose the exact solution 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega) of (23) has time derivatives σtL1((0,T),HrΛ)\sigma_{t}\in L^{1}((0,T),H^{r}\Lambda^{-}), μtL1((0,T),HrΛ)\mu_{t}\in L^{1}((0,T),H^{r}\Lambda) and ωtL1((0,T),HrΛ+)\omega_{t}\in L^{1}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Then, for any 1mr1\leq m\leq r and t[0,T]t\in[0,T], we have the bound

h(,t)hm0T(𝒖tm+dσm+dμm)dt.\|\mathcal{E}_{h}(\cdot,t)\|\lesssim h^{m}\int_{0}^{T}\left(\|\boldsymbol{u}_{t}\|_{m}+\|\,{\rm d}^{-}\sigma\|_{m}+\|\,{\rm d}\mu\|_{m}\right)\,{\rm d}t.
Proof.

The fact that h(,0)=0\mathcal{E}_{h}(\cdot,0)=0 and Theorem 3.6 implies

sup0tTh(,t)20TΘh,t+𝑮dt.\sup\limits_{0\leq t\leq T}\|\mathcal{E}_{h}(\cdot,t)\|\leq 2\int_{0}^{T}\|\Theta_{h,t}+\boldsymbol{G}\|\,{\rm d}t.

Using the triangle inequality and the approximation properties of IhI_{h}, the desired result follows. ∎

We then obtain the following estimates by the triangle inequality, Lemma 3.5, and 3.6.

Theorem 3.9.

Suppose the exact solution 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} of (23) has time derivatives σtL1((0,T),HrΛ)\sigma_{t}\in L^{1}((0,T),H^{r}\Lambda^{-}), μtL1((0,T),HrΛ)\mu_{t}\in L^{1}((0,T),H^{r}\Lambda) and ωtL1((0,T),HrΛ+)\omega_{t}\in L^{1}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Let 𝐮h\boldsymbol{u}_{h} be the exact solution of (38). Then, for any 1mr1\leq m\leq r and t[0,T]t\in[0,T], we have the bound

𝒖(,t)𝒖h(,t)\displaystyle\|\boldsymbol{u}(\cdot,t)-\boldsymbol{u}_{h}(\cdot,t)\| hm(𝒖L(Hm)+0T(𝒖tm+dσm+dμm)dt).\displaystyle\lesssim h^{m}\left(\|\boldsymbol{u}\|_{L^{\infty}(H^{m})}+\int_{0}^{T}(\|\boldsymbol{u}_{t}\|_{m}+\|\,{\rm d}^{-}\sigma\|_{m}+\|\,{\rm d}\mu\|_{m})\,{\rm d}t\right).
Remark 3.10.

In this theorem, the convergence order mm is determined by the polynomial order of the finite element spaces preserved.

Remark 3.11.

It should be point out that in [26], Quenneville has obtained an error estimates for the semi-discretization in the form

𝒖𝒖hL(L2)\displaystyle\|\boldsymbol{u}-\boldsymbol{u}_{h}\|_{L^{\infty}(L^{2})} πh𝒖0𝒖0,h+πh𝒖𝒖L(L2)\displaystyle\leq\|\pi_{h}\boldsymbol{u}_{0}-\boldsymbol{u}_{0,h}\|+\|\pi_{h}\boldsymbol{u}-\boldsymbol{u}\|_{L^{\infty}(L^{2})}
+(1+T)((πh𝒖𝒖)(,0)+πh𝒖t𝒖tL1(L2)),\displaystyle\quad+(1+T)(\|(\pi_{h}\boldsymbol{u}-\boldsymbol{u})(\cdot,0)\|+\|\pi_{h}\boldsymbol{u}_{t}-\boldsymbol{u}_{t}\|_{L^{1}(L^{2})}),

where πh\pi_{h} is an elliptic projection operator. Comparing with this result, ours do not have the factor 1+T1+T and thus is more robust to the time variable.

We now give to the error estimates in the energy norm 𝒜𝒖𝒜h𝑼h\|\mathcal{A}\boldsymbol{u}-\mathcal{A}_{h}\boldsymbol{U}_{h}\|, which is equivalent to dσdσh+dμdμh+δμδhμ+δ+ωδh+ωh\|d^{-}\sigma-d^{-}\sigma_{h}\|+\|\,{\rm d}\mu-\,{\rm d}\mu_{h}\|+\|\delta\mu-\delta_{h}\mu\|+\|\delta^{+}\omega-\delta_{h}^{+}\omega_{h}\|. Note that it is possible that the L2L^{2}-norm is small but the energy norm is larger due to the small oscillation in the error. We shall show the energy norm is still of the same order of convergence. We give the estimate of 𝒜hh\|\mathcal{A}_{h}\mathcal{E}_{h}\| first. By Lemma 3.5 and 3.6, we have the following estimate.

Lemma 3.12.

Suppose the exact solution 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} of (23) has time derivatives σttL1((0,T),HrΛ)\sigma_{tt}\in L^{1}((0,T),H^{r}\Lambda^{-}), μttL1((0,T),HrΛ)\mu_{tt}\in L^{1}((0,T),H^{r}\Lambda) and ωttL1((0,T),HrΛ+)\omega_{tt}\in L^{1}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Then, for any 1mr1\leq m\leq r and t[0,T]t\in[0,T], we have the bound

𝒜hh(,t)\displaystyle\|\mathcal{A}_{h}\mathcal{E}_{h}(\cdot,t)\| hm(𝒖tL(Hm)+dσL(Hm)+dμL(Hm))\displaystyle\lesssim h^{m}(\|\boldsymbol{u}_{t}\|_{L^{\infty}(H^{m})}+\|\,{\rm d}^{-}\sigma\|_{L^{\infty}(H^{m})}+\|\,{\rm d}\mu\|_{L^{\infty}(H^{m})})
+hm0T(𝒖ttm+dσtm+dμtm)dt\displaystyle\quad+h^{m}\int_{0}^{T}\left(\|\boldsymbol{u}_{tt}\|_{m}+\|\,{\rm d}^{-}\sigma_{t}\|_{m}+\|\,{\rm d}\mu_{t}\|_{m}\right)\,{\rm d}t
Proof.

Using the fact that h(,0)=0\mathcal{E}_{h}(\cdot,0)=0 and Theorem 3.6, we have

𝒜hh(,t)4Θh,t+𝑮L(L2)+20TΘh,tt+𝑮tdt.\|\mathcal{A}_{h}\mathcal{E}_{h}(\cdot,t)\|\leq 4\|\Theta_{h,t}+\boldsymbol{G}\|_{L^{\infty}(L^{2})}+2\int_{0}^{T}\|\Theta_{h,tt}+\boldsymbol{G}_{t}\|\,{\rm d}t.

Triangle inequality and Lemma 3.5 imply the desired result. ∎

Theorem 3.13.

Suppose the exact solution 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} of (23) has time derivatives σttL1((0,T),HrΛ)\sigma_{tt}\in L^{1}((0,T),H^{r}\Lambda^{-}), μttL1((0,T),HrΛ)\mu_{tt}\in L^{1}((0,T),H^{r}\Lambda) and ωttL1((0,T),HrΛ+)\omega_{tt}\in L^{1}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Then, for any 1mr1\leq m\leq r and t[0,T]t\in[0,T], we have the bound

𝒜h𝒖h(,t)𝒜𝒖(,t)\displaystyle\|\mathcal{A}_{h}\boldsymbol{u}_{h}(\cdot,t)-\mathcal{A}\boldsymbol{u}(\cdot,t)\|\lesssim hm[𝒖tL(Hm)+dσL(Hm)+dμL(Hm)\displaystyle h^{m}{\large[}\,\|\boldsymbol{u}_{t}\|_{L^{\infty}(H^{m})}+\|\,{\rm d}^{-}\sigma\|_{L^{\infty}(H^{m})}+\|\,{\rm d}\mu\|_{L^{\infty}(H^{m})}
𝒜𝒖L(Hm)+0T(𝒖ttm+dσtm+dμt)dt].\displaystyle\|\mathcal{A}\boldsymbol{u}\|_{L^{\infty}(H^{m})}+\int_{0}^{T}(\|\boldsymbol{u}_{tt}\|_{m}+\|\,{\rm d}^{-}\sigma_{t}\|_{m}+\|\,{\rm d}\mu_{t}\|)\,{\rm d}t\,{\large]}.
Proof.

