• 検索結果がありません。

The basis functions exploit an orthogonal decomposition of the trial subspace to minimize the energy and are expressed in terms of local eigenproblems

N/A
N/A
Protected

Academic year: 2022

シェア "The basis functions exploit an orthogonal decomposition of the trial subspace to minimize the energy and are expressed in terms of local eigenproblems"

Copied!
24
0
0

読み込み中.... (全文を見る)

全文

(1)

ERROR ESTIMATES FOR A TWO-DIMENSIONAL SPECIAL FINITE ELEMENT METHOD BASED ON COMPONENT MODE SYNTHESIS

ULRICH HETMANIUKANDAXEL KLAWONN

Abstract. This paper presents a priori error estimates for a special finite element discretization based on compo- nent mode synthesis. The basis functions exploit an orthogonal decomposition of the trial subspace to minimize the energy and are expressed in terms of local eigenproblems. The a priori error bounds state the explicit dependency of constants with respect to the mesh size and the first neglected eigenvalues. A residual-based a posteriori error indicator is derived. Numerical experiments on academic problems illustrate the sharpness of these bounds.

Key words. domain decomposition, finite elements, eigendecomposition, a posteriori error estimation AMS subject classifications. 35J20, 65F15, 65N25, 65N30, 65N55

1. Introduction. Classical Lagrangian finite element methods are challenged by prob- lems

−∇ ·(A(x)∇u(x)) =f(x) inΩ, u= 0 on∂Ω, (1.1)

where the coefficient matrixAis rough or highly oscillating so that a standard application of the finite element method needs a highly refined mesh to reach sufficient accuracy. Over the last couple of years, many discretization methods have been proposed to enable the ac- curate, efficient, and robust solution of these complex problems. Approximation subspaces that incorporate specialized knowledge of the coefficient matrixAgive rise to effective fi- nite element methods. Examples include the multiscale finite element [15,21], the mixed multiscale finite element [1], the heterogeneous multiscale finite element [14], adaptive mul- tiscale methods [28], and the generalized finite element method [3,4,6]. Babuˇska, Caloz, and Osborn [5, p. 947] denote such finite element methods special.

Hetmaniuk and Lehoucq [20] proposed to build a conforming approximation space by local eigenfunctions for the partial differential operator in (1.1). Eigenbases are often efficient in terms of Kolmogorov n-width (see Melenk [26]), and local eigenfunctions are supposed to span a good approximation space. The discretization in [20] is based upon the classic idea of component mode synthesis (CMS), introduced in [13,23] and used, e.g., by Gervasio et al. [16] in the spectral projection decomposition method. Starting from a partition of the do- mainΩ, component mode synthesis methods exploit an orthogonal decomposition ofH01(Ω) to solve the minimization problem

(1.2) argmin

v∈H01(Ω)

1 2 Z

(∇v(x))TA(x)∇v(x)dx− Z

f(x)v(x)dx

.

For two-dimensional problems, the conforming approximation space proposed in [20] com- bines bubble eigenfunctions (localized inside one element), energy-minimizing extensions of vertex-specific trace functions (localized on the elements sharing the vertex), and energy- minimizing extensions of edge-bubble eigenfunctions (localized on an edge and the adja- cent elements). Numerical experiments in [20,24] illustrate the efficacy of this CMS-based

Received August 17, 2013. Accepted March 28, 2014. Published online on June 20, 2014. Recommended by Y. Achdou. The work is partly supported by NSF grant DMS-0914876.

Department of Applied Mathematics, University of Washington, Box 353925, Seattle, WA 98195 (hetmaniu@uw.edu).

Mathematisches Institut, Universit¨at zu K¨oln, Weyertal 86-90, 50931 K¨oln, Germany (klawonn@math.uni-koeln.de).

109

(2)

approach. The first goal of this paper is to present a priori error estimates for this local eigenfunction-based discretization. The error bounds state the explicit dependency of con- stants with respect to the mesh size and the first neglected eigenvalues.

Special finite element methods allow great flexibility in their definition. These numerical methods often contain a parameter, such as the vertex-specific trace function or the number of eigenfunctions, motivated by heuristics arguments. An efficient choice of parameter(s) may not be known in advance and could be estimated adaptively during the computations. The second objective of this paper is to derive an a posteriori error indicator that could guide the selection of the number of bubble eigenfunctions and edge-bubble eigenfunctions.

