• Aucun résultat trouvé

On the approximation of electromagnetic fields by edge finite elements. Part 4: analysis of the model with one sign-changing coefficient

N/A
N/A
Protected

Academic year: 2022

Partager "On the approximation of electromagnetic fields by edge finite elements. Part 4: analysis of the model with one sign-changing coefficient"

Copied!
32
0
0

Texte intégral

(1)

HAL Id: hal-03273264

https://hal.inria.fr/hal-03273264

Preprint submitted on 29 Jun 2021

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

On the approximation of electromagnetic fields by edge finite elements. Part 4: analysis of the model with one

sign-changing coefficient

Patrick Ciarlet

To cite this version:

Patrick Ciarlet. On the approximation of electromagnetic fields by edge finite elements. Part 4:

analysis of the model with one sign-changing coefficient. 2021. �hal-03273264�

(2)

(will be inserted by the editor)

On the approximation of electromagnetic fields by edge finite elements. Part 4: analysis of the model with one sign-changing coefficient

Patrick Ciarlet Jr.

This version: June 29, 2021

Abstract In electromagnetism, in the presence of a negative material sur- rounded by a classical material, the electric permittivity, and possibly the magnetic permeability, can exhibit a sign-change at the interface. In this set- ting, the study of electromagnetic phenomena is a challenging topic. We focus on the time-harmonic Maxwell equations in a bounded setofR3, and more precisely on the numerical approximation of the electromagnetic fields by edge finite elements. Special attention is paid to low-regularity solutions, in terms of the Sobolev scale (Hs(Ω))s>0. With the help of T-coercivity, we address the case of one sign-changing coefficient, both for the model itself, and for its discrete version. Optimal a priori error estimates are derived.

Introduction

We study the numerical approximation by finite elements of electromagnetic fields governed by the time-harmonic Maxwell equations, in the presence of a negative material surrounded by a classical material. A negative material can be a metal at optical frequencies, or a metamaterial, see for instance [46, 1]. So, in this setting, the electric permittivity, and possibly the magnetic permeabil- ity, can exhibit a sign-change at the interface between the two materials. We consider such a model in a bounded set ofR3, supplemented with a vanishing boundary condition on the tangential trace. To the author’s knowledge, the first attempt to address this situation theoretically can be found in [10, 9]; see also [15, 42]. However, little is known regarding the numerical approximation of the model. In the present paper, we provide the numerical analysis for a model with one sign-changing coefficient. In a similar setting, we refer to [34] for the analysis of the two-dimensional time-harmonic Maxwell equations, and to [22]

for the analysis of the three-dimensional, div-curl, or div-curlcurl, problem.

POEMS, CNRS, INRIA, ENSTA Paris, Institut Polytechnique de Paris, 91120 Palaiseau, France (E-mail: patrick.ciarlet@ensta-paris.fr).

(3)

For the numerical approximation, we use (low-order) edge finite elements. We use some recent results [21, 31, 22] to interpolate low-regularity solutions that can occur both in a classical setting, that is for a model with fixed-sign, piece- wise smooth coefficients [24, 6, 23], and in the presence of an interface between a classical material and a negative material.

In what follows, we shall assume that the electric permittivity εhas a sign- change, while the magnetic permeability µ has a fixed sign (when the roles of ε and µ are reversed, we refer to section 8). Typically, this corresponds to an interface model between a metal surrounded by a classical material (in some ad hoc frequency range). Classically [2,§8], for solving the time-harmonic Maxwell equations, one can choose first-order formulations in both the electric and magnetic fields, or second-order formulations in the electric field only, or in the magnetic field only. Our choice will be a second-order formulation in the electric field.

The outline is as follows. We begin by introducing some notations, together with a precise definition of the mathematical framework considered hereafter.

Before investigating the solution of this problem, we propose some comments in section 2 to help identify the difficulties to be addressed. For that, we rely on some well-known facts regarding the classical setting (fixed-sign coefficient), that we shall apply to the new model. We introduce the companion scalar problem and tools, such as the T-coercivity to realize the inf-sup condition.

In section 3, we explain how to solve the time-harmonic Maxwell equations.

Next, in section 4, we recall the numerical approximation via edge finite el- ements, and in particular how one can interpolate the electric field, which can (possibly) be of low-regularity. To prove the results regarding convergence of the numerical method, we use some results regarding practical discrete T- coercivity (for the companion scalar problem) which is achieved with the help of T-conform meshes. These are recalled in the appendix A. As a matter of fact, these results allow us to prove the uniform discrete inf-sup condition for the time-harmonic Maxwell equations: this is the object of the next two sec- tions. Then, in section 7, we provide a numerical illustration to check that the expected convergence order is achieved, and how the use of T-conform meshes may impact the convergence rate. In section 8, we outline how one can solve theoretically and numerically the case ofµhaving a sign-change, andεhaving a fixed sign. Finally, we give some concluding remarks in the last section.