The triangle inequality and Lemma 3.4 imply that

𝒜h𝒖h(,t)𝒜𝒖(,t)\displaystyle\|\mathcal{A}_{h}\boldsymbol{u}_{h}(\cdot,t)-\mathcal{A}\boldsymbol{u}(\cdot,t)\| d(σIhσ)+d(μIhμ)+δμδhIhμ\displaystyle\lesssim\|\,{\rm d}^{-}(\sigma-I_{h}^{-}\sigma)\|+\|\,{\rm d}(\mu-I_{h}\mu)\|+\|\delta\mu-\delta_{h}I_{h}\mu\|
+δ+ωδh+Ih+ω+𝒜hh(,t)\displaystyle\quad+\|\delta^{+}\omega-\delta^{+}_{h}I_{h}^{+}\omega\|+\|\mathcal{A}_{h}\mathcal{E}_{h}(\cdot,t)\|
=d(IIh)σ+d(μIhμ)+(IQh)δμ\displaystyle=\|\,{\rm d}^{-}(I-I_{h}^{-})\sigma\|+\|\,{\rm d}(\mu-I_{h}\mu)\|+\|(I-Q_{h}^{-})\delta\mu\|
+(IQh)δ+ω+𝒜hh(,t).\displaystyle\quad+\|(I-Q_{h})\delta^{+}\omega\|+\|\mathcal{A}_{h}\mathcal{E}_{h}(\cdot,t)\|.

Using the properties of the L2L^{2} projection operators, Lemma 3.5 and 3.12, we get the desired results. ∎

4. Full-discretization

In this section, we will consider the full discretization. We will use a second order continuous time Galerkin method [16] to discretize time variable and will obtain the energy estimates and optimal error estimates.

Energy conservation numerical schemes can have a crucial influence on the quality of the numerical simulations. In long-time simulations, energy-preserving can have a dramatic effect on stability and global error growth. The numerical schemes are not automatically inherit from the semi-discretization and a lot of time discretization methods cannot preserve the energies exactly. These led us to pay more attentions on the time discretization.

4.1. Time discretization

Let 𝒯Δt\mathcal{T}_{\,\Delta t} denote the equispaced partition of the interval (0,T)(0,T) with Δt=T/N\,\Delta t=T/N and NN the number of elements in 𝒯Δt\mathcal{T}_{\,\Delta t}. For 1iN1\leq i\leq N, we denote ti=iΔtt_{i}=i\,\Delta t and τi=(ti1,ti)\tau_{i}=(t_{i-1},t_{i}) with t0=0t_{0}=0. For any quantity v(t)v(t), we denote vi=v(ti)v^{i}=v(t_{i}). Define 𝒫1(𝒯Δt)\mathcal{P}_{1}(\mathcal{T}_{\,\Delta t}) (abbr. 𝒫1\mathcal{P}_{1}) as the set of continuous piecewise linear polynomials with respect to the time variable tt on 𝒯Δt\mathcal{T}_{\,\Delta t} and 𝒫0(𝒯Δt)\mathcal{P}_{0}(\mathcal{T}_{\,\Delta t}) (abbr. 𝒫0\mathcal{P}_{0}) as the set of piecewise constant with respect to the time variable tt on 𝒯Δt\mathcal{T}_{\,\Delta t}. For any Sobolev space SS associates with the spatial variables, we use 𝒫1(S)\mathcal{P}_{1}(S) to denote the set of functions that are continuous piecewise linear polynomials with respect to the time variable tt and in the Sobolev space SS with respect to the spatial variables. 𝒫0(S)\mathcal{P}_{0}(S) is defined similarly.

The full discrete formulation of the Hodge wave equation (23) can be written as: Find 𝑼h=(σ~h,μ~h,ω~h)𝒫1(𝑾h)\boldsymbol{U}_{h}=(\tilde{\sigma}_{h},\tilde{\mu}_{h},\tilde{\omega}_{h})^{\intercal}\in\mathcal{P}_{1}(\boldsymbol{W}_{h}) such that

(43) 0T(𝑼h,t,𝑽h+𝒜h𝑼h,𝑽h)dt=0T𝑭,𝑽hdt𝑽h𝒫0(𝑾h).\int_{0}^{T}\left(\langle\boldsymbol{U}_{h,t},\boldsymbol{V}_{h}\rangle+\langle\mathcal{A}_{h}\boldsymbol{U}_{h},\boldsymbol{V}_{h}\rangle\right)\,{\rm d}t=\int_{0}^{T}\langle\boldsymbol{F},\boldsymbol{V}_{h}\rangle\,{\rm d}t\qquad\forall~{}~{}~{}\boldsymbol{V}_{h}\in\mathcal{P}_{0}(\boldsymbol{W}_{h}).
Remark 4.1.

The full discrete formulation (43) is equivalent to

ti1ti(𝑼h,t,𝑽h+𝒜h𝑼h,𝑽h)dt=ti1ti𝑭,𝑽hdt𝑽h𝒫0(𝑾h), 1iN.\int_{t_{i-1}}^{t_{i}}(\langle\boldsymbol{U}_{h,t},\boldsymbol{V}_{h}\rangle+\langle\mathcal{A}_{h}\boldsymbol{U}_{h},\boldsymbol{V}_{h}\rangle)\,{\rm d}t=\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{F},\boldsymbol{V}_{h}\rangle\,{\rm d}t\quad\forall~{}\boldsymbol{V}_{h}\in\mathcal{P}_{0}(\boldsymbol{W}_{h}),\ 1\leq i\leq N.

The fact that

ti1ti𝑼h,t,𝑽hdt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{U}_{h,t},\boldsymbol{V}_{h}\rangle\,{\rm d}t =𝑼hi𝑼hi1,𝑽h,\displaystyle=\langle\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1},\boldsymbol{V}_{h}\rangle,

and

ti1ti𝒜h𝑼h,𝑽hdt=Δt2𝒜h(𝑼hi+𝑼hi1),𝑽h,\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{U}_{h},\boldsymbol{V}_{h}\rangle\,{\rm d}t=\frac{\,\Delta t}{2}\langle\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}),\boldsymbol{V}_{h}\rangle,

implies the full discrete formulation (43) is essentially a Crank-Nicolson scheme with exact time integration of the right hand side.

We have the following energy estimates for (43).

Theorem 4.2.

Let 𝐔h=(σ~h,μ~h,ω~h)𝒫1(𝐖h)\boldsymbol{U}_{h}=(\tilde{\sigma}_{h},\tilde{\mu}_{h},\tilde{\omega}_{h})^{\intercal}\in\mathcal{P}_{1}(\boldsymbol{W}_{h}) be the solutions of (43). Assume that fL((0,T),L2Λ)f\in L^{\infty}((0,T),L^{2}\Lambda), then there hold the following energy bound

(44) max0iN𝑼hi𝑼h0+20T𝑭dt.\max\limits_{0\leq i\leq N}\|\boldsymbol{U}_{h}^{i}\|\leq\|\boldsymbol{U}_{h}^{0}\|+2\int_{0}^{T}\|\boldsymbol{F}\|\,{\rm d}t.

When F=0F=0, the inequality becomes equality and we have the energy conservation

𝑼hi=𝑼h0, 1iN.\|\boldsymbol{U}_{h}^{i}\|=\|\boldsymbol{U}_{h}^{0}\|,\quad\forall\ 1\leq i\leq N.
Proof.