The rest of the paper is organized as follows. Section2reviews notations and the local eigenfunction-based discretization. Section3presents a priori error estimates and a residual- based a posteriori error indicator. Finally, numerical experiments illustrate the sharpness of these bounds.

2. Review of a special finite element method based on component mode synthesis.

LetΩbe a bounded polygonal domain in the planeR2whose boundary∂Ωis composed of straight lines. On this domain, the Sobolev spacesHk(Ω)andH0k(Ω)are defined in a stan- dard way (withk >0). Fractional order Sobolev spacesHs(Ω)are defined by interpolation.

Denote

a(u, v) = Z

(∇u(x))TA(x)∇v(x)dx ∀u, v∈H01(Ω),

the bilinear form induced by (1.1). The coefficient matrix Ais assumed to be symmetric positive definite, to beC1onΩ, and to satisfy

(2.1) 0< αminξTξ≤ξTA(x)ξ≤αmaxξTξ ∀x∈Ωandξ∈R2\ {0}. Givenf ∈L2(Ω), the problem (1.2) is rewritten as

argmin

v∈H01(Ω)

1

2a(v, v)−(f, v)

,

where(·,·)denotes the standard inner product onL2(Ω). The associated optimality system is the variational formulation of (1.1): findu∈H01(Ω)such that

(2.2) a(u, v) = (f, v) ∀v∈H01(Ω).

We refer to the solutions of (1.1), (1.2), and (2.2) as equivalent in a formal sense. Throughout the paper, the regularity assumption is:

ASSUMPTION1. Givenf ∈L2(Ω), there existss0> 32such that the solutionubelongs toHs0(Ω)∩H01(Ω).

This regularity assumption implies some conditions for the domainΩ. For example, whenΩ is convex, Assumption1holds withs0= 2; see Grisvard [17, Theorem 3.2.1.2].

Consider a family(Th)hof conforming partitions ofΩinto a finite number of triangles or convex quadrilaterals with straight edges. The mesh size his the maximal diameter of the elementsK inTh. Here every elementK is assumed to be a non-empty bounded open set. The family(Th)h is assumed to be shape regular, i.e., the ratio of the diameter of any element K inTh to the diameter of its largest inscribed ball is bounded by a constantσ independent ofKand ofTh. The interfaceΓis defined as

Γ = [

K∈Th

∂K

!

\∂Ω.

(3)

Given two distinct elementsKandKinTh, the intersectionK∩Kis empty, a vertex, or a complete edge with two vertices.

LetVKbe the subspace of local functions whose restrictions toKbelong toH01(K)and which are trivially extended throughoutΩ,

VK =n

v∈H01(Ω) :v|K ∈H01(K) and v|\K = 0o .

Any member function ofVK has a zero trace on the boundary∂Ωand on the interfaceΓ.

LetWΓ be the subspace of trace functions onΓfor all functions inH01(Ω). DenoteVΓ the subspace of energy-minimizing extensions of trace functions onΓ,

VΓ =

Eτ ∈H01(Ω) :τ ∈WΓ , where the extensionE(τ)solves the minimization problem

v∈Hinf01(Ω)a(v, v) subject to v|Γ=τ.

The energy-minimizing extensionE(τ)satisfies, in the weak sense,

−∇ ·(A(x)∇Eτ(x)) = 0 inK,∀K∈ Th, Eτ=τ onΓ,

Eτ= 0 on∂Ω.

(2.3)

This property indicates that functions inVΓare governed by the underlying partial differential equation. Note that any non-zero member function ofVΓhas a non-zero trace onΓ.

A key result is the orthogonal decomposition

(2.4) H01(Ω) = M

K∈Th

VK

!

⊕VΓ.

The decomposition (2.4) is orthogonal with respect to the inner producta(·,·)because a(v, w) = 0 ∀v∈VK,∀w∈VK,(K6=K),

a(v, vΓ) = 0 ∀v∈VK,∀vΓ∈VΓ.

The former equality follows because the supports of the two functionsvandware disjoint.

The latter equality follows by definition of the extension (2.3). Although not often stated in this form, result (2.4) is at the heart of the analysis and development of domain decomposition methods for elliptic partial differential equations [16,29,31] and modern component mode synthesis methods [7,10].