1 Setting of the problem

As in [21], we denote constant fields by the symbolcst. Vector-valued (respec- tively tensor-valued) function spaces are written in boldface character (resp.

blackboard bold characters). Unless otherwise specified, we consider spaces of real-valued functions. Given a non-empty open setOofR3, we use the notation (·|·)0,O (respectivelyk · k0,O) for theL2(O) and theL2(O) := (L2(O))3 inner products (resp. norms). More generally, (·|·)s,O andk·ks,O (respectively|·|s,O) denote the inner product and the norm (resp. semi-norm) of the Sobolev spaces

(4)

Hs(O) andHs(O) := (Hs(O))3 for s∈ R(resp. for s>0). The index zmv indicates zero-mean-value fields. If moreover the boundary∂Ois Lipschitz,n denotes the unit outward normal vector field to ∂O. It is assumed that the reader is familiar with function spaces related to Maxwell’s equations, such as H(curl;O), H0(curl;O), H(div;O), H0(div;O) etc. A priori, H(curl;O) is endowed with the “natural” normv 7→(kvk20,O+kcurlvk20,O)1/2, etc. We refer to the monographs of Monk [40] and Assous et al [2] for details.

The symbolCis used to denote a generic positive constant which is indepen- dent of the meshsize, the mesh and the fields of interest ; C may depend on the geometry, or on the coefficients defining the model. We use the notation A.B for the inequalityACB, whereA andB are two scalar fields, and C is a generic constant.

Letbe adomaininR3, ie. an open, connected and bounded subset ofR3 with a Lipschitz-continuous boundary∂Ω. The domainΩcan betopologically trivial(tt) ornon-trivial(tnt) [33]. This means that we assume that one of the two conditions below holds:

(tt) ’for all curl-free vector field vC1(Ω), there exists pC0(Ω)such that v=∇pin Ω’;

(tnt)’there existInon-intersecting, piecewise plane manifolds,j)j=1,···,I, with boundaries ∂Σi∂Ω, such that, if we letΩ˙ =\SI

i=1Σi, for all curl-free vector fieldv, there exists p˙∈C0( ˙Ω)such that v=∇p˙ inΩ’.˙ To simplify the computations (without restricting the scope of the study), we assume that the boundary∂Ω isconnected.

We let be surrounded by a perfect conductor. We recall that, for a given pulsationω >0, the time-harmonic Maxwell equations set incan be expressed in terms of the complex-valued electric fieldeonly. They write

FindeH0(curl;Ω)such that curl(µ−1curle)ω2εe=ıωj in divεe=%in Ω.

(1) Above, the real-valued coefficientεis the electric permittivity tensor and the real-valued coefficient µ is the magnetic permeability tensor. The complex- valued source termsjand%are respectively the current density and the charge density. They are related by the charge conservation equation

−ıω%+ divj= 0 inΩ. (2)

Classically, in (1), the equation divεe = % is implied by the second-order equationcurl(µ−1curle)−ω2εe=ıωj, together with the charge conservation equation (2), so it is omitted from now on. We fix thea priori regularity of the current density to jL2(Ω), which implies that %H−1(Ω), with dependencek%kH−1(Ω)=ω−1kdivjkH−1(Ω)ω−1kjkL2(Ω).

(5)

Finally, note that one can split the problem into two parts, where <(e)∈ H0(curl;Ω) is related to−=(j)∈L2(Ω), resp.=(e)∈H0(curl;Ω) is related to <(j) ∈ L2(Ω). So, we carry on with e standing either for <(e) or =(e), resp.f standing for−ω−1=(j) orω−1<(j), that is withreal-valued fields. One can check that the equivalent variational formulation inH0(curl;Ω) writes

FindeH0(curl;Ω) such that

aω(e,v) =ω2(f|v)0,Ω, ∀v∈H0(curl;Ω), (3) where

aω(u,v) := (µ−1curlu|curlv)0,Ωω2(εu|v)0,Ω, ∀u,vH0(curl;Ω).

Note that with these notations, one has divεe=−divf inΩ.

Then, the real-valued coefficient ξ ∈ {ε, µ} fulfills one of the two sets of conditions below, which we refer to as theclassical caseand theinterface case hereafter.

Classical case:

ξ is a real-valued, symmetric, measurable tensor field onΩ,

∃ξ, ξ+>0,∀z∈R3, ξ|z|2ξz·zξ+|z|2 a.e. inΩ. (4) Interface case:Ωis partitioned into the non-trivial partitionP := (Ωp)p=+,−, where± are domains, andδξfulfills (4), withδ|Ω+= +1 andδ|Ω =−1.

For our studies of the time-harmonic Maxwell equations in the electric field, we assume from now on that

εis as in theinterface case;µis as in theclassical case.