Taking 𝑽h\boldsymbol{V}_{h} in (43) as

𝑽h|τi=𝑼hi+𝑼hi1and𝑽h|𝒯Δtτi=0,\boldsymbol{V}_{h}|_{\tau_{i}}=\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\qquad\text{and}\qquad\boldsymbol{V}_{h}|_{\mathcal{T}_{\,\Delta t}\setminus\tau_{i}}=0,

we obtain

ti1ti𝑼h,t,𝑼hi+𝑼hi1dt+ti1ti𝒜h𝑼h,𝑼hi+𝑼hi1dt=ti1ti𝑭,𝑼hi+𝑼hi1dt.\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{U}_{h,t},\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\rangle\,{\rm d}t+\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{U}_{h},\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\rangle\,{\rm d}t=\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{F},\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\rangle\,{\rm d}t.

The fact that

ti1ti𝒜h𝑼h,𝑼hi+𝑼hi1dt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{U}_{h},\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\rangle\,{\rm d}t =Δt2𝒜h(𝑼hi+𝑼hi1),𝑼hi+𝑼hi1=0\displaystyle=\frac{\,\Delta t}{2}\langle\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}),\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\rangle=0

and

ti1ti𝑼h,t,𝑼hi+𝑼hi1dt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{U}_{h,t},\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\rangle\,{\rm d}t =𝑼hi2𝑼hi12\displaystyle=\|\boldsymbol{U}_{h}^{i}\|^{2}-\|\boldsymbol{U}_{h}^{i-1}\|^{2}

imply

𝑼hi2𝑼hi12\displaystyle\|\boldsymbol{U}_{h}^{i}\|^{2}-\|\boldsymbol{U}_{h}^{i-1}\|^{2} =ti1ti𝑭,𝑼hi+𝑼hi1dt2max0iN𝑼hiti1ti𝑭dt.\displaystyle=\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{F},\boldsymbol{U}_{h}^{i}+\boldsymbol{U}_{h}^{i-1}\rangle\,{\rm d}t\leq 2\max\limits_{0\leq i\leq N}\|\boldsymbol{U}_{h}^{i}\|\int_{t_{i-1}}^{t_{i}}\|\boldsymbol{F}\|\,{\rm d}t.

Summing over ii from 11 to mNm\leq N, we get

𝑼hm2𝑼h022max0iN𝑼hi0T𝑭dt.\displaystyle\|\boldsymbol{U}_{h}^{m}\|^{2}-\|\boldsymbol{U}_{h}^{0}\|^{2}\leq 2\max\limits_{0\leq i\leq N}\|\boldsymbol{U}_{h}^{i}\|\int_{0}^{T}\|\boldsymbol{F}\|\,{\rm d}t.

Therefore, the desired result follows by Lemma 2.1. ∎

Remark 4.3.

Since (43) is a linear system, therefore Theorem 4.2 implies the existence and uniqueness of the solution for the full discrete form (43).

Theorem 4.4.

Let 𝐔h=(σ~h,μ~h,ω~h)𝒫1(𝐖h)\boldsymbol{U}_{h}=(\tilde{\sigma}_{h},\tilde{\mu}_{h},\tilde{\omega}_{h})^{\intercal}\in\mathcal{P}_{1}(\boldsymbol{W}_{h}) be the solution of (43). Assume that fW1,1((0,T),L2Λ)f\in W^{1,1}((0,T),L^{2}\Lambda), then there holds the following energy bound

(45) max0iN𝒜h𝑼hi𝒜h𝑼h0+4𝑭L(L2)+20T𝑭tdt.\max\limits_{0\leq i\leq N}\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i}\|\leq\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{0}\|+4\|\boldsymbol{F}\|_{L^{\infty}(L^{2})}+2\int_{0}^{T}\|\boldsymbol{F}_{t}\|\,{\rm d}t.

When F=0F=0, the inequality becomes equality and we have the energy conservation

𝒜h𝑼hi=𝒜h𝑼h0 1iN.\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i}\|=\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{0}\|\quad\forall\ 1\leq i\leq N.
Proof.

Taking 𝑽h\boldsymbol{V}_{h} in (43) as

𝑽h|τi=𝒜h(𝑼hi𝑼hi1)and𝑽h|𝒯Δtτi=0,\boldsymbol{V}_{h}|_{\tau_{i}}=\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1})\quad\text{and}\quad\boldsymbol{V}_{h}|_{\mathcal{T}_{\,\Delta t}\setminus\tau_{i}}=0,

we have

ti1ti𝑼h,t,𝒜h(𝑼hi𝑼hi1)dt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{U}_{h,t},\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1})\rangle\,{\rm d}t +ti1ti𝒜h𝑼h,𝒜h(𝑼hi𝑼hi1)dt\displaystyle+\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{U}_{h},\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1})\rangle\,{\rm d}t
=ti1ti𝑭,𝒜h(𝑼hi𝑼hi1)dt.\displaystyle=\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{F},\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1})\rangle\,{\rm d}t.

Using the fact that

ti1ti𝑼h,t,𝒜h(𝑼hi𝑼hi1)dt=𝑼hi𝑼hi1,𝒜h(𝑼hi𝑼hi1)=0\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{U}_{h,t},\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1})\rangle\,{\rm d}t=\langle\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1},\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1})\rangle=0

and

ti1ti𝒜h𝑼h,𝒜h(𝑼hi𝑼hi1)=Δt2(𝒜h𝑼hi2𝒜h𝑼hi12),\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{U}_{h},\mathcal{A}_{h}(\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1})\rangle=\frac{\,\Delta t}{2}(\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i}\|^{2}-\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i-1}\|^{2}),

we get

𝒜h𝑼hi2𝒜h𝑼hi12\displaystyle\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i}\|^{2}-\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i-1}\|^{2} =2ti1ti𝑭,𝒜h𝑼hi𝑼hi1Δtdt=2ti1ti𝑭,𝒜h𝑼h,tdt.\displaystyle=2\int_{t_{i-1}}^{t_{i}}\left\langle\boldsymbol{F},\mathcal{A}_{h}\frac{\boldsymbol{U}_{h}^{i}-\boldsymbol{U}_{h}^{i-1}}{\,\Delta t}\right\rangle\,{\rm d}t=2\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{F},\mathcal{A}_{h}\boldsymbol{U}_{h,t}\rangle\,{\rm d}t.

Summing over ii from 11 to mNm\leq N, we obtain

𝒜h𝑼hm2\displaystyle\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{m}\|^{2} =𝒜h𝑼h02+20tm𝑭,𝒜h𝑼h,tdt\displaystyle=\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{0}\|^{2}+2\int_{0}^{t_{m}}\langle\boldsymbol{F},\mathcal{A}_{h}\boldsymbol{U}_{h,t}\rangle\,{\rm d}t
=𝒜h𝑼h02+2𝑭m,𝒜h𝑼hm2𝑭0,𝒜h𝑼h020tm𝑭t,𝒜h𝑼hdt\displaystyle=\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{0}\|^{2}+2\langle\boldsymbol{F}^{m},\mathcal{A}_{h}\boldsymbol{U}_{h}^{m}\rangle-2\langle\boldsymbol{F}^{0},\mathcal{A}_{h}\boldsymbol{U}_{h}^{0}\rangle-2\int_{0}^{t_{m}}\langle\boldsymbol{F}_{t},\mathcal{A}_{h}\boldsymbol{U}_{h}\rangle\,{\rm d}t
𝒜h𝑼h02+2max0iN𝒜h𝑼hi(2𝑭L(L2)+0T𝑭tdt).\displaystyle\leq\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{0}\|^{2}+2\max\limits_{0\leq i\leq N}\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i}\|\left(2\|\boldsymbol{F}\|_{L^{\infty}(L^{2})}+\int_{0}^{T}\|\boldsymbol{F}_{t}\|\,{\rm d}t\right).

Then the desired result follows by a direct using of Lemma 2.1. ∎

4.2. Error analysis of the full discretization

In this subsection, we turn to the error estimates of the full discrete formulation (43). We bound the error of the full discrete formulation in various norms. Let

𝒆h=h𝒖𝑼h.\displaystyle\boldsymbol{e}_{h}=\mathcal{I}_{h}\boldsymbol{u}-\boldsymbol{U}_{h}.