An approximating subspace consistent with the decomposition (2.4) arises from selecting basis functions in the subspacesVK andVΓ. To build this approximating subspace, we in- troduce two different sets of eigenvalue problems. First, we define fixed-interface eigenvalue problems: find(z,K, λ,K)∈VK×Rsuch that

a(z∗,K, v) =λ∗,K(z∗,K, v) ∀v∈VK.

Next, for any open edge e ⊂ Γ, the edge-based coupling eigenvalue problem is: find (τ,e, λ,e)∈H

1 2

00(e)×Rsuch that

a(E(˜τ,e), E(˜η)) =λ,e

Z

e

τ,eηde ∀η∈H

1 2

00(e),

(4)

FIG. 2.1. Example of an edge-bubble eigenfunction along an interior edgee.

FIG. 2.2. Trace ofϕPalongΓfor a domain partitioned into 16 elements.

whereη˜denotes the trivial extension ofηby 0 onΓ. The eigenvalues{λi,K}i=1and{λi,e}i=1

are assumed to be ordered into nondecreasing sequences. The eigenmodesz∗,Kandτ∗,eform orthonormal bases for theL2-inner product on the elementKand the edgee, respectively.

Figure2.1illustrates an example for an eigenfunctionτe.

To complete the approximating subspace, each vertex-specific functionϕP is defined as the harmonic extension satisfying

−∇ ·(A(x)∇ϕP(x)) = 0 inK, ϕP = 0 on∂Ω, ϕP 6= 0 onΓ, ϕP(P) =δP,P,

for any element K, where δP,P is the Kronecker delta function. Here ϕP is chosen to be linear on each edgee1. OnΓ, the trace forϕP has local support along the boundaries of elements sharing the vertexP. The resulting functionϕP will also have as support the elements sharing the pointP. Figure2.2illustrates an example of the trace ofϕP.

1Efendiev and Hou [15] discuss other choices forϕP.

(5)

The conforming discretization spaceVACMS, proposed in [20], is consistent with the or- thogonal decomposition (2.4) and is defined as follows:

VACMS = M

K∈Th

span{zi,K; 1≤i < IK}

!

"

M

P∈

span{ϕP}

!

⊕ M

e⊂Γ

span{E(˜τi,e) ; 1≤i < Ie}

!#

,

whereIKandIeare positive integers2. The letter A in ACMS stands for approximate. Note that the verticesP and the edgeseare taken in the interior ofΩ. The basis functions have local support and the homogeneous Dirichlet boundary condition is built intoVACMS.

In summary, the conforming finite-dimensional subspaceVACMS ⊂H01(Ω)exploits the orthogonal decomposition (2.4) for incorporating information from the variational forma(·,·).

The subspaceVACMScontains information within elements via the bubble eigenfunctions. The functionsϕP andE(˜τi,e)carry information among several and two elements, respectively.

3. Error estimates. The goal of this section is to derive error estimates for the differ- ence of the exact solutionuof (2.2) and the approximate solutionuACMS ∈VACMSdefined by

(3.1) a(uACMS, v) = (f, v) ∀v∈VACMS. The orthogonal decomposition (2.4) implies that

(3.2) a(u−uACMS, u−uACMS) =a(uB−uACMS,B, uB−uACMS,B)

+a(uΓ−uACMS,Γ, uΓ−uACMS,Γ), where the solutionusatisfies

u=uB+uΓ, uB∈ M

K∈Th

VK

!

and uΓ ∈VΓ, and the approximationuACMS∈VACMSis written as

uACMS=uACMS,B+uACMS,Γ, uACMS,B∈ M

K∈Th

VK

!

and uACMS,Γ∈VΓ. The two error terms in (3.2) are treated separately.

LEMMA3.1. The componentsuBanduACMS,Bsatisfy

(3.3) a(uB−uACMS,B, uB−uACMS,B)≤ X

K∈Th

kfk2L2(K)

λIK,K ≤Ch2 X

K∈Th

kfk2L2(K)

αmin,KIK,

whereCis a constant andαmin,Kverifies

(3.4) 0< αmin,KξTξ≤ξTA(x)ξ ∀x∈Kandξ∈R2\ {0}.

2WhenIKis 1, the subspacespan

zi,K; 1i < IK is equal to{0}(the same convention holds forIe).

(6)

Proof. By Galerkin orthogonality, the error satisfies

a(uB−uACMS,B, uB−uACMS,B)≤a(uB−w, uB−w)

∀w∈ M

K∈Th

span{zi,K; 1≤i < IK}

! .