2 Some comments

Observe that if the electric field is curl-free, ie. curle = 0, then it may be written as e=∇pe for some peH01(Ω) (cf. Theorem 3.3.9 in [2], as ∂Ω is connected). Moreover, pe is such that divε∇pe =−divf in H−1(Ω). So to ensure well-posedness, one must make an assumption on thecompanion scalar problemwith Dirichlet boundary condition:

Find sH01(Ω)such that (ε∇s|∇q)0,Ω=hg, qiH1

0(Ω), ∀q∈H01(Ω), (5)

namely, that this scalar problem is well-posed. In other words,

∃C?>0, ∀g∈H−1(Ω), ∃!ssolution to (5), withkskH1

0(Ω)C?kgk−1,Ω. (6) To measure elements ofH01(Ω), we choose the normq7→ kqkH1

0(Ω):=k∇qk0,Ω. If the permittivityε were to fulfill (4), well-posedness of the scalar problem wouldautomaticallyhold, as an obvious consequence of the fact that (q, q0)7→

(6)

(ε∇q|∇q0)0,Ω defines an inner product on H01(Ω), whose associated norm is equivalent to thek · kH1

0(Ω)-norm.

However, in the present setting, since εis as in the interface case, this is an additional assumption, which is addressed with the help of T-coercivity [12, 8]. We recall the abstract framework below, see [20, 18] for details. Let V be a Hilbert space with normk · kV, anda(·,·) asymmetric, continuous bilinear form onV ×V. Then, the well-posedness of the problem

Find uV such thata(u, v) =hf, viV, ∀v∈V, (7) which reads

∃C >0, ∀f ∈V0, ∃!usolution to (7), withkukVCkfkV0, (8) can be addresssed as follows. One has to prove that the formaisT-coercive, cf. Theorem 1 and Remark 2 of [18]:

∃α >0, ∃T ∈ L(V), ∀v∈V, |a(v, T v)| ≥αkvk2V. (9) In other words, the operatorT realizes the classical inf-sup condition (see eg.

[5])explicitly.

Hence, for the scalar problem (5), and because ε is a symmetric tensor field, well-posedness is equivalent to (q, q0) 7→ (ε∇q|∇q0)0,Ω fulfilling an inf- sup condition:

∃γ0>0, ∀q∈H01(Ω), sup

q0∈H01(Ω)\{0}

|(ε∇q|∇q0)0,Ω| kq0kH1

0(Ω)

γ0kqkH1

0(Ω). (10) Or, as noted above, this is equivalent to

∃α0>0, ∃T0∈ L(H01(Ω)),∀q∈H01(Ω), |(ε∇q|∇(T0q))0,Ω| ≥α0k∇qk20,Ω. Note that the absolute value can be removed. Indeed, the quadratic mapping q7→(ε∇q|∇(T0q))0,Ω is continuous inH01(Ω) and vanishes only forq= 0, so it takes either positive, or negative, values everywhere in H01(Ω). Thus, (10) is also equivalent to

∃α0>0, ∃T0∈ L(H01(Ω)),

∀q∈H01(Ω), (ε∇q|∇(T0q))0,Ωα0k∇qk20,Ω. (11) To recapitulate, we assume from now on that

(10)-(11) holds for the companion scalar problem (5).

When we perform the numerical analysis, and in order to obtain explicit convergence rates between the exact and approximate solution to the time- harmonic Maxwell equations, we shall maketwo additional assumptions:

(7)

the coefficientsε, µare piecewise smooth: there exists a partition{Ωp}p=1,···,P

of Ω, made of disjoint domains (Ωp)p=1,···,P, with=∪Pp=1p, and such that ε|Ωp, µ|ΩpW1,∞(Ωp) for p= 1,· · ·, P. In relation to the partition and fors≥0, we define

P Hs(Ω) :={v∈L2(Ω) : v|ΩpHs(Ωp), 1≤pP}, (12) endowed with the “natural” normkvkP Hs(Ω):= X

1≤p≤P

kvpk2s,Ωp1/2

. the dataf has extra-regularity, in the sense that

divfH−1+τ0(Ω), withτ0∈(0,1] given. (13) For further analysis, let us introduce the scalar problem withmodified right- hand side

Find sH01(Ω)such that (ε∇s|∇q)0,Ω=hg, qiH1

0(Ω)+ (εg,∇q)0,Ω, ∀q∈H01(Ω). (14) Ifεwere as in the classical case [24, 43, 29, 38, 35, 6, 27], one could prove a shift theorem for the problem (14) when the data (g,g) has extra-regularitylike

gH−1+τ0(Ω), gH1(Ω), withτ0∈(0,1] given.

In the interface case, there exist similar results in this direction. We refer to [25, 13, 17, 16, 11] for a piecewise constant coefficientε. So we introduceτDir∈(0,1]

depending only on the geometry and onεsuch that

∀s∈[0, τDir)\ {1/2}, ∀(g,g)H−1+s(Ω)×H1(Ω), the solutionsto (14) is such thatsP H1+s(Ω), and kskP H1+s(Ω).(kgk−1+s,Ω+kgk1,Ω).

Above, the constant hidden in.may depend ons, but not ong nor ong. By a slight abuse of vocabulary, we call this result theshift theorem, respectively τDirthelimit regularity exponent. We assume from now on that

a shift theorem holds with τDir∈(0,1] for the modified scalar problem (14).

