• Aucun résultat trouvé

Superconvergence Analysis for Time-Dependent Maxwell’s Equations in Metamaterials

N/A
N/A
Protected

Academic year: 2022

Partager "Superconvergence Analysis for Time-Dependent Maxwell’s Equations in Metamaterials"

Copied!
23
0
0

Texte intégral

(1)

Superconvergence Analysis for Time-Dependent Maxwell’s Equations in Metamaterials

Yunqing Huang,1Jichun Li,2Qun Lin3

1Hunan Key Laboratory for Computation and Simulation in Science and Engineering, Xiangtan University, Xiangtan, China

2Department of Mathematical Sciences, University of Nevada Las Vegas, Nevada 89154-4020

3LSEC, ICMSEC, Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing 100080, China

Received 31 July 2010; accepted 22 June 2011

Published online 1 September 2011 in Wiley Online Library (wileyonlinelibrary.com).

DOI 10.1002/num.20703

In this article, we consider the time-dependent Maxwell’s equations modeling wave propagation in meta- materials. One-order higher global superclose results in theL2norm are proved for several semidiscrete and fully discrete schemes developed for solving this model using nonuniform cubic and rectangular edge elements. Furthermore,Lsuperconvergence at element centers is proved for the lowest order rectangular edge element. To our best knowledge, such pointwise superconvergence result and its proof are original, and we are unaware of any other publications on this issue. © 2011 Wiley Periodicals, Inc. Numer Methods Partial Differential Eq 28: 1794–1816, 2012

Keywords: Maxwell’s equations; metamaterial; mixed finite element method; superconvergence

I. INTRODUCTION

The metamaterials we are interested in are those artificially constructed electromagnetic nanoma- terials with negative refraction index. The successful demonstration of such metamaterials in 2000 triggered a big wave in further study of such metamaterials and exploration of their applications in diverse areas such as subwavelength imaging and design of invisibility cloak. More details can be found in monographs such as [1–4] and references cited therein.

Because of the tremendous cost of metamaterials, numerical simulation plays a very important role in the investigation of wave propagation involving metamaterials. However, simulations are almost exclusively based on either the classic finite-difference time-domain (FDTD) method or

Correspondence to:Jichun Li, Department of Mathematical Sciences, University of Nevada Las Vegas, NV 89154-4020 (e-mail: jichun@unlv.nevada.edu)

Contract grant sponsor: NSFC; contract grant number: 11031006

Contract grant sponsor: Hunan Provicial NSF; contract grant number: 10JJ7001

Contract grant sponsor: National Science Foundation; contract grant number: DMS-0810896

© 2011 Wiley Periodicals, Inc.

(2)

commercial packages such as HFSS and COMSOL [5, p. 12]. Because of their limited capabilities (e.g., FDTD is not well suited for problems involving complex geometries, and COMSOL runs out memory so quickly for 3D simulations), there is an urgent need for developing more efficient and reliable software for metamaterial simulations [5, p. 28]. To our best knowledge, although there are many excellent work for the Maxwell’s equations in vacuum (see, e.g., articles [6–11], books [12–14], and references cited therein), there is not much work devoted to the development and analysis of finite element methods (FEMs) for the Maxwell’s equations involving metamaterials.

Fernandes and Raffetto [15, 16] initiated the study of well posedness and finite element analy- sis for time-harmonic Maxwell’s equations involving metamaterials. In recent years, we made some initial effort in developing and analyzing some FEM for time-domain Maxwell’s equations involving metamaterials [17–19].

In our recent numerical experiments [17, 18], we found that superconvergence phenomena exist for metamaterial simulations using FEMs. It is well known [20–22] that superconvergence results often happen when the underlying differential equations have smooth solutions and are solved on very structured meshes such as rectangular grids or strongly regular triangular grids. Many excellent superconvergence results have been obtained for elliptic and parabolic problems (e.g., [23–26]). However, there exist not many superconvergence results for Maxwell’s equations. In 1994, Monk [27] obtained the first superconvergence result for Maxwell’s equations in vacuum.