For every elementK, define the projection operatorPIKas follows (3.5) ∀v∈L2(K) : PIK(v) =

IXK1 i=1

Z

K

zi,Kv

zi,K.

ReplacingwbyPIK(uB), the projection error foruBverifies a(uB− PIK(uB), uB− PIK(uB))

= X

K∈Th

Z

K

(∇uB− ∇PIK(uB))TA(∇uB− ∇PIK(uB)). On elementK, properties of the family of eigenfunctions(zi,K)+i=1indicate that

Z

K

(∇uB− ∇PIK(uB))TA(∇uB− ∇PIK(uB)) =

+

X

i=IK

λi,K Z

K

uBzi,K 2

.

For every eigenvectorzi,K, we have Z

K

uBzi,K= 1 λi,K

Z

K

(∇uB)TA∇zi,K= 1 λi,K

Z

K

(−∇ ·(A∇uB))zi,K

= 1

λi,K Z

K

f zi,K. Hence, we get

Z

K

(∇uB− ∇PIK(uB))TA(∇uB− ∇PIK(uB))

=

+

X

i=IK

1 λi,K

Z

K

f zi,K 2

≤kf− PIK(f)k2L2(K)

λIK,K ≤kfk2L2(K)

λIK,K . Thus, the projection erroruB− PIK(uB)satisfies

a(uB− PIK(uB), uB− PIK(uB)) = X

K∈Th

kfk2L2(K)

λIK,K

.

By (3.4), the eigenvalueλi,Kis larger thanαmin,Ktimes thei-th eigenvalue of the Lapla- cian onK. By combining the bound of Bourquin on eigenvalues for the Laplacian [9, p. 74]

and the shape regularity of the family(Th)h, there exists a constantCindependent ofKandi such that

(3.6) λi,K≥Cαmin,K

i h2. This estimate concludes the proof.

(7)

This result uses only the regularity assumption that−∇ ·(A∇u) =fbelongs toL2(Ω).

Whenfis more regular, a sharper bound for the projection error exists. The lower bound (3.6) is valid for all eigenvalues, while Weyl’s formula for eigenvalues is asymptotic; see Bour- quin [10, Equation (95)].

Next, the error inVΓis estimated.

LEMMA3.2. The componentsuΓanduACMS,Γsatisfy (3.7) a(uΓ−uACMS,Γ, uΓ−uACMS,Γ)≤Cs0,σ,Ah2s03 X

K∈Th

kuk2Hs0(K)

mine⊂∂K∩ΓλIe,e, when the solutionubelongs toHs0(Ω)∩H01(Ω).

Proof. By Galerkin orthogonality, the error satisfies

a(uΓ−uACMS,Γ, uΓ−uACMS,Γ)≤a(uΓ−w, uΓ−w) ∀w∈VΓ.

Recall that the functionuΓis equal toE(u|Γ). Using the same characterization forwyields a(uΓ−w, uΓ−w) = X

K∈Th

Z

K

(∇E(u|Γ−w|Γ))TA∇E(u|Γ−w|Γ).

When the restrictionu|e−w|ebelongs toH

1 2

00(e)for every edgee⊂Γ, we have onK E(u|Γ−w|Γ) =E(u|∂K−w|∂K) = X

e∂K

E

u|^e−w|e

,

whereu|^e−w|eis the trivial extension of(u−w)|eby 0 onΓ. This relation yields a(uΓ−w, uΓ−w)

= X

K∈Th

Z

K

X

e∂K

∇E

u|^e−w|e

!T

A X

e∂K

∇E

u|^e−w|e

!

and

a(uΓ−w, uΓ−w)

≤C X

K∈Th

X

e∂K

Z

K

∇E

u|^e−w|e

T A

∇E

u|^e−w|e

, (3.8)

where the Cauchy-Schwarz inequality has been used. The support ofu|^e−w|eis included ine. Its energy-minimizing extension has a local support inKe,1∪Ke,2, whereKe,1andKe,2

are the two elements whose boundaries share the edgee. Rearranging the terms in (3.8) gives a(uΓ−w, uΓ−w)

≤CX

eΓ

Z

Ke,1Ke,1

∇E

u|^e−w|e

T A

∇E

u|^e−w|e

. (3.9)

To construct such a functionw, we proceed as follows. LetIh be the piecewise linear interpolation operator onΓand define the projection operatorΠIe, for each interior edgee, as follows

∀η∈L2(e) : ΠIe(η) =

IXe1 i=1

Z

e

τi,eη

τi,e.