Simple caculation shows that for any 𝑽h𝒫0(𝑾h)\boldsymbol{V}_{h}\in\mathcal{P}_{0}(\boldsymbol{W}_{h}), 𝒆h\boldsymbol{e}_{h} satisfies the following equation

(46) 0T(𝒆h,t,𝑽h+𝒜h𝒆h,𝑽h)dt=0TΘh+𝑮,𝑽hdt,𝑽h𝒫0(𝑾h).\int_{0}^{T}(\langle\boldsymbol{e}_{h,t},\boldsymbol{V}_{h}\rangle+\langle\mathcal{A}_{h}\boldsymbol{e}_{h},\boldsymbol{V}_{h}\rangle)\,{\rm d}t=\int_{0}^{T}\langle\Theta_{h}+\boldsymbol{G},\boldsymbol{V}_{h}\rangle\,{\rm d}t,\quad\forall~{}~{}\boldsymbol{V}_{h}\in\mathcal{P}_{0}(\boldsymbol{W}_{h}).

Then, we have the following estimates of 𝒆h\boldsymbol{e}_{h}.

Lemma 4.5.

Let 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} be the solution of (23) and 𝐔h=(σ~h,μ~h,ω~h)𝒫1(𝐖h)\boldsymbol{U}_{h}=(\tilde{\sigma}_{h},\tilde{\mu}_{h},\tilde{\omega}_{h})^{\intercal}\in\mathcal{P}_{1}(\boldsymbol{W}_{h}) be the solution of (43). Assume that σttL1((0,T),HrΛ)\sigma_{tt}\in L^{1}((0,T),H^{r}\Lambda^{-}), μttL1((0,T),HrΛ)\mu_{tt}\in L^{1}((0,T),H^{r}\Lambda) and ωttL1((0,T),HrΛ+)\omega_{tt}\in L^{1}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Then, for any 1mr1\leq m\leq r, we have the bound

max0iN𝒆hi\displaystyle\max\limits_{0\leq i\leq N}\|\boldsymbol{e}_{h}^{i}\|\lesssim hm0T(𝒖hm+dσm+dμm)dt+Δt20T𝒜𝒖ttdt.\displaystyle h^{m}\int_{0}^{T}(\|\boldsymbol{u}_{h}\|_{m}+\|\,{\rm d}^{-}\sigma\|_{m}+\|\,{\rm d}\mu\|_{m})\,{\rm d}t+\,\Delta t^{2}\int_{0}^{T}\|\mathcal{A}\boldsymbol{u}_{tt}\|\,{\rm d}t.
Proof.

Taking 𝑽h\boldsymbol{V}_{h} in (46) as

𝑽h|τi=𝒆hi1+𝒆hiand𝑽h|𝒯Δtτi=0,\boldsymbol{V}_{h}|_{\tau_{i}}=\boldsymbol{e}_{h}^{i-1}+\boldsymbol{e}_{h}^{i}\quad\text{and}\quad\boldsymbol{V}_{h}|_{\mathcal{T}_{\,\Delta t}\setminus\tau_{i}}=0,

we obtain

ti1ti𝒆h,t,𝒆hi1+𝒆hidt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{e}_{h,t},\boldsymbol{e}_{h}^{i-1}+\boldsymbol{e}_{h}^{i}\rangle\,{\rm d}t +ti1ti𝒜h𝒆h,𝒆hi+𝒆hi+1dt=ti1tiΘh+𝑮,𝒆hi1+𝒆hidt.\displaystyle+\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{e}_{h},\boldsymbol{e}_{h}^{i}+\boldsymbol{e}_{h}^{i+1}\rangle\,{\rm d}t=\int_{t_{i-1}}^{t_{i}}\langle\Theta_{h}+\boldsymbol{G},\boldsymbol{e}_{h}^{i-1}+\boldsymbol{e}_{h}^{i}\rangle\,{\rm d}t.

Note that

ti1ti𝒆h,t,𝒆hi1+𝒆hidt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{e}_{h,t},\boldsymbol{e}_{h}^{i-1}+\boldsymbol{e}_{h}^{i}\rangle\,{\rm d}t =𝒆hi2𝒆hi12,\displaystyle=\|\boldsymbol{e}_{h}^{i}\|^{2}-\|\boldsymbol{e}_{h}^{i-1}\|^{2},

and

ti1ti𝒜h𝒆h,𝒆hi1+𝒆hidt=ti1ti𝒜h(IJΔt)𝒆h,𝒆hi1+𝒆hidt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{e}_{h},\boldsymbol{e}_{h}^{i-1}+\boldsymbol{e}_{h}^{i}\rangle\,{\rm d}t=\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}(I-J_{\,\Delta t})\boldsymbol{e}_{h},\boldsymbol{e}_{h}^{i-1}+\boldsymbol{e}_{h}^{i}\rangle\,{\rm d}t
\displaystyle\leq 2max0iN𝒆hiti1ti𝒜h(IJΔt)𝒆hdtCΔt2max0ileqN𝒆hiti1ti𝒜hh𝒖ttdt\displaystyle 2\max\limits_{0\leq i\leq N}\|\boldsymbol{e}_{h}^{i}\|\int_{t_{i-1}}^{t_{i}}\|\mathcal{A}_{h}(I-J_{\,\Delta t})\boldsymbol{e}_{h}\|\,{\rm d}t\leq C\,\Delta t^{2}\max\limits_{0\leq ileqN}\|\boldsymbol{e}_{h}^{i}\|\int_{t_{i-1}}^{t_{i}}\|\mathcal{A}_{h}\mathcal{I}_{h}\boldsymbol{u}_{tt}\|\,{\rm d}t
\displaystyle\leq CΔt2max0ileqN𝒆hiti1ti𝒜𝒖ttdt,\displaystyle C\,\Delta t^{2}\max\limits_{0\leq ileqN}\|\boldsymbol{e}_{h}^{i}\|\int_{t_{i-1}}^{t_{i}}\|\mathcal{A}\boldsymbol{u}_{tt}\|\,{\rm d}t,

we then have

𝒆hi2𝒆hi12Cmax0iN𝒆hi(ti1ti(Θh+𝑮)dt+Δt2ti1ti𝒜𝒖ttdt)\displaystyle\|\boldsymbol{e}_{h}^{i}\|^{2}-\|\boldsymbol{e}_{h}^{i-1}\|^{2}\leq C\max\limits_{0\leq i\leq N}\|\boldsymbol{e}_{h}^{i}\|\left(\int_{t_{i-1}}^{t_{i}}(\|\Theta_{h}\|+\|\boldsymbol{G}\|)\,{\rm d}t+\,\Delta t^{2}\int_{t_{i-1}}^{t_{i}}\|\mathcal{A}\boldsymbol{u}_{tt}\|\,{\rm d}t\right)
\displaystyle\lesssim max0iN𝒆hi(hmti1ti(𝒖hm+dσm+dμm)dt+Δt2ti1ti𝒜𝒖ttdt).\displaystyle\max\limits_{0\leq i\leq N}\|\boldsymbol{e}_{h}^{i}\|\left(h^{m}\int_{t_{i-1}}^{t_{i}}(\|\boldsymbol{u}_{h}\|_{m}+\|\,{\rm d}^{-}\sigma\|_{m}+\|\,{\rm d}\mu\|_{m})\,{\rm d}t+\,\Delta t^{2}\int_{t_{i-1}}^{t_{i}}\|\mathcal{A}\boldsymbol{u}_{tt}\|\,{\rm d}t\right).

Summing over ii from 11 to mNm\leq N, we get

𝒆hm2\displaystyle\|\boldsymbol{e}_{h}^{m}\|^{2} max0iN𝒆hi(hm0T(𝒖m+dσm+dμm)dt+Δt20T𝒜𝒖ttdt).\displaystyle\lesssim\max\limits_{0\leq i\leq N}\|\boldsymbol{e}_{h}^{i}\|\left(h^{m}\int_{0}^{T}(\|\boldsymbol{u}\|_{m}+\|\,{\rm d}^{-}\sigma\|_{m}+\|\,{\rm d}\mu\|_{m})\,{\rm d}t+\,\Delta t^{2}\int_{0}^{T}\|\mathcal{A}\boldsymbol{u}_{tt}\|\,{\rm d}t\right).