Later, Brandts [28] presented another superconvergence analysis for 2D Maxwell’s equations in vacuum. Also, Lin and Coworkers [29–31] systematically developed some global superconver- gence results using the so-called Lin’s integral identity technique [21, 32–34] developed in early 1990s. In 2008, Lin and Li [35] extended the superconvergence result for the vacuum case to three popular dispersive media models. However, all existing results are limited to semi-discrete schemes and obtained only in theL2-norm. Here, we extend our previous work [35] to the meta- material case. Superconvergence results are obtained for both semidiscrete and fully discrete schemes based on nonuniform cubic meshes. Furthermore, the present results are improved in that the convergence constant depends on time linearly, instead of exponentially because of the use of Gronwall’s inequality.

In this article,C > 0 denotes a generic constant, which is independent of the finite element mesh sizehand time step sizeτ. We also introduce some common notation [14]

H (div;)=

v(L2())3; ∇ ·v(L2())3 , H (curl;)=

v(L2())3; ∇ ×v(L2())3 , H0(curl;)= {vH (curl;); n×v=0 on},

whereα ≥ 0 is a real number, andis a rectangular-type domain inR3with boundary ∂.

Let (Hα())3 be the standard Sobolev space equipped with the norm · α and seminorm

| · |α. Specifically, · 0 will mean the (L2())3-norm. We equip H (curl;) with the norm v0,curl =(v20+ curlv20)1/2. For clarity, in the rest of this article we introduce the vector notation

L2()= L2()3

, Hα()=(Hα())3.

The rest of this article is organized as follows. In Section II, we present the governing equa- tions for metamaterials. In Section III, we develop two different semidiscrete schemes and prove the superclose results for both schemes. Then, in Section IV, we present a fully discrete scheme and its superclose analysis. Because of some major differences between the 2D and 3D cases,

(3)

in Section V, we extend the 3D results to 2D. Especially, we prove theLsuperconvergence at element centers for the lowest order rectangular edge element. Finally, we conclude the article in Section VI.

II. THE GOVERNING EQUATIONS

The governing equations for modeling wave propagation in metamaterials are [19]:

0∂E

∂t = ∇ ×HJ, in×(0,T ), (1) μ0

∂H

∂t = −∇ ×EK, in×(0,T ), (2)

∂J

∂t +eJ =0ω2peE, in×(0,T ), (3)

∂K

∂t +mK =μ0ω2pmH, in×(0,T ), (4) where0andμ0are the permittivity and permeability in vacuum respectively,ωpe andωpm are the electric and magnetic plasma frequencies, respectively,eandmare the electric and mag- netic damping frequencies, respectively,E(x,t )andH(x,t )are the electric and magnetic fields, respectively, andJ(x,t )andK(x,t )are the induced electric and magnetic currents, respectively.

For simplicity, we assume that the boundary ofis perfect conducting so that

n×E=0 on∂, (5)

wherenis the unit outward normal to∂. Furthermore, we assume that the initial conditions are E(x, 0)=E0(x), H(x, 0)=H0(x), (6) J(x, 0)=J0(x), K(x, 0)=K0(x), (7) whereE0,H0,J0, andK0are some given functions.

To simplify the presentation and numerical simulation, we first rescale the governing equations (1)–(4). By introducing the notation

˜ t= t

0μ0,˜m=m

0μ0,˜e =e

0μ0,ω˜2m=0μ0ωpm2 ,ω˜2e =0μ0ω2pe, E˜ =√

0μ0E, H˜ =μ0H, J˜ =μ0J, K˜ =√ 0μ0K,

it is not difficult to check that the Eqs. (1)–(4) can be written as

E˜

∂t˜ = ∇ × ˜H − ˜J, (8)

H˜

∂t˜ = −∇ × ˜E− ˜K, (9)

(4)

J˜

∂t˜ + ˜eJ˜ = ˜ω2eE,˜ (10)

K˜

∂t˜ + ˜mK˜ = ˜ω2mH˜, (11) which have the same form as the original governing equations (1)–(4) if we set0=μ0=1 in (1)–(4). In the rest of this article, our discussion is based on the nondimensionalized form (8)–(11) by dropping all those tildes.