Remark 1 We shall also need a shift theorem for the scalar problem involving the magnetic permeabilityµwith Neumann boundary condition, see (31) be- low. Result can be found in the above-mentioned references, becauseµ is as

in the classical case. ut

So far we focused on curl-free fields. To tackle fields with a non-vanishing curl, we use an ad hoc splitting inH0(curl;Ω). Define

KN(Ω, ε) :={v∈H0(curl;Ω) : divεv = 0}.

An equivalent (variational) definition is

KN(Ω, ε) :={v∈H0(curl;Ω) : (εv|∇q)0,Ω= 0, ∀q∈H01(Ω)}.

(8)

Proposition 1 One has the continuous, direct sum

H0(curl;Ω) =∇[H01(Ω)]⊕KN(Ω, ε). (15) Proof Obviously, ∇[H01(Ω)] +KN(Ω, ε) is a subset of H0(curl;Ω). Let vH0(curl;Ω). According to (6), there existspvH01(Ω) such that

(ε∇pv|∇q)0,Ω= (εv|∇q)0,Ω, ∀q∈H01(Ω). (16) Now, letkv =v− ∇pv, one haskvKN(Ω, ε) by construction. It follows thatH0(curl;Ω) =∇[H01(Ω)] +KN(Ω, ε).

Next, let z ∈ ∇[H01(Ω)]∩KN(Ω, ε) be given. There exists sH01(Ω) such that z =∇s and, by definition of KN(Ω, ε), s is governed by (5) with zero right-hand side. By uniqueness of the solution, one has s= 0 and so z = 0:

the sum is direct.

Finally, by definition (16) ofpv and according to (11), one hasα0k∇pvk20,Ω≤ (ε∇pv|∇(T0pv))0,Ω= (εv|∇(T0pv))0,Ω≤ kεvk0,Ωk∇(T0pv)k0,Ω, so that

k∇pvkH(curl;Ω)=k∇pvk0,Ωα−10 kεk∞,ΩkT0kL(H1

0(Ω))kvk0,Ω, and kkvkH(curl;Ω)≤(1 +α−10 kεk∞,ΩkT0kL(H1

0(Ω)))kvkH(curl;Ω).

So the sum is continuous. ut

In other words, we may introduce the operators of L(H0(curl;Ω), H01(Ω)), resp. ofL(H0(curl;Ω))

π1 :

H0(curl;Ω)H01(Ω) v7→pv

, π2 :

H0(curl;Ω)KN(Ω, ε) v7→kv

and write, for allvH0(curl;Ω),v =∇(π1v) +π2v. Note that (π2)2=π2. We finally recall an important result on the measure of elements ofKN(Ω, ε).

For its proof, we refer the reader to Corollary 5.2 of [9].

Theorem 1 Elements of KN(Ω, ε) can be measured with the kcurl·k0,Ω- norm:

∃CW >0, ∀k∈KN(Ω, ε), kkk0,ΩCWkcurlkk0,Ω, (17)

∃CW0 >1, ∀k∈KN(Ω, ε), kkkH(curl;Ω)CW0 kcurlkk0,Ω. (18)

3 Solving the exact problem

Recall thatµis as in the classical case (cf. (4)), resp.εis as in the interface case, and assumption (10)-(11) holds. Using operators π1 and π2, one can provide an equivalent reformulation of the variational formulation (3). Its solution e may be split as

e=e0+∇φ, withe0=π2eandφ=π1e. (19)

(9)

By using the (variational) definition ofKN(Ω, ε) (recall thatεis a symmetric tensor field), we notice thate0 andφare respectively governed by

Find e0KN(Ω, ε)such that

aω(e0,v) =ω2(f|v)0,Ω, ∀v ∈KN(Ω, ε). (20) Find φH01(Ω)such that

(ε∇φ|∇q)0,Ω=hdivf, qiH1

0(Ω), ∀q∈H01(Ω). (21) Actually, there is an equivalence result (the proof is left to the reader).

Proposition 2 A fieldeis a solution to (3) if, and only if,π2e is a solution to (20) and π1e is a solution to (21).

According to the assumption on ε, we already know that problem (21) is well-posed. Hence proving the well-posedness of (3) amounts to proving the well-posedness of (20). We recall Theorem 8.15 of [9].

Theorem 2 The imbedding of KN(Ω, ε)in L2(Ω)is compact.

As a consequence (cf. Theorem 8.16 of [9]), one has the

Corollary 1 The variational formulation (20) with unknown e0 enters the Fredholm alternative:

either the problem (20) is well-posed, ie. it admits a unique solution e0 in H0(curl;Ω), which depends continuously on the dataf:

ke0kH(curl;Ω).kfk0,Ω;

or, the problem (20) has solutions if, and only if, f satisfy a finite num- ber of compatibility conditions; in this case, the space of solutions is an affine space of finite dimension, and the component of the solution which is orthogonal (in the sense of theH0(curl;Ω)inner product) to the corre- sponding linear vector space, depends continuously on the data f.