Then the desired result follows. ∎

Using triangle inequality, Lemma 4.5 and the properties of IhI_{h}, we have the following L2L^{2} error estimate.

Theorem 4.6.

Let 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} be the solution of (23) and 𝐔h=(σ~h,μ~h,ω~h)𝒫1(𝐖h)\boldsymbol{U}_{h}=(\tilde{\sigma}_{h},\tilde{\mu}_{h},\tilde{\omega}_{h})^{\intercal}\in\mathcal{P}_{1}(\boldsymbol{W}_{h}) be the solution of (43). Assume that σttL1((0,T),HrΛ)\sigma_{tt}\in L^{1}((0,T),H^{r}\Lambda^{-}), μttL1((0,T),HrΛ)\mu_{tt}\in L^{1}((0,T),H^{r}\Lambda) and ωttL1((0,T),HrΛ+)\omega_{tt}\in L^{1}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Then, for any 1mr1\leq m\leq r we have the bound

max0iN𝒖i𝑼hi\displaystyle\max\limits_{0\leq i\leq N}\|\boldsymbol{u}^{i}-\boldsymbol{U}_{h}^{i}\|\lesssim hm(𝒖L(Hm)+𝒖L1(Hm)+dσL1(Hm)+dμL1(Hm))\displaystyle h^{m}\left(\|\boldsymbol{u}\|_{L^{\infty}(H^{m})}+\|\boldsymbol{u}\|_{L^{1}(H^{m})}+\|\,{\rm d}^{-}\sigma\|_{L^{1}(H^{m})}+\|\,{\rm d}\mu\|_{L^{1}(H^{m})}\right)
+Δt2𝒜𝒖ttL1(L2)\displaystyle+\,\Delta t^{2}\|\mathcal{A}\boldsymbol{u}_{tt}\|_{L^{1}(L^{2})}

Now, we turn to the estimates of the energy norm 𝒜()\|\mathcal{A}(\cdot)\| error. We first estimate 𝒜h𝒆h\|\mathcal{A}_{h}\boldsymbol{e}_{h}\|.

Lemma 4.7.

Let 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} be the solution of (23) and 𝐔h=(σ~h,μ~h,ω~h)𝒫1(𝐖h)\boldsymbol{U}_{h}=(\tilde{\sigma}_{h},\tilde{\mu}_{h},\tilde{\omega}_{h})^{\intercal}\in\mathcal{P}_{1}(\boldsymbol{W}_{h}) be the solution of (43). Assume that σttL((0,T),HrΛ)\sigma_{tt}\in L^{\infty}((0,T),H^{r}\Lambda^{-}), μttL((0,T),HrΛ)\mu_{tt}\in L^{\infty}((0,T),H^{r}\Lambda) and ωttL((0,T),HrΛ+)\omega_{tt}\in L^{\infty}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Then, for any 1mr1\leq m\leq r we have the bound

𝒜h𝒆h\displaystyle\|\mathcal{A}_{h}\boldsymbol{e}_{h}\| hm(𝒖tL1(Hm)+dσtL1(Hm)+dμtL1(Hm))+Δt2𝒜𝒖ttL(L2).\displaystyle\lesssim h^{m}\left(\|\boldsymbol{u}_{t}\|_{L^{1}(H^{m})}+\|\,{\rm d}^{-}\sigma_{t}\|_{L^{1}(H^{m})}+\|\,{\rm d}\mu_{t}\|_{L^{1}(H^{m})}\right)+\,\Delta t^{2}\|\mathcal{A}\boldsymbol{u}_{tt}\|_{L^{\infty}(L^{2})}.
Proof.

Taking 𝑽h\boldsymbol{V}_{h} in (46) as

𝑽h|τi=𝒜h(𝒆hi𝒆hi1)and𝑽h|𝒯Δtτi=0,\boldsymbol{V}_{h}|_{\tau_{i}}=\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\quad\text{and}\quad\boldsymbol{V}_{h}|_{\mathcal{T}_{\,\Delta t}\setminus\tau_{i}}=0,

we obtain

ti1ti𝒆h,t,𝒜h(𝒆hi𝒆hi1)dt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{e}_{h,t},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t +ti1ti𝒜h𝒆h,𝒜h(𝒆hi𝒆hi1)dt\displaystyle+\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{e}_{h},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t
=ti1tiΘh+𝑮,𝒜h(𝒆hi𝒆hi1)dt.\displaystyle=\int_{t_{i-1}}^{t_{i}}\langle\Theta_{h}+\boldsymbol{G},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t.

Note that

ti1ti𝒆h,t,𝒜h(𝒆hi𝒆hi1)dt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\boldsymbol{e}_{h,t},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t =(𝒆hi𝒆hi1),𝒜h(𝒆hi𝒆hi1)=0,\displaystyle=\langle(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1}),\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle=0,

and

ti1ti𝒜h𝒆h,𝒜h(𝒆hi𝒆hi1)dt\displaystyle\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}\boldsymbol{e}_{h},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t =Δt2(𝒜h𝒆hi2𝒜h𝒆hi12)\displaystyle=\frac{\,\Delta t}{2}\left(\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{i}\|^{2}-\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{i-1}\|^{2}\right)
+ti1ti𝒜h(IJΔt)𝒆h,𝒜h(𝒆hi𝒆hi1)dt.\displaystyle\quad+\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}(I-J_{\,\Delta t})\boldsymbol{e}_{h},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t.

Therefore,

𝒜h𝒆hi2𝒜h𝒆hi12\displaystyle\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{i}\|^{2}-\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{i-1}\|^{2}
=\displaystyle= 2Δtti1tiΘh+𝑮,𝒜h(𝒆hi𝒆hi1)dt2Δtti1ti𝒜h(IJΔt)𝒆h,𝒜h(𝒆hi𝒆hi1)dt\displaystyle\frac{2}{\,\Delta t}\int_{t_{i-1}}^{t_{i}}\langle\Theta_{h}+\boldsymbol{G},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t-\frac{2}{\,\Delta t}\int_{t_{i-1}}^{t_{i}}\langle\mathcal{A}_{h}(I-J_{\,\Delta t})\boldsymbol{e}_{h},\mathcal{A}_{h}(\boldsymbol{e}_{h}^{i}-\boldsymbol{e}_{h}^{i-1})\rangle\,{\rm d}t
=\displaystyle= 2ti1tiΘh+𝑮,𝒜ht(JΔt𝒆h)dt2ti1ti𝒜h(IJΔt)𝒆h,𝒜ht(JΔt𝒆h)dt.\displaystyle 2\int_{t_{i-1}}^{t_{i}}\left\langle\Theta_{h}+\boldsymbol{G},\mathcal{A}_{h}\frac{\partial}{\partial t}(J_{\,\Delta t}\boldsymbol{e}_{h})\right\rangle\,{\rm d}t-2\int_{t_{i-1}}^{t_{i}}\left\langle\mathcal{A}_{h}(I-J_{\,\Delta t})\boldsymbol{e}_{h},\mathcal{A}_{h}\frac{\partial}{\partial t}(J_{\,\Delta t}\boldsymbol{e}_{h})\right\rangle\,{\rm d}t.