We like to remark that solving the metamaterial model (8)–(11) is quite challenging, because the governing equations cannot be reduced to a simple vector wave equation as in vacuum. Solving (10) yields

J(x,t )=eetJ0+ω2e t

0

ee(ts)E(x,s)ds≡ −f + ˜˜J(E). (12) Similarly, from (11) we have

K(x,t )=emtK0+ωm2 t

0

em(ts)H(x,s)ds≡ −g+ ˜˜K(H). (13) Differentiating (8) with respect totand substituting (9) and (12) into it, we obtain

0=Et t− ∇ ×Ht+Jt

=Et t+ ∇ × ∇ ×E+ ∇ ×K+ω2eEe

eetJ0+ωe2 t

0

ee(t−s)E(x,s)ds

. (14) Taking curl of (13) and substituting (8) into it, we have

∇ ×K=emt∇ ×K0+ωm2 t

0

em(t−s)(Et+J)(x,s)ds. (15) Then replacingJof (15) by (12) and substituting the result into (14), we can obtain an equation involving one variableE. As the resulting equation is so complicated, it is not a good idea of solving just one variable for the Maxwell’s equations when metamaterials are involved. Hence, we better resort to the mixed FEM instead of the standard FEM.

III. 3D SUPERCLOSE ANALYSIS FOR SEMIDISCRETE SCHEMES

In this section, we will discuss two different ways of solving the Maxwell’s equations involving metamaterials. One way is to solve (8)–(11) directly. Another way is to reduce the system of (8)–(11) to two unknowns by using (12) and (13), i.e., in the integral-differential equation form:

∂E

∂t = ∇ ×H − ˜˜J(E)+f, (16)

∂H

∂t = −∇ ×E− ˜˜K(H)+g. (17)

(5)

To design our mixed FEM, we partition by a family of regular cubic meshes Th with maximum mesh sizeh. Consider the Raviart-Thomas-Nédélec cubic elements (e.g., [36] and [14]):

Uh =

ψhH (div;): ψh|KQk,k1,k1×Qk1,k,k1×Qk1,k1,k, ∀KTh , Vh =

φhH (curl;): φh|KQk−1,k,k×Qk,k−1,k×Qk,k,k−1, ∀KTh .

Here,Qi,j,kdenotes the space of polynomials whose degrees are less than or equal toi,j,kin variablesx,y,z, respectively. It is easy to see that

∇ ×VhUh. (18)

For superclose analysis, we need to define two operators. The first one is the standard L2()-projection operator: For anyHL2(),PhHUhsatisfies

(PhHH,ψh)=0, ∀ψhUh.

Another one is the Nédélec interpolation operator h, which is defined as: For any EH (curl;),hEVhsatisfies [36, p. 331]:

li

(EhE)·tqdl=0, ∀qPk−1(li),i=1,. . ., 12,

σi

((EhE)×n·qdσ =0, ∀qQk−2,k−1i)×Qk−1,k−2i),i=1,. . ., 6,

K

(EhE)·qdK=0, ∀qQk1,k2,k2×Qk2,k1,k2×Qk2,k2,k1,

whereli andσi are the edges and faces of an elementK,tis the unit tangent vector along the edgeli, andnis the unit normal vector on faceσi.

Furthermore, our superclose analysis depends on the following two fundamental results.

Lemma 3.1. [30, Lemma 3.1]. On any cubic elementK, for anyEHk+2(K), we have

K

∇ ×(EhE)·ψdxdydz=O(hk+1)Ek+2,Kψ0,K, ∀ψ|KUh(K).

Lemma 3.2. [30, Lemma 3.2]. On any cubic elementK, for anyEHk+1(K), we have

K

(EhE)·φdxdydz=O(hk+1)Ek+1,Kφ0,K, ∀φ|KVh(K)

Note that in [30], the results of Lemmas 3.1 and 3.2 were stated for the whole domain. But, the proofs of [30] actually show that the results hold true element-wisely. Readers can consult the original proofs of [30] and our detailed proof for rectangular elements presented in Section V. Below, we will discuss two different ways of solving the Maxwell’s equations (8)–(11).