Finally, each alternative occurs simultaneously for variational formulation (20) and variational formulation (3) with unknowne.

From now on, we assume that variational formulation (3) iswell-posed:

∀f ∈L2(Ω), ∃!e∈H0(curl;Ω) solnto (3) andkekH(curl;Ω).kfk0,Ω. (22)

(10)

4 Approximation by N´ed´elec’s finite elements

For the ease of exposition1, we assume thatand{Ωp}p=1,···,P are Lipschitz polyhedra. We consider a family of simplicial meshes ofΩ, and we choose the N´ed´elec’s first family of edge finite elements [41, 40] to define finite dimensional subspaces (Vh)h of H0(curl;Ω). So is triangulated by a shape regular family of meshes (Th)h, made up of (closed) simplices, generically denoted by K. Each mesh is indexed by h := maxKhK (the meshsize), where hK is the diameter ofK. And meshes are conforming with respect to the partition {Ωp}p=1,···,P induced by the coefficients ε, µ: namely, for all h and all K ∈ Th, there exists p ∈ {1,· · ·, P} such that Kp. N´ed´elec’s H(curl;Ω)- conforming (first family, first-order) finite element spaces are then defined by

Vh:={vhH0(curl;Ω) : vh|K ∈ R1(K), ∀K∈ Th}, whereR1(K) is the vector space of polynomials onK defined by

R1(K) :={v ∈P1(K) : v(x) =a+b×x, a,bR3}.

To approximate the curl-free fields, we need to define a suitable approximation of elements ofH01(Ω). So we introduce finite dimensional subspaces (Mh)h of H01(Ω). Lagrange’s first-order finite element spaces are defined by

Mh:={qhH01(Ω) : qh|KP1(K), ∀K∈ Th}.

The discrete companion scalar problems are Find shMh such that

(ε∇sh|∇qh)0,Ω=hg, qhiH1

0(Ω), ∀qhMh. (23)

For approximation purposes, one can use the Lagrange interpolation operator ΠhL, or the Scott-Zhang interpolation operator ΠhSZ. The latter allows one to interpolate any element of H01(Ω), with values in Mh, at the expense of local interpolation operators that are not localized to each tetraedron, but are localized to the union of the tetrahedron and its neighbouring tetrahedra. We refer to [30] for details. Unless otherwise specified, we chooseΠhgrad=ΠhSZ.

Forhgiven, the discrete variational formulation of the time-harmonic prob- lem (3) is

Find ehVh such that

aω(eh,vh) = (f|vh)0,Ω, ∀vhVh. (24)

1 The results obtained in this paper carry over to curved polyhedra, that is domains with piecewise smooth boundaries (see eg. p. 81 in [2] for a precise definition). When dealing with the discretization by first-order edge finite elements inH0(curl;Ω), one may use [26].

Respectively, when dealing with the discretization by Lagrange’s first-order finite elements inH01(Ω), one may use [32]. In particular, it is proven there that optimal interpolation properties hold, ie. one may recover up toO(h) accuracy, provided the field to be interpolated is sufficiently smooth.

(11)

To obtain explicit error estimates for the time-harmonic Maxwell equations, a natural idea is to use the interpolation of its solutione. This requires some additional definitions and a priori analysis of the regularity ofe, and of its curl.

Let Πhcurl be the classical global Raviart-Thomas-N´ed´elec interpolant in H0(curl;Ω) with values in Vh [41]. We then denote by Πhdiv the classi- cal global Raviart-Thomas-N´ed´elec interpolation operator inH0(div;Ω) with values inWh [45, 41], where (Wh)hare designed with the help ofH(div;Ω)- conforming, first-order finite element spaces:

Wh:={whH0(div;Ω) : wh|K∈ D1(K), ∀K∈ Th}, whereD1(K) is the vector space of polynomials onK defined by

D1(K) :={v∈P1(K) : v(x) =a+bx, aR3, bR}.

Let us recall a few useful properties (see Chapter 5 in [40]). To start with, Proposition 3 For all h, it holds that

∇[Mh]⊂Vh; (25)

∀vhVh, Πhcurlvh=vh; (26)

curl[Vh]⊂Wh; (27)

∀whWh, Πhdivwh=wh; (28)

∀v∈H0(curl;Ω)s.t.Πhcurlv exists, Πhdiv(curlv) =curl(Πhcurlv).(29) There are useful additional properties regarding Πhcurl listed below. Below, when we refer to piecewise-Hs fields, the partition is understood as in (12).

Proposition 4 (discrete exact sequence [41]) Let h be given, and let vH0(curl;Ω) that can be written asv =∇q in Ω, for some qH01(Ω).

Then if Πhcurlv is well-defined, there exists qhMh such that Πhcurlv=∇qh

inΩ.