Summing over ii from 11 to mNm\leq N, we get

𝒜h𝒆hm2\displaystyle\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{m}\|^{2}
=\displaystyle= 20tmΘh+𝑮,𝒜ht(JΔt𝒆h)dt20tm𝒜h(IJΔt)𝒆h,𝒜ht(JΔt𝒆h)dt\displaystyle 2\int_{0}^{t_{m}}\left\langle\Theta_{h}+\boldsymbol{G},\mathcal{A}_{h}\frac{\partial}{\partial t}(J_{\,\Delta t}\boldsymbol{e}_{h})\right\rangle\,{\rm d}t-2\int_{0}^{t_{m}}\left\langle\mathcal{A}_{h}(I-J_{\,\Delta t})\boldsymbol{e}_{h},\mathcal{A}_{h}\frac{\partial}{\partial t}(J_{\,\Delta t}\boldsymbol{e}_{h})\right\rangle\,{\rm d}t
=\displaystyle= 2Θhm+𝑮m,𝒜h𝒆hm20tmΘh,t+𝑮t,𝒜hJΔt𝒆hdt\displaystyle 2\langle\Theta_{h}^{m}+\boldsymbol{G}^{m},\mathcal{A}_{h}\boldsymbol{e}_{h}^{m}\rangle-2\int_{0}^{t_{m}}\langle\Theta_{h,t}+\boldsymbol{G}_{t},\mathcal{A}_{h}J_{\,\Delta t}\boldsymbol{e}_{h}\rangle\,{\rm d}t
+20tm𝒜ht(IJΔt)𝒆h,𝒜h(JΔt𝒆h)dt\displaystyle+2\int_{0}^{t_{m}}\left\langle\mathcal{A}_{h}\frac{\partial}{\partial t}(I-J_{\,\Delta t})\boldsymbol{e}_{h},\mathcal{A}_{h}(J_{\,\Delta t}\boldsymbol{e}_{h})\right\rangle\,{\rm d}t
\displaystyle\leq 2max0iN𝒜h𝒆hi(max0iNΘhi+𝑮i+0TΘh,t+𝑮tdt+0T𝒜ht(IJΔt)𝒆hdt)\displaystyle 2\max\limits_{0\leq i\leq N}\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{i}\|\left(\max\limits_{0\leq i\leq N}\|\Theta_{h}^{i}+\boldsymbol{G}^{i}\|+\int_{0}^{T}\|\Theta_{h,t}+\boldsymbol{G}_{t}\|\,{\rm d}t+\int_{0}^{T}\left\|\mathcal{A}_{h}\frac{\partial}{\partial t}(I-J_{\,\Delta t})\boldsymbol{e}_{h}\right\|\,{\rm d}t\right)
\displaystyle\leq Cmax0iN𝒜h𝒆hi0T(Θh,t+𝑮t+𝒜ht(IJΔt)𝒆h)dt\displaystyle C\max\limits_{0\leq i\leq N}\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{i}\|\int_{0}^{T}\left(\|\Theta_{h,t}\|+\|\boldsymbol{G}_{t}\|+\left\|\mathcal{A}_{h}\frac{\partial}{\partial t}(I-J_{\,\Delta t})\boldsymbol{e}_{h}\right\|\right)\,{\rm d}t
\displaystyle\leq Cmax0iN𝒜h𝒆hi[hm(𝒖tL1(Hm)+dσtL1(Hm)+dμtL1(Hm)+Δt2𝒜𝒖ttL(L2))],\displaystyle C\max\limits_{0\leq i\leq N}\|\mathcal{A}_{h}\boldsymbol{e}_{h}^{i}\|\left[h^{m}\left(\|\boldsymbol{u}_{t}\|_{L^{1}(H^{m})}+\|\,{\rm d}^{-}\sigma_{t}\|_{L^{1}(H^{m})}+\|\,{\rm d}\mu_{t}\|_{L^{1}(H^{m})}+\,\Delta t^{2}\|\|\mathcal{A}\boldsymbol{u}_{tt}\|_{L^{\infty}(L^{2})}\right)\right],

where in the last inequality, we have used the properties of the one dimensional interpolation operator. Then the desired result follows. ∎

We summarize the error estimate for the full discretization below.

Theorem 4.8.

Let 𝐮=(σ,μ,ω)\boldsymbol{u}=(\sigma,\mu,\omega)^{\intercal} be the solution of (23) and 𝐔h=(σ~h,μ~h,ω~h)𝒫1(𝐖h)\boldsymbol{U}_{h}=(\tilde{\sigma}_{h},\tilde{\mu}_{h},\tilde{\omega}_{h})^{\intercal}\in\mathcal{P}_{1}(\boldsymbol{W}_{h}) be the solution of (43). Assume that σtttL((0,T),HrΛ)\sigma_{ttt}\in L^{\infty}((0,T),H^{r}\Lambda^{-}), μtttL((0,T),HrΛ)\mu_{ttt}\in L^{\infty}((0,T),H^{r}\Lambda) and ωtttL((0,T),HrΛ+)\omega_{ttt}\in L^{\infty}((0,T),H^{r}\Lambda^{+}) with r1r\geq 1. Then, for any 1mr1\leq m\leq r we have the bound

max0iN𝒜𝒖i𝒜h𝑼hi\displaystyle\max\limits_{0\leq i\leq N}\|\mathcal{A}\boldsymbol{u}^{i}-\mathcal{A}_{h}\boldsymbol{U}_{h}^{i}\|\lesssim hm[𝒖tL1(Hm)+𝒜𝒖L(Hm)\displaystyle\,h^{m}\left[\|\boldsymbol{u}_{t}\|_{L^{1}(H^{m})}+\|\mathcal{A}\boldsymbol{u}\|_{L^{\infty}(H^{m})}\right.
+dσtL1(Hm)+dμtL1(Hm)]\displaystyle+\left.\|\,{\rm d}^{-}\sigma_{t}\|_{L^{1}(H^{m})}+\|\,{\rm d}\mu_{t}\|_{L^{1}(H^{m})}\right]
+\displaystyle+ Δt2𝒜𝒖ttL(L2).\displaystyle\,\Delta t^{2}\|\mathcal{A}\boldsymbol{u}_{tt}\|_{L^{\infty}(L^{2})}.

5. Numerical experiments

In this section, we will give some simple numerical examples to illustrate the theoretical results. We consider the Hodge wave equation on the unit square Ω=(0,1)2\Omega=(0,1)^{2}, i.e., n=2n=2 and compute the cases k=0k=0, k=1k=1 and k=2k=2.

5.1. The case k=0k=0

The Hodge wave equation presents in the standard H1(Ω)H^{1}(\Omega) or H(curl)H({\rm curl\,}) language reads as (note that σ=δμ=0\sigma=\delta\mu=0): Find μH01\mu\in H_{0}^{1} and 𝝎H0(rot)\boldsymbol{\omega}\in H_{0}(\operatorname*{rot}) or μH0(curl)\mu\in H_{0}({\rm curl\,}) and 𝝎H0(div)\boldsymbol{\omega}\in H_{0}(\operatorname{div}) such that

(47) (μt,v)+(𝝎,gradv)\displaystyle(\mu_{t},v)+(\boldsymbol{\omega},{\rm grad\,}v) =(f,v)𝒗H01(Ω),\displaystyle=(f,v)\qquad\forall~{}~{}\boldsymbol{v}\in H_{0}^{1}(\Omega),
(48) (𝝎t,ϕ)(gradμ,ϕ)\displaystyle(\boldsymbol{\omega}_{t},\boldsymbol{\phi})-({\rm grad\,}\mu,\boldsymbol{\phi}) =0ϕH0(rot).\displaystyle=0\qquad\forall~{}~{}\boldsymbol{\phi}\in H_{0}(\operatorname*{rot}).

or

(49) (μt,v)+(𝝎,curlv)\displaystyle(\mu_{t},v)+(\boldsymbol{\omega},{\rm curl\,}v) =(f,v)𝒗H0(curl),\displaystyle=(f,v)\qquad\forall~{}~{}\boldsymbol{v}\in H_{0}({\rm curl\,}),
(50) (𝝎t,ϕ)(curlμ,ϕ)\displaystyle(\boldsymbol{\omega}_{t},\boldsymbol{\phi})-({\rm curl\,}\mu,\boldsymbol{\phi}) =0ϕH0(div).\displaystyle=0\qquad\forall~{}~{}\boldsymbol{\phi}\in H_{0}(\operatorname{div}).

We only give the numerical results for (49)-(50). We choose the exact solutions as

μ(x,y,t)\displaystyle\mu(x,y,t) =etsin(πx)sin(πy),\displaystyle=e^{-t}\sin(\pi x)\sin(\pi y),
𝝎(x,y,t)\displaystyle\boldsymbol{\omega}(x,y,t) =πet(sin(πx)cos(πy)cos(πx)sin(πy)).\displaystyle=-\pi e^{-t}\begin{pmatrix}\sin(\pi x)\cos(\pi y)\\ -\cos(\pi x)\sin(\pi y)\end{pmatrix}.