(6)

A. The Semidiscrete Scheme for an Integral-Differential Formulation

The first way is to solve the system (8)–(11) in the reduced integral-differential equation form:

∂E

∂t = ∇ ×H − ˜˜J(E)+f, (19)

∂H

∂t = −∇ ×E− ˜˜K(H)+g. (20) Assuming existence of smooth solutions to (19) and (20), we can obtain its corresponding weak formulation: For anyt(0,T], find the solution(E,H)H0(curl;)×L2()such that (Et,φ)(H,∇ ×φ)+(J˜˜(E),φ)=(f,φ),φH0(curl;), (21) (Ht,ψ)+(∇ ×E,ψ)+(K(H˜˜ ),ψ)=(g,ψ),ψL2(), (22) with the initial conditions

E(x, 0)=E0(x), H(x, 0)=H0(x),x. (23) Note that the above derivation used the Stokes’ formula

∇ ×φ·Hdx=

φ· ∇ ×Hdx+

n×φ·Hds. (24) LetV0h= {vVh :v×n=0 on}. Now we can construct a semidiscrete mixed FEM for solving (21) and (22): For anyt(0,T], find the solution(Eh,Hh)V0h×Uhsuch that

Eht,φh

(Hh,∇ ×φh)+(J˜˜(Eh),φh)=(f,φh),φhV0h, (25) Hht,ψh

+(∇ ×Eh,ψh)+(K˜˜(Hh),ψh)=(g,ψh),ψhUh, (26) with the initial conditions

Eh(x, 0)=hE0(x), Hh(x, 0)=PhH0(x),x. (27) For this scheme, we have the following superclose result:

Theorem 3.1. Let(E,H)and(Eh,Hh)be the analytic and finite element solutions of (21) and (22) and (25) and (26) at timet(0,T],respectively. Under the regularity assumptions

EtL(0,T;Hk+1()), EL(0,T;Hk+2()),

wherek≥1is the order of the basis functions in spacesUhandVh. Then, there exists a constant C >0independent of the mesh sizehbut linearly dependent onT, such that

hEEhL(0,T;L2())+ PhHHhL(0,T;L2())

Chk+1 EtL(0,T;Hk+1())+ EL(0,T;Hk+2())

.

(7)

Proof. Denoteξ =hEEh,η=PhHHh. Choosingφ=φh=ξ andψ =ψh=ηin (21) and (25), (22), and (26), respectively, and rearranging the resultants, we obtain the following error equations

t,ξ )(η,∇ ×ξ )+(J˜˜(ξ ),ξ )=((hEE)t,ξ )(PhHH,∇ ×ξ ) +(J˜˜(hEE),ξ ),

t,η)+(∇ ×ξ,η)+(K(η),˜˜ η)=((PhHH)t,η)

+(∇ ×(hEE),η)+(K(P˜˜ hHH),η).

Using theL2-projection property, and the fact that∇ ×VhUh, we have 1

2 d dt

ξ20+ η20

+(J˜˜(ξ ),ξ )+(K(η),˜˜ η)

((hEE)t,ξ )+(J˜˜(hEE),ξ )+(∇ ×(hEE),η). (28) By Lemma 3.2 and the Cauchy-Schwarz inequality, we have

((hEE)t,ξ )Chk+1EtL(0,T;Hk+1())ξL(0,T;L2()), and

(J˜˜(hEE),ξ )=ω2e t

0

ee(t−s)(hEE,ξ )ds

ω2e t

0

ee(t−s)·Chk+1Ek+1ξ0ds

Chk+1EL(0,T;Hk+1())ξL(0,T;L2()), where in the last step we used the estimate

t 0

ee(t−s)ds = 1

e(1eet)≤ 1 e

and absorbed the physical parametersωeandeinto the generic constantC.

Similarly, by Lemma 3.1 we have

(∇ ×(hEE),η)Chk+1EL(0,T;Hk+2())ηL(0,T;L2()).