(8)

We replace the functionwby

w=E Ih(uΓ) +X

eΓ

ΠeIe(uΓ− Ih(uΓ))

!

∈ VΓ∩VACMS,

whereΠeIe(η)is the extension by 0 ofΠIe(η)onΓ. For this choice ofw, we have u|e−w|e∈H

1 2

00(e) for every edgee⊂Γ.

Assumption 1 indicates that E(u|Γ) belongs to Hs0(Ω). Hence, the restriction u|e− Ih(u)|eis contained in H

1 2

00(e)∩H1(e)for every edgee ⊂ Γ. The relations (3.9) and (A.2) yield

(3.10) a(uΓ−w, uΓ−w)≤Cs0,σ,A

X

eΓ

ku− Ih(u)k2H1(e)

λIe,e . Properties of the interpolation operatorIhgive

ku− Ih(u)k2H1(e)≤Ch2(s032)|u|2Hs0−1

2(e)≤Ch2(s032)kuk2Hs0−1 2(e); see Steinbach [30]. Relation (3.10) becomes

a(uΓ−w, uΓ−w)≤Cs0,σ,Ah2(s032) X

K∈Th

P

e∂Kkuk2Hs0−1 2(e)

mine∂KΓλIe,e

.

A theorem of Arnold et al. [2, Theorem 6.1] indicates that we have X

e∂K

kuk2Hs0−1

2(e)≤Ckuk2Hs0(K)

becauseuis continuous onΩand satisfies the conditions for traces on a polygon. Finally we get

a(uΓ−w, uΓ−w)≤Cs0,σ,Ah2s03 X

K∈Th

kuk2Hs0(K)

mine⊂∂K∩ΓλIe,e.

To the best of the authors’ knowledge, a lower bound on all the edge-bubble eigenval- uesλ,eis not available. Based on the discussion in Bourquin [9, p. 89] and on egde-related eigenvalues for particular geometries (see, for example, [10, p. 412]), one could expect that

(3.11) λl,e≥Cαmin

l h,

where the constantCdoes not depend oneorl. The error (3.7) would become (3.12) a(uΓ−uACMS,Γ, uΓ−uACMS,Γ)≤Cs0,σ,A

h2s02 αmin

X

K∈Th

kuk2Hs0(K)

mine⊂∂K∩ΓIe, providing a rate ofh2whenubelongs toH2(Ω).

(9)

REMARK3.3. The result in Lemma3.2does not exhibit an optimal behavior with respect to the edge-based coupling eigenvalues whens0>32. Indeed, bounds on the eigendecompo- sition do not take into account the smoothness ofu|ΓbeyondH1(Γ). Such analysis for the Steklov-Poincar´e operator seems difficult to establish.

Combining (3.2) and the previous two lemmas yields the error estimate foru.

PROPOSITION 3.4. Assume that the solutionuof (2.2) belongs toH01(Ω)∩Hs0(Ω), with s0 > 32. Then the error between the solution u and the approximate solution uACMS∈VACMSsatisfies

a(u−uACMS, u−uACMS)≤ X

K∈Th

kfk2L2(K)

λIK,K

+Cs0,σ,Ah2s03 X

K∈Th

kuk2Hs0(K)

mine∂KΓλIe,e

,

where the constantCs0,σ,Adoes not depend onuandh.

Note that the approximationuACMSconverges toueven without any bubble eigenfunction (i.e.,IK = 1). For every elementK, the first eigenvalueλ1,Kverifiesλ1,K ≥Cαhmin2 ,which yields

a(u−uACMS, u−uACMS)≤C h2

αminkfk2L2(Ω)

+Cs0,σ,Ah2s03 X

K∈Th

kuk2Hs0(K)

mine∂KΓλIe,e. WhenIK =Ie= 1, the approximationuACMSstill converges touthanks to the vertex-specific functions. This particular case was proved in [12,22].

REMARK3.5. The error estimates in Proposition3.4are closely related to the pioneering work of Bourquin [8,9,10] on component mode synthesis. The main difference lies in the way the information is transferred among elements. Bourquin uses eigenmodes on Γ for the Steklov-Poincar´e operator. Here the vertex-specific functionsϕP and the edge-bubble eigenfunctions carry information among elements and have local support.