The initial conditions are μ0=μ(x,y,0)\mu_{0}=\mu(x,y,0) and 𝝎0=𝝎(x,y,0)\boldsymbol{\omega}_{0}=\boldsymbol{\omega}(x,y,0). We use piecewise continuous second order polynomial to discrete μ\mu and use RT1RT_{1} element [27] to discrete ω\omega, the numerical results are listed in Table 1.

Table 1. Errors and convergence orders in various norms with Δt=0.0001\,\Delta t=0.0001 at t=0.0004t=0.0004.
hh μμh\|\mu-\mu_{h}\| curl(μμh)\|{\rm curl\,}(\mu-\mu_{h})\| ωωh\|\omega-\omega_{h}\|
1/4 3.8253e-03 1.3417e-01 1.3164e-01
1/8 5.0314e-04 3.5025e-02 3.3567e-02
1/16 6.7850e-05 9.5850e-03 8.4467e-03
order 2.914 1.904 1.981

From Table 1, we can see that the mixed finite element method is of second order convergence rate for the variables curlμ{\rm curl\,}\mu and ω\omega, and is of third order convergence rate for μ\mu. All these variables have optimal convergence order.

5.2. The case k=1k=1

The Hodge wave equation is: Find σH0(curl)\sigma\in H_{0}({\rm curl\,}), 𝝁H0(div)\boldsymbol{\mu}\in H_{0}(\operatorname{div}) and ωL02(Ω)\omega\in L_{0}^{2}(\Omega) such that

(σt,τ)(𝝁,curlτ)\displaystyle(\sigma_{t},\tau)-(\boldsymbol{\mu},{\rm curl\,}\tau) =0τH0(curl),\displaystyle=0\qquad\forall~{}~{}\tau\in H_{0}({\rm curl\,}),
(𝝁t,𝒗)+(curlσ,𝒗)+(ω,div𝒗)\displaystyle(\boldsymbol{\mu}_{t},\boldsymbol{v})+({\rm curl\,}\sigma,\boldsymbol{v})+(\omega,\operatorname{div}\boldsymbol{v}) =(𝒇,𝒗)𝒗H0(div),\displaystyle=(\boldsymbol{f},\boldsymbol{v})\qquad\forall~{}~{}\boldsymbol{v}\in H_{0}(\operatorname{div}),
(ωt,ϕ)(div𝝁,ϕ)\displaystyle(\omega_{t},\phi)-(\operatorname{div}\boldsymbol{\mu},\phi) =0ϕL02(Ω).\displaystyle=0\qquad\forall~{}~{}\phi\in L_{0}^{2}(\Omega).

This formulation can be viewed as the mixed method for the time-harmonic Maxwell’s equations with divergence free constrain on both 𝝁\boldsymbol{\mu} and 𝒇\boldsymbol{f}. The formulation can also be viewed as the mixed method for the elastic wave equation. We use continuous piecewise quadratic polynomial to discrete σ\sigma, use RT1RT_{1} element [27] to discrete μ\mu and use discontinuous piecewise linear polynomial to discrete ω\omega. Firstly, we choose the exact solutions as

(51) σ(x,y,t)\displaystyle\sigma(x,y,t) =2et(πsin2(πx)sin(πy)cos(πy)x(x1)(2x1)y2(y1)2)\displaystyle=2e^{-t}\left(\pi\sin^{2}(\pi x)\sin(\pi y)\cos(\pi y)-x(x-1)(2x-1)y^{2}(y-1)^{2}\right)
(52) 𝝁(x,y,t)\displaystyle\boldsymbol{\mu}(x,y,t) =et(sin2(πx)sin2(πy)x2(x1)2y2(y1)2),\displaystyle=e^{-t}\begin{pmatrix}\sin^{2}(\pi x)\sin^{2}(\pi y)\\ x^{2}(x-1)^{2}y^{2}(y-1)^{2}\end{pmatrix},
(53) ω(x,y,t)=2et(πsin(πx)cos(πx)sin2(πy)+x2(x1)2y(y1)(2y1)),\displaystyle\begin{split}\omega(x,y,t)&=-2e^{-t}\left(\pi\sin(\pi x)\cos(\pi x)\sin^{2}(\pi y)\right.\\ &\quad\left.+x^{2}(x-1)^{2}y(y-1)(2y-1)\right),\end{split}

with initial conditions σ0=σ(x,y,0)\sigma_{0}=\sigma(x,y,0), 𝝁0=𝝁(x,y,0)\boldsymbol{\mu}_{0}=\boldsymbol{\mu}(x,y,0) and ω0=ω(x,y,0)\omega_{0}=\omega(x,y,0). The numerical results are listed in Table 2. We also test the long time robustness of our algorithm, the numerical results are listed in Table 3.

Table 2. Errors and convergence orders in various norms with Δt=0.0001\,\Delta t=0.0001 at t=0.0004t=0.0004.
hh σσh\|\sigma-\sigma_{h}\| curl(σσh)\|{\rm curl\,}(\sigma-\sigma_{h})\| μμh\|\mu-\mu_{h}\| div(μμh)\|\operatorname{div}(\mu-\mu_{h})\| ωωh\|\omega-\omega_{h}\|
1/4 4.8563e-02 1.6691e+00 2.9085e-02 0.1391 0.1388
1/8 6.4968e-03 4.4475e-01 7.6216e-03 0.0364 0.0367
1/16 8.2584e-04 1.1298e-01 1.9354e-03 0.0093 0.0093
order 2.949 1.942 1.950 1.950 1.9763
Table 3. Long time problem with Δt=0.1\,\Delta t=0.1 and h=1/16h=1/16.
TT σσh\|\sigma-\sigma_{h}\| curl(σσh)\|{\rm curl\,}(\sigma-\sigma_{h})\| μμh\|\mu-\mu_{h}\| div(μμh)\|\operatorname{div}(\mu-\mu_{h})\| ωωh\|\omega-\omega_{h}\|
10 5.4138e-04 1.5087e-02 3.7500e-01 1.3603 0.2374
30 4.8684e-04 1.8186e-02 3.7502e-01 1.3604 0.2486
50 6.1267e-04 1.4339e-02 3.7502e-01 1.3604 0.4158

Then, we chose 𝒇=0\boldsymbol{f}=0 and the initial conditions σ0=σ(x,y,0)\sigma_{0}=\sigma(x,y,0), 𝝁0=𝝁(x,y,0)\boldsymbol{\mu}_{0}=\boldsymbol{\mu}(x,y,0) and ω0=ω(x,y,0)\omega_{0}=\omega(x,y,0) with σ\sigma, 𝝁\boldsymbol{\mu} and ω\omega defined as in (51) - (53). We compute the energies 𝑼hi\|\boldsymbol{U}_{h}^{i}\| and 𝒜h𝑼hi\|\mathcal{A}_{h}\boldsymbol{U}_{h}^{i}\| on different time levels, the numerical results are showing in Fig. 1.

Refer to caption
Figure 1. Energies 𝑼h\|\boldsymbol{U}_{h}\| and 𝒜h𝑼h\|\mathcal{A}_{h}\boldsymbol{U}_{h}\| in different times with h=1/16h=1/16 and Δt=0.25\,\Delta t=0.25.

From this example, we have the following observations.

  1. (1)

    The mix finite element method is of second order convergence rate for the variables curlσ{\rm curl\,}\sigma, μ\mu, divμ\operatorname{div}\mu and ω\omega, and is of third order convergence rate for the variable σ\sigma. All of these variables have optimal convergence order.

  2. (2)

    From Table 3, we can see that the mixed finite element method is robust for long-time problem.

  3. (3)

    From Fig. 1, we can see that the mixed finite element method conserves the energies 𝑼h\|\boldsymbol{U}_{h}\| and 𝒜h𝑼h\|\mathcal{A}_{h}\boldsymbol{U}_{h}\| exactly.