Proposition 5 (classical interpolation results)Assume thatvP Hs(Ω) andcurlvP Hs0(Ω)for somes>1/2,s0 >0. Then one can defineΠhcurlv and, in addition, one has the approximation result [4]:

kv−ΠhcurlvkH(curl;Ω).hmin(s,s0,1){kvkP Hs(Ω)+kcurlvkP Hs0

(Ω)}. (30) Furthermore, if curlv is piecewise constant onTh, one has the improved ap- proximation result (cf. Theorem 5.41 in [40]):

kv−ΠhcurlvkH(curl;Ω).hmin(s,1){kvkP Hs(Ω)+kcurlvk0,Ω}.

(12)

Remark 2 When2 is a domain ofR2, note that one can define the Raviart- Thomas-N´ed´elec interpolant of a fieldvH(curl;2) as soon asvP Hs(Ω2) for somes>0 (there is no requirement on the regularity of curlv). This re- sult is proven in [3] for fields in H(div;2), and it obviously carries over to fields in H(curl;2) by appropriate coordinates transform. Further, one has the approximation result:

kv−ΠhcurlvkH(curl;Ω2).hmin(s,1){kvkP Hs(Ω2)+kcurlvk0,Ω2}. ut Our aim is to applyΠhcurl to the electric fieldegoverned by (3).

First, one must havecurleP Hs0(Ω) for somes0 >0. SincefL2(Ω), we immediately find thatc=µ−1curlebelongs to

XT(Ω, µ) :={v∈H(curl;Ω) : µvH0(div;Ω)}.

Then, using a shift theorem for the companion scalar problem with Neumann boundary condition

FindsHzmv1 (Ω)such that (µ∇s|∇q)0,Ω =hg0, qiH1

zmv(Ω), ∀q∈Hzmv1 (Ω), (31) and a regular plus gradient splitting (see eg. [24, 21]), we introduceτN eu∈(0,1]

depending only on the geometry and onµsuch that XT(Ω, µ)⊂ ∩s0∈[0,τN eu)P Hs0(Ω),

with continuous imbedding for alls0∈[0, τN eu). Furthermore, using a Weber inequality (cf. Theorem 6.2.5 in [2]), one has that for alls0 ∈[0, τN eu),

∀v∈XT(Ω, µ), kvkP Hs0

(Ω).kcurlvk0,Ω+kdivµvk0,Ω+ X

1≤i≤I

|hµv·n,1iH1/2i)|. (32) As a consequence, we note that sinceµis piecewise smooth, it also holds that curle∈ ∩s0∈[0,τN eu)P Hs0(Ω). (33) Then, to guarantee thatΠhcurlcan be applied to the electric fielde, one must check whethereP Hs(Ω) for some s>1/2. To evaluate the exponentsa priori, we use the following decomposition (see Lemma 2.4 of [36]).

Proposition 6 There exist operators

P ∈ L(H0(curl;Ω),H1(Ω)), Q∈ L(H0(curl;Ω), H01(Ω)), such that

∀v∈H0(curl;Ω), v=P v+∇(Qv). (34) This yields some useful results for elements ofKN(Ω, ε).

(13)

Corollary 2 The a priori regularity of elements ofKN(Ω, ε) is governed by the imbedding:

KN(Ω, ε)⊂ ∩s∈[0,τDir)P Hs(Ω).

Moreover, for alls∈[0, τDir),

∀k∈KN(Ω, ε), kkkP Hs(Ω).kcurlkk0,Ω. (35) Proof Let kKN(Ω, ε). According to proposition 6, one can write k = k?+∇sk with k?H1(Ω), resp. skH01(Ω), and it holds that kk?k1,Ω+ kskkH1

0(Ω) .kkkH(curl;Ω). In particular, div(ε∇sk) =−divεk? in Ω, so sk solves the modified scalar problem (14) with data (0,−k?). Thanks to the shift theorem, we know that, for all s ∈ [0, τDir),sk belongs to P H1+s(Ω), with the boundkskkP H1+s(Ω).kk?k1,Ω.Using the triangle inequality, we conclude that

∀s∈[0, τDir), kP Hs(Ω), andkkkP Hs(Ω).kkkH(curl;Ω).

This proves the first part of the corollary. Using finally theorem 1 on the equivalence of norms inKN(Ω, ε), we conclude that (35) holds. ut Hence, it may happen that the field to be interpolated, eg. the electric field e, does not belong tos>1/2P Hs(Ω). In the classical case, the occurence of such a situation is explained in Section 7 of [24]. In the interface case, this can be inferred from the results obtained in [8, 11].

On the other hand, to interpolate such a low regularity field, one may still choose the quasi-interpolation operator of [31], or the combined interpolation operator of [21, 22]. We choose the latter. To get a definition for the combined interpolation operator, denoted by Πhcomb, one needs to be able to split low regularity fields defined onΩ. To that end, we apply proposition 6.

Definition 1 (combined interpolation operator) LetvH0(curl;Ω), withcurlvHs0(Ω) for somes0 >0. We define

Πhcombv:=Πhcurl(P v) +∇(Πhgrad(Qv)).

Then, the approximation results for the combined interpolation are a straight- forward consequence of the available results forΠhcurl andΠhgrad.