The choice of basis functions in VACMS determine the efficiency of the discretization method. The number of eigenfunctions cannot be known in advance and should be estimated adaptively during the computations. The following proposition introduces an a posteriori error indicator that could guide how to select the number of bubble eigenfunctions and edge- bubble eigenfunctions.

PROPOSITION3.6. The error betweenuanduACMSsatisfies pa(u−uACMS, u−uACMS)≤Cε,σ,A

(X

K∈Th

kf− PIK(f)k2L2(K)

λIK,K

+h X

K∈Th

kf− PIK(f)k2L2(K)

X

e⊂∂K∩Γ

1 λ2Ie,e

!

+hX

eΓ

Je νTeA∇uACMS2

L2(e)

λ1Ie,e



1 2

, (3.13)

(10)

whereε >0andJe(ψ)denotes the jump of a given functionψacross the edgeein the direc- tion of the unit normal vectorνe. The constantCε,σ,Adepends onε,σ, and the coefficient matrixA.

The proof is given in AppendixB. Bound (3.13) indicates that the right hand side defines an a posteriori error indicator. This error indicator is reliable, i.e., the error is bounded from above by multiples of the indicator. Proving the effectivity of the error indicator remains an open question.

REMARK3.7. In practice, the basis functions are computed numerically by introducing a nested finer grid. The selection of this nested finer grid impacts both the accuracy and the complexity of the algorithm. Finding error estimates and a posteriori error indicators for such a two-grid scheme remains an open problem that is beyond the scope of this paper; see [11]

and [19] for a recent study applied to the multiscale finite element method. A complexity comparison between a two-grid scheme and the standard application of the finite element method would require a specific study with careful numerical experiments. However, to estimate the merit of the two-grid scheme over the standard application of the finite element method, flop count expressions are briefly discussed in the same style as the comparison of Hou and Wu [21, Section 4.2].

If Mh denotes the fine mesh size, then the fine grid yieldsO M2h2

degrees of free- dom. The computational complexity associated with the standard application of the finite element method over the fine grid is dominated by the operation count for solving the linear system,

O (M2h2)α

=O(Mh),

whereα∈[1,3]depends on the specific linear solver used3. The complexity for the two-grid scheme based on component mode synthesis is

O(h) + max

O(Mh2),O(M6h2),O(M2α+1h2) ,

where O(h) is the cost of solving the algebraic equation (3.1). The other term esti- mates the cost for computing the basis functionsϕP,z,K,andE(˜τ,e), respectively. The complexity for computing all the vertex-specific functionsϕP isO(Mh2). The bubble eigenfunctionsz,K require, at most,O(M6h2)operations. Note that this cost is an over- estimate because it does not exploit the fact that onlyIK ≪ M2 eigenmodes of a sparse pencil are needed. O(M2α+1h2)estimates the complexity for computing the edge-bubble eigenfunctionsE(˜τ,e).

Whenα = 1, the two-grid scheme is not attractive from an operation count point of view. However, solvers withα = 1are not common or available for a general coefficient matrixA. As soon as α > 1, a two-grid scheme has some merit, especially when M is smaller thanh13.

4. Numerical experiments. In this section, numerical experiments illustrate the sharp- ness of the previous bounds at academic examples. When the exact solution is not known explicitly, the energy,

E(v) = 1 2

Z

(∇v(x))TA(x)∇v(x)dx− Z

f(x)v(x)dx,

represents an intrinsic metric for comparing the quality of approximations to the exact solu- tion. Computing the difference between the energy of the computed solution and the energy

3For a finite element discretization in two dimensions, a sparse solver is usually characterized byα= 32; see, for example, Heath [18, Table 11.4].

(11)

100 101 102 103 104 10−4

10−3 10−2

Number of bubble eigenfunctions H1 semi−norm of absolute error

Discretization Error Line of Slope −1/2

FIG. 4.1. Convergence curve for solution (4.1) withα= 1.51andh= 1(squared domain).

of the exact solution,E = E(u), is equivalent to computing the norm of the error for the energy inner product,

E(v)− E=a(u−v, u−v)

2 .

The minimum energyEis obtained by extrapolating energies for finite element solutions on fine meshes.

4.1. Convergence towards a smooth solution. In this section, consider the problem

−∆u=f in Ω = [0,1]×[0,1], u= 0 on ∂Ω, where the domain[0,1]×[0,1]is partitioned by square elements.