Integrating (28) from 0 tot, using the fact thatξ(0)=η(0)=0 and substituting the above estimates into the resultant, we obtain

1 2

ξ(t )20+ η(t )20

t·Chk+1 EtL(0,T;Hk+1())ξL(0,T;L2())

+EL(0,T;Hk+1())ξL(0,T;L2())+ EL(0,T;Hk+2())ηL(0,T;L2())

,

(8)

where we used the inequality t 0

(J˜˜(ξ ),ξ )dt ≥0, t

0

(K(η),˜˜ η)dt ≥0, (29) which results from Remark 3.1.

Finally, using the arithmetic-geometry mean inequality(ab1a2+δb2), then taking the maximum with respect tot, we have

ξ2

L(0,T;L2())+ η2

L(0,T;L2())Ch2(k+1) Et2

L(0,T;Hk+1())+ E2

L(0,T;Hk+2())

, whereCis linearly dependent onT2. The proof completes by using the triangle inequality and the standard interpolation and projection error estimates.

Remark 3.1. According to [37, (1.2)], a real-valued kernel β(t )L1(0,T ) is called postive-definite if for eachT >0,βsatisfies

T 0

φ (t ) t

0

β(ts)φ (s)dsdt ≥0, ∀φC[0,T]. (30) From Plancherel’s theorem, for any kernelβL1(0,), (30) holds if and only if

0

β(t )cos(αt )dt≥0, ∀α >0. (31) It is easy to check that the kernelβ(t )=eetsatisfies (31), hence theβinJ˜˜ is a positive-definite kernel.

Remark 3.2. Applying the interpolation estimates and inverse inequality to the superclose result, we can easily obtain the following optimalLerror estimate

EEhL(0,T;L())+ HHhL(0,T;L())Chk.

Using the postprocessing operators ˆ2h and 2h defined in [30, 35], we can achieve the following global superconvergence result:

ˆ2hEEhL(0,T;L2())+ 2hHHhL(0,T;L2())Chk+1.

B. The Semidiscrete Scheme for a Pure Differential Formulation

Here, we consider a semidiscrete scheme for solving the system (8)–(11) directly. Its correspond- ing weak formulation is as follows: For anyt(0,T], find the solutionsEH0(curl;), JH (curl;),H, andKL2()such that

(Et,φ)(H,∇ ×φ)+(J,φ)=0, ∀φH0(curl;), (32) (Ht,ψ)+(∇ ×E,ψ)+(K,ψ)=0, ∀ψL2(), (33) (Jt,φ)˜ +e(J,φ)˜ω2e(E,φ)˜ =0, ∀φ˜H (curl;), (34) (Kt,ψ)˜ +m(K,ψ˜)ω2m(H,ψ˜)=0, ∀ψ˜L2(), (35) with the initial conditions (6) and (7).

(9)

We can construct a semidiscrete mixed method for solving (32)–(35): For anyt(0,T], find the solutionsEhV0h,JhVh,Hh,KhUhsuch that

Eht,φh

(Hh,∇ ×φh)+(Jh,φh)=0, ∀φhV0h, (36) Hht,ψh

+(∇ ×Eh,ψh)+(Kh,ψh)=0, ∀ψhUh, (37) Jht,φ˜h

+e(Jh,φ˜h)ω2e(Eh,φ˜h)=0, ∀φ˜hVh, (38) Kht,ψ˜h

+m(Kh,ψ˜h)ω2m(Hh,ψ˜h)=0, ∀ψ˜hUh, (39) with the initial approximations

E0h(x)=hE0(x), H0h(x)=PhH0(x), J0h(x)=hJ0(x), K0h(x)=PhK0(x).