Proposition 7 (combined interpolation results) Let vH0(curl;Ω), with QvP H1+s(Ω) and curlvP Hs0(Ω) for some s≥0,s0 >0. One has the approximation result:

kv−ΠhcombvkH(curl;Ω).hmin(s,s0,1){kvkH(curl;Ω) (36) +kQvkP H1+s(Ω)+kcurlvkP Hs0

(Ω)}.

Furthermore, if curlv is piecewise constant onTh, one has the improved ap- proximation result:

kv−ΠhcombvkH(curl;Ω).hmin(s,1){kvkH(curl;Ω)+kQvkP H1+s(Ω)}.

(14)

Together with this definition of the combined interpolation operator, we have the results below, to be compared with the well-known results (26) and (29) for the classical interpolation operator.

Proposition 8 For all h, it holds that

∀vhVh,∃qhMh, Πhcombvh=vh+∇qh; (37)

∀v∈H0(curl;Ω) s.t.curlvHs0(Ω) for somes0 >0, (38) Πhdiv(curlv) =curl(Πhcombv).

Proof LetvhVh. We note that becausevh is piecewise smooth onTh, one hasvh,curlvhP Ht(Ω) for allt∈[0,1/2). HenceΠhcombvhis well-defined according to definition 1. If we write vh= (vh)?+∇svh, with (vh)? =P vh, resp.svh =Qvh, we haveΠhcombvh:=Πhcurl(vh)?+∇(Πhgradsvh).

On the other hand,∇svh =vh−(vh)?. SinceΠhcurl(vh−(vh)?) is well-defined, so isΠhcurl(∇svh) and, according to proposition 4, there exists qh0Mhsuch thatΠhcurl(∇svh) =∇q0h. Applying nowΠhcurlto (vh)?=vh−∇svh, it follows that

Πhcurl(vh)?=Πhcurlvh− ∇q0h=vh− ∇q0h,

where the second equality now follows from (26). One concludes that Πhcombvh:=vh+∇(Πhgradsvhqh0),

which is precisely (37) withqh=Πhgradsvhq0hMh.

To check (38), letvbe split asv=P v+∇(Qv). Sincecurl(P v)Hs0(Ω), ac- cording to proposition 5 one may apply (29) toP v, leading toΠhdiv(curl(P v)) = curl(Πhcurl(P v)). On the other hand, because of the definition 1 ofΠhcombv= Πhcurl(P v) +∇(Πhgrad(Qv)) one has

curl(Πhcurl(P v)) =curl(Πhcombv− ∇(Πhgrad(Qv))) =curl(Πhcombv).

Using finally the equalitycurlv =curl(P v) leads to the claim. ut We now have all the required results to bound the interpolation error of the electric fielde.

Proposition 9 Letebe the solution to the time-harmonic Maxwell equations.

Let the extra-regularity of the data f be as in (13) with τ0 > 0 given. One can define Πhcombe, and moreover one has the approximation result, for all s∈[0,min(τ0, τDir, τN eu)),

ke−ΠhcombekH(curl;Ω).hs{kdivfk−1+s,Ω+kfk0,Ω}.

Proof Lets∈[0,min(τ0, τDir, τN eu)) ; becauseτDir≤1, one hass<1.

According to proposition 6, we may writee=e?+∇sewithe?H1(Ω),seH01(Ω), andke?k1,Ω+ksekH1

0(Ω).kekH(curl;Ω). By construction,sesolves the modified scalar problem (14) with data (divf,−e?). Buts<min(τ0, τDir) so,

(15)

thanks to the shift theorem,seP H1+s(Ω), with the boundksekP H1+s(Ω). kdivfk−1+s,Ω+ke?k1,Ω. Using (22) for the last inequality below, we find

ksekP H1+s(Ω).kdivfk−1+s,Ω+kekH(curl;Ω).kdivfk−1+s,Ω+kfk0,Ω. On the other hand, one hascurleP Hs(Ω), cf. (33). Then, with the help of the bound (32) on kµ−1curlekP Hs(Ω), noting that divcurle= 0 inΩ, and hcurle·n,1iH1/2i)= 0 for 1≤iI (see Remark 3.5.2 in [2]), we find

kcurlekP Hs(Ω).kµ−1curlekP Hs(Ω).kcurl(µ−1curle)k0,Ω .kek0,Ω+kfk0,Ω,

where we used the relation curl(µ−1curle) =ω2εe+ω2f in Ω. Therefore, we one can defineΠhcombe and, using (36) and (22) once more, one finds now ke−ΠhcombekH(curl;Ω).hs{kekH(curl;Ω)+ksekP H1+s(Ω)+kcurlekP Hs(Ω)}