5.3. The case k=2k=2

The Hodge wave equation presents in the H(div)H(\operatorname{div}) and L2L^{2} language reads as (note that ω=dμ=0\omega=\,{\rm d}\mu=0): Find 𝝈H0(div)\boldsymbol{\sigma}\in H_{0}(\operatorname{div}) and μL02(Ω)\mu\in L_{0}^{2}(\Omega) such that

(54) (𝝈t,𝝉)+(μ,div𝝉)\displaystyle(\boldsymbol{\sigma}_{t},\boldsymbol{\tau})+(\mu,\operatorname{div}\boldsymbol{\tau}) =0𝝉H0(div),\displaystyle=0\qquad\forall~{}~{}\boldsymbol{\tau}\in H_{0}(\operatorname{div}),
(55) (μt,v)(div𝝈,v)\displaystyle(\mu_{t},v)-(\operatorname{div}\boldsymbol{\sigma},v) =(f,v)vL02(Ω).\displaystyle=(f,v)\qquad\forall~{}~{}v\in L_{0}^{2}(\Omega).

This formulation is the mixed method for acoustic wave equations [22]. We choose the exact solutions as

(56) 𝝈(x,y,t)\displaystyle\boldsymbol{\sigma}(x,y,t) =πet(cos(πx)sin(πy)sin(πx)cos(πy)),\displaystyle=-\pi e^{-t}\begin{pmatrix}\cos(\pi x)\sin(\pi y)\\ \sin(\pi x)\cos(\pi y)\end{pmatrix},
(57) μ(x,y,t)\displaystyle\mu(x,y,t) =etsin(πx)sin(πy),\displaystyle=e^{-t}\sin(\pi x)\sin(\pi y),

and pick initial conditions 𝝈0=𝝈(x,y,0)\boldsymbol{\sigma}_{0}=\boldsymbol{\sigma}(x,y,0) and μ0=μ(x,y,0)\mu_{0}=\mu(x,y,0). We use RT1RT_{1} element to discrete 𝝈\boldsymbol{\sigma} and use discontinuous piecewise linear polynomials to discrete μ\mu, the numerical results are listed in Table 4.

Table 4. Errors and convergence orders in various norms with Δt=0.0001\,\Delta t=0.0001 at t=0.0004t=0.0004.
hh σσh\|\sigma-\sigma_{h}\| div(σσh)\|\operatorname{div}(\sigma-\sigma_{h})\| μμh\|\mu-\mu_{h}\|
1/4 5.6058e-02 3.8201e-01 1.9350e-02
1/8 1.4002e-02 9.6901e-02 4.9051e-03
1/16 3.4958e-03 2.4360e-02 1.2312e-03
order 2.002 1.985 1.992

From Table 4, we can see that the mixed finite element method is of second order convergence rate for all the variables.

Acknowledgments

We would like to thank Professor Long Chen from University of California at Irvine for valuable discussion and suggestions.

References

  • [1] D. N. Arnold, R. S. Falk, and R. Winther. Finite element exterior calculus, homological techniques, and applications. Acta Numerica, 15:1–155, 2006.
  • [2] D. N. Arnold, R. S. Falk, and R. Winther. Finite element exterior calculus: from hodge theory to numerical stability. Bulletin of the American Mathematical Society, 47(2):281–354, 2010.
  • [3] D. N. Arnold. Finite element exterior calculus. SIAM, 2018.
  • [4] G. A. Baker. Error estimates for finite element methods for second order hyperbolic equations. SIAM Journal on Numerical Analysis, 13(4):564–576, 1976.
  • [5] F. Brezzi and M. Fortin. Mixed and hybrid finite element methods. Springer-Verlag, New York, 1991.
  • [6] L. Chen and Y. Wu. Convergence of adaptive mixed finite element methods for the hodge laplacian equation: without harmonic forms. SIAM Journal on Numerical Analysis, 55(6):2905–2929, 2017.
  • [7] L. Chen, Y. Wu, L. Zhong, and J. Zhou. Multigrid preconditioners for mixed finite element methods of the vector laplacian. Journal of Scientific Computing, 77:101–128, 2018.
  • [8] S. H. Christiansen, H. Z. Munthe-Kaas and B. Owren. Topics in structure-preserving discretization. Acta Numerica, pp 1-119, 2011.
  • [9] E. T. Chung and B. Engquist. Optimal discontinuous galerkin methods for the acoustic wave equation in higher dimensions. SIAM Journal on Numerical Analysis, 47(5):3820–3848, 2009.
  • [10] M. Costabel and A. McIntosh. On Bogovskil̆ and regularized Poincaré integral operators for de Rahm complexes on Lipschitz domains. Math. Z., 265: 297–320, 2010.
  • [11] L. C. Cowsat, T. F. Dupont, and M. F. Wheeler. A priori estimates for mixed finite element methods for the wave equation. Computer Methods in Applied Mechanics and Engineering, 82(1-3):205–222, 1990.
  • [12] L. Demkowicz, P. Monk, L. Vardapetyan, and W. Rachowicz. De Rham diagram for hphp finite element spaces. Math. Comput. Applications, 39(7-8):29-38, 2000.
  • [13] L. Demkowicz, I. Babuška. Optimal pp interpolation error estimates for edge finite elements of variable order in 2D. SIAM J. Numer. Anal., 41(4):1195–1208, 2003.
  • [14] L. Demkowicz and A. Buffa. H1H^{1}, H(curl)H({\rm curl\,}) and H(div)H(\operatorname{div})-conforming projection-based interpolation in three dimensions Quasi-optimal pp-interpolation estimates. Comput. Methods Appl. Mech. Engrg, 194: 267 – 296, 2005.
  • [15] T. Dupont. L2-estimates for Galerkin methods for second order hyperbolic equations. SIAM Journal on Numerical Analysis, 10(5):880–889, 1973.
  • [16] D. French and T. Peterson. A continuous space-time finite element method for the wave equation. Mathematics of Computation, 65(214):491–506, 1996.
  • [17] T. Geveci. On the application of mixed finite element methods to the wave equations. ESAIM: Mathematical Modelling and Numerical Analysis, 22(2):243–250, 1988.
  • [18] V. Girault and P.-A. Raviart. Finite element methods for Navier-Stokes equations: theory and algorithms, volume 5. Springer Science & Business Media, 2012.
  • [19] R. Glowinski and T. Rossi. A mixed formulation and exact controllability approach for the computation of the periodic solutions of the scalar wave equation. (i): Controllability problem formulation and related iterative solution. Academie Des Sciences Comptes Rendus Mathematique, 343(7):493–498, 2006.
  • [20] R. Griesmaier and P. Monk. Discretization of the wave equation using continuous elements in time and a hybridizable discontinuous Galerkin method in space. J Sci Comput, 58: 472-498.
  • [21] E. W. Jenkins. Numerical solution of the acoustic wave equation using raviart–thomas elements. Journal of Computational and Applied Mathematics, 206(1):420–431, 2007.
  • [22] R. C. Kirby and T. T. Kieu. Symplectic-mixed finite element approximation of linear acoustic wave equations. Numerische Mathematik, 130(2):257–291, Oct 2014.
  • [23] D. Mitrea, M. Mitrea, and M. Taylor. Layer potentials, the Hodge Laplacian, and global boundary problems in nonsmooth Riemannian manifolds, volume 713. American Mathematical Soc., 2001.
  • [24] M. Mitrea. Dirichlet integrals and gaffney-friedrichs inequalities in convex domains. Forum Math, 13:531–567, 2001.
  • [25] J. T. Oden, L. Demkowicz, W. Rachowicz, and T. A. Westermann. Towards a universal hh-pp adaptive finite element strategy, Part2. A posteriori error estimation. Comput. Methods Appl. Meth. Eng., 77: 113–180, 1989.
  • [26] V. Quenneville-Bélair. A new approach to finite element simulations of general relativity. Ph. D. thesis, University of Minnesota, 2015.
  • [27] P. A. Raviart and J. M. Thomas. A mixed finite element method for 2-nd order elliptic problems. Mathematical aspects of finite element methods, pages 292–315, 1977.
  • [28] B. Riviére and M. F. Wheeler. A priori error estimates for mixed finite element approximations of the acoustic wave equation. SIAM Journal on Numerical Analysis, 40(5):1698–1715, 2002.