For this scheme, we have the following superclose result:

Theorem 3.2. Let (E,H,J,K) and (Eh,Hh,Jh,Kh) be the analytic and finite element solutions of (32)–(35) and (36)–(39) at time t(0,T], respectively. Under the regularity assumptions

Et,Jt,JL(0,T;Hk+1()), EL(0,T;Hk+2()),

wherek≥1is the order of the basis functions in spacesUhandVh. Then, there exists a constant C >0independent of the mesh sizehbut linearly dependent onT, such that

hEEhL(0,T;L2())+ PhHHhL(0,T;L2())

+ 1

ωehJJhL(0,T;L2())+ 1

ωmPhKKhL(0,T;L2())

Chk+1 EtL(0,T;Hk+1())+JL(0,T;Hk+1())+EL(0,T;Hk+2())+JtL(0,T;Hk+1())

. Proof. Denoteξ =hEEh,η =PhHHh,ξ˜ =hJJh, andη˜ = PhKKh. Choosingφ =φh =ξ in (32) and (36),ψ =ψh =ηin (33) and (37),φ˜ =φ˜h = ˜ξ in (34) and (38), andψ˜ =ψ˜h = ˜ηin (35) and (39), respectively, and rearranging the resultants, we obtain the error equations

t,ξ )(η,∇ ×ξ )+(˜ξ,ξ )=((hEE)t,ξ )(PhHH,∇ ×ξ )+(hJJ,ξ ), t,η)+(∇ ×ξ,η)+(η,˜ η)=((PhHH)t,η)+(∇ ×(hEE),η)+(PhKK,η), (˜ξt,ξ )˜ +e(˜ξ,ξ )˜ −ωe2,ξ )˜ =((hJJ)t,ξ )˜ +e(hJJ,ξ )˜ −ωe2(hEE,ξ ),˜ ˜t,η)˜ +m(η,˜ η)˜ −ωm2(η,η)˜ =((PhKK)t,η)˜ +m(PhKK,η)˜ −ω2m(PhHH,η).˜

(10)

Dividing the last two equations byω2eandωm2, respectively, then adding the above four equations together, we obtain

1 2

d dt

ξ20+ η20+ 1

ω2e˜ξ20+ 1 ω2m ˜η20

+ e

ω2e˜ξ20+m ω2m ˜η20

=((hEE)t,ξ )+(hJJ,ξ )+(∇ ×(hEE),η) + 1

ωe2((hJJ)t,ξ )˜ +e

ωe2(hJJ,ξ )˜ −(hEE,ξ ),˜ (40) where we used theL2-projection property in the above derivation.

Using Lemmas 3.1 and 3.2 and the Cauchy-Schwarz inequality, we have ((hEE)t,ξ )Chk+1EtL(0,T;Hk+1())ξL(0,T;L2()), (hJJ,ξ )Chk+1JL(0,T;Hk+1())ξL(0,T;L2()), (∇ ×(hEE),η)Chk+1EL(0,T;Hk+2())ηL(0,T;L2()),

1

ω2e((hJJ)t,ξ )˜ ≤ 1

ωe2 ·Chk+1JtL(0,T;Hk+1())˜ξL(0,T;L2()), e

ω2e(hJJ,ξ )˜ ≤ e

ω2e ·Chk+1JL(0,T;Hk+1())˜ξL(0,T;L2()), (hEE,ξ )˜ ≤Chk+1EL(0,T;Hk+1())˜ξL(0,T;L2()).

Substituting the above estimates into (40) and following the same technique as we used in Theorem 3.1, we obtain

ξ2

L(0,T;L2())+ η2

L(0,T;L2())+ 1 ω2e˜ξ2

L(0,T;L2())+ 1 ω2m ˜η2

L(0,T;L2())

Ch2(k+1) Et2

L(0,T;Hk+1())+ J2

L(0,T;Hk+1())

+E2

L(0,T;Hk+2())+ Jt2

L(0,T;Hk+1())

, where the constantCis linearly dependent onT2.

Remark 3.3. The same remark as Remark 3.2 holds true for this case.

IV. 3D SUPERCLOSE ANALYSIS FOR FULLY DISCRETE SCHEMES

To define our fully discrete scheme, we divide the time interval(0,T )intoMuniform subintervals by points 0 = t0 < t1 <· · · < tM = T, wheretm =, and denote them-th subinterval by Im= [tm−1,tm]. Moreover, we defineum=u(·,mτ )for 0≤mM, and the notation:

δτum=umum−1

τ , u¯m= 1

2(um+um−1).