.hs{kdivfk−1+s,Ω+kfk0,Ω},

which is the desired estimate. ut

Remark 3 In particular, we note that even when the electric fieldedoes not belong to∪s>1/2P Hs(Ω), one may use the combined interpolation operator and still obtain “best” interpolation error. On the other hand, it is well-known by using classical interpolation that, whenτDir=τN eu= 1, and for aregular datafH(div;Ω), the interpolation error behaves likeO(h). ut From this point on, to obtain the well-posedness result for the discretized problems, and finally convergence to the exact solutione, one needs to prove a uniform discrete inf-sup condition. For that, we mimic in section 5 the two in- gredients that were used to solve the exact variational formulation: uniformly stable discrete decompositions in the spirit of proposition 1 ; uniform equiva- lence of norms in the spirit of theorem 1. The key ingredient is the study of the approximation of the companion scalar problem (5). And, since it was origi- nally solved with the help of T-coercivity, we consider two situations regarding its approximation. We refer to the appendix A for details. Either we have at hand a ”full” T-coercivity involution operator T0 to solve (5), that can also be used to establish to establish the uniform discrete T-coercivity (58)-(59) of the discrete scalar problems (23). Or, we only have at hand a ”weak” explicit T-coercivity involution operatorT, cf. (57). The first situation is addressed in subsection 5.1, whereas the second situation is addressed in subsection 5.2.

5 Uniform estimates

5.1 Case of a ”full” T-coercivity operator

We assume in this section that we have at hand a ”full” T-coercivity involu- tion operatorT0to solve the companion scalar problem (5) (see section A.1),

(16)

and that the meshes are T-conform, such that (58)-(59) are fulfilled, with consequences listed in section A.2. Define, for anyh,

Kh(ε) :={vhVh : (εvh|∇qh)0,Ω= 0, ∀qhMh}. (39) Proposition 10 Assume that (58) holds. For allh, one has the direct sum

Vh=∇[Mh]⊕Kh(ε). (40) Proof Let h be given. Thanks to (25), we know that ∇[Mh] + Kh(ε) is a subset ofVh. Then forvhVhand because the discrete scalar problem (23) is well-posed, there exists one, and only one,pvhMh such that

(ε∇pvh|∇qh)0,Ω= (εvh|∇qh)0,Ω, ∀qhMh. (41) And one has

kvh =vh− ∇pvhKh(ε), (42) soVh=∇[Mh] +Kh(ε). Using (58), the fact that the sum is direct is derived exactly as in the continuous case (see the proof of proposition 1). ut For all h, we can use the splitting (40) and the explicit definitions (41)-(42) to introduce the operators

π1h :

VhMh

vh7→pvh

, π2h :

VhKh(ε) vh7→kvh

. (43)

In other words, one may write, for all h, for all vhVh, vh =∇(π1hvh) + π2hvh. Also, one has for all h, (π2h)2 =π2h. Below, we prove the uniform stability of the decomposition (40).

Proposition 11 Assume that (58) holds. The continuity moduli of the oper- ators1h)h,2h)h are bounded independently ofh.

Proof Given handvhVh, one has according to (58) and (41)

α00k∇(π1huh)k20,Ω≤(ε∇(π1huh)|∇(T01huh)))0,Ω= (εuh|∇(T01huh)))0,Ω

≤ kεuhk0,Ωk∇(T01huh))k0,Ω

≤ kεuhk0,ΩkT0kL(H1

0(Ω))k∇(π1huh)k0,Ω, so that

k∇(π1huh)k0,Ω≤(α00)−1kεk∞,ΩkT0kL(H1

0(Ω))kuhk0,Ω. And then

2huhkH(curl;Ω)≤(1 + (α00)−1kεk∞,ΩkT0kL(H1

0(Ω)))kuhkH(curl;Ω),

so the claim follows. ut

Next, one has to check that kh 7→ kcurlkhk0,Ω defines a norm on Kh(ε).

And, if the answer is positive, whether this norm of uniformly equivalent in h(ie. with constants that are independent ofh) to thek · kH(curl;Ω)-norm on Kh(ε).

Références

Documents relatifs

Maxwell’s equations, interface problem, edge elements, sensitivity to coefficients, 18.. error

In this paper we considered the edge finite element approximation of the Maxwell resolvent operator and we proved the uniform convergence in the L 2 (Ω) 3 -norm.. The proof

Among those stands the adiabatic variable method; we present in this paper a mathematical analysis of this approximation and propose, in particular, an a posteriori estimate that

Our proof clearly implies that the discrete compactness property holds true also for the pure p method on a conforming rectangular mesh (i.e., without hanging node) with the

It is known that the discrete and semi-discrete models obtained by dis- cretizing the wave equation with the classical finite difference or finite element method do not

[2004, 2006, 2008, etc.] Simulate an “array” of n chains in “parallel.” At each step, use an RQMC point set Pn to advance all the chains by one step, while inducing global

Keywords: Dynamical systems; Diophantine approximation; holomorphic dynam- ics; Thue–Siegel–Roth Theorem; Schmidt’s Subspace Theorem; S-unit equations;.. linear recurrence

Dur- ing MGMT amplification, five DNAs (cases n°7, 19, 26, 37, 38) showed uncertain or failed controls of conver- sion but valid controls of conversion for LINE-1 and MLH1