(11)

Now we can formulate the Crank-Nicolson mixed finite element scheme for solving (32)–(35):

Form=1, 2,. . .,M, findEmhV0h,JmhVh,Hmh,KmhUhsuch that δτEmh,φh

H¯mh,∇ ×φh

+J¯mh,φh

=0, ∀φhV0h, (41) δτHmh,ψh

+

∇ × ¯Emh,ψh

+K¯mh,ψh

=0, ∀ψhUh, (42) δτJmh,φ˜h

+eJ¯mh,φ˜h

ω2eE¯mh,φ˜h

=0, ∀φ˜hVh, (43) δτKmh,ψ˜h

+mK¯mh,ψ˜h

ω2mH¯mh,ψ˜h

=0, ∀ψ˜hUh, (44) subject to the initial approximations

E0h(x)=hE0(x), H0h(x)=PhH0(x), (45) J0h(x)=hJ0(x), K0h(x)=PhK0(x). (46) For this fully discrete scheme, we have the following superclose result:

Theorem 4.1. Let (Em,Hm,Jm, and Km) and (Emh,Hmh,Jmh, and Kmh) be the analytic and finite element solutions of (32)–(35) and (41)–(44) at timetm, respectively. Under the regularity assumptions

Et,Jt,JL(0,T;Hk+1()), EL(0,T;Hk+2()), Et t,Ht t,Jt t,Kt t,∇ ×Et t,∇ ×Ht tL(0,T;L2()),

wherek≥1is the order of the basis functions in spacesUhandVh. Then, there exists a constant C >0independent of the mesh sizehbut linearly dependent onT, such that

1maxmM

hEmEmh

0+PhHmHmh

0+hJmJmh

0+PhKmKmh

0)

Chk+1 EtL(0,T;Hk+1())+JL(0,T;Hk+1())+EL(0,T;Hk+2())+JtL(0,T;Hk+1())

+2 ∇ ×Ht tL(0,T;L2())+ Jt tL(0,T;L2())+ ∇ ×Et tL(0,T;L2())

+ Kt tL(0,T;L2())+ Et tL(0,T;L2())+ Ht tL(0,T;L2())

. Proof. Integrating (32)–(35) in time overImand dividing all byτ, we have

τEm,φ)− 1

τ

Im

H(s)ds,∇ ×φ

+ 1

τ

Im

J(s)ds,φ

=0, (47)

τHm,ψ)+

∇ ×1 τ

Im

E(s)ds,ψ

+ 1

τ

Im

K(s)ds,ψ

=0, (48)

τJm,φ)˜ +e 1

τ

Im

J(s)ds,φ˜

ωe2 1

τ

Im

E(s)ds,φ˜

=0, (49)

τKm,ψ˜)+m 1

τ

Im

K(s)ds,ψ˜

ωm2 1

τ

Im

H(s)ds,ψ˜

=0. (50)

Références

Documents relatifs

The perturbation Stochastic Finite Element Method (SFEM) is used to determine the mean and the variance of the thermoelastic quality factor and is compared to direct

1.  êà÷åñòâå ïðèìåðà íà ðèñ. Ïîäìíîæåñòâà ÿ÷ååê ñåòêè, ïîëó÷åííûå ïðè áëî÷íîì âàðèàíòå

We solve this problem with RT 0 and BDM 1 mixed finite elements on triangular meshes, which are obtained by regular refinement with the initial mesh showed in Figure 8.. The same as

It has been shown that these techniques and arguments are powerful for convergence analysis especially for improving accuracy analysis in mixed finite element methods for solving

— An asymptotic expansion of the mixed finite element solution of a linear parabolic problem is used to dérive superconvergence results.. Optimal order error estimâtes

In this paper, we consider a similar problem described in [8] for a Galerkin method to the elliptic équations on the unit square with Dirichlet boundary condition based on

For the approximation of solutions for radiation problems using intégral équations we refer to [3], [9], [16], [17] in the case of smooth boundaries and to [21] in the case of

Abstract — Auxihary theorems allowing to extend the theory of curved finite éléments introduced m [2], [3] to the case ofboundary value problems with varions stable and