First, the functionf is chosen so that the exact solution is

(4.1) u(x, y) = x−x2

y−y2α

,

whereα > 32. Figure4.1illustrates the convergence when only one element is used and the number of bubble eigenfunctions is increased. Whenα= 1.51≈32+ε, the right hand sidef belongs toL2(Ω). The convergence curve exhibits a decrease proportional to 1λ

I, which is predicted by the bound (3.3).

Whenf = 1, the right hand side now belongs toH12(Ω). In Figure4.2, the convergence, when only one element is used and the number of bubble eigenfunctions is increased, exhibits a higher convergence rate, which is described by the projection error off,

kf − PIfkL2(Ω)

√λI ≤kfkL2(Ω)

√λI .

Keepingf = 1and using only one bubble eigenfunction and one edge-bubble eigen- function, Figure4.3illustrates the convergence when the number of elements is increased.

As expected, the convergence curve exhibits a decrease proportional to the mesh sizeh.

The next study keepsf = 1and usesh= 12 and4096bubble eigenfunctions for every element. Figure4.4illustrates the convergence when the number of edge-bubble eigenfunc- tions is increased. The convergence curve exhibits a plateau because the number of bubble

(12)

100 101 102 103 10−3

10−2 10−1 100 101

Number of bubble eigenfunctions

Absolute error

L2 norm of projection error for f H1 semi−norm of discretization error Line of slope −1/4

Line of slope −3/4

FIG. 4.2. Convergence curve whenf= 1andh= 1(squared domain).

10−3 10−2 10−1 100

10−4 10−3 10−2 10−1

h

Absolute error

H1 semi−norm

FIG. 4.3. Convergence curve whenf= 1(squared domain).

100 101 102 103

10−6 10−5 10−4 10−3 10−2

Number of edge−bubble eigenfunctions

Absolute error

H1 semi−norm Line of slope −3/2

FIG. 4.4. Convergence curve whenf= 1,h= 12, and 4096 bubble eigenfunctions are used (squared domain).

(13)

10−4 10−3 10−2 10−1 100 10−4

10−3 10−2 10−1 100

Mesh size

Absolute error

H1 semi−norm A posteriori error indicator Line y = Ch2/3

FIG. 4.5. Convergence curve for a fixed number of bubble and edge-bubble eigenfunctions (L-shaped domain, f= 1).

eigenfunctions is fixed. Before reaching this asymptote, the curve decreases likeI

3 2

e . This rate is higher than the prediction in (3.12). Bourquin [9, p. 45] indicates that, for smooth functions, a superconvergence phenomenon is expected with the precise rateI

3 2

e . 4.2. Problem on a L-shaped domain. In this section, consider the problem

−∆u= 1 in Ω = ([0,1]×[0,1])\ 1

2,1

× 1

2,1

, u= 0 on ∂Ω, where the domain Ω is partitioned by square elements. The exact solution belongs to H53(Ω)∩H01(Ω). For this problem, the approximate value forEis

E=−6.689868958058575×103.

Proposition3.4, bound (3.6), and conjecture (3.11) indicate that the error is bounded as fol- lows

(4.2) a(u−uACMS, u−uACMS)≤C h2

maxKIK kfk2L2(Ω)+C h43

maxeIekuk2H53(Ω). The following experiments illustrate the sharpness of this result.

Using only one bubble eigenfunction and one edge-bubble eigenfunction, Figure 4.5 illustrates the convergence when the number of elements is increased. As expected, the con- vergence curve exhibits a decrease proportional toh23. The a posteriori error indicator (3.13) (withε= 0) decreases also proportionally toh23. The ratio between the error indicator and the semi-norm varies between 4 and 10.

Next, only one bubble eigenfunction is used while the mesh sizehis decreased. The number of edge-bubble eigenfunctions is set to the integer part of1h. Figure4.6illustrates the convergence when the number of elements is increased. SincemaxeIe=O(1h), bound (4.2) suggests a convergence rate ofh, which is matched by the numerical experiment. The plot confirms that the impact of bubble eigenfunctions depends only on the regularity of the right hand sidef.

(14)

10−2 10−1 100 10−4

10−3 10−2 10−1

Mesh size

Absolute error

H1 semi−norm A posteriori error indicator Line y = Ch

FIG. 4.6. Convergence curve whenf= 1for a fixed number of bubble eigenfunctions (L-shaped domain).

100 101 102 103

10−4 10−3 10−2 10−1

Number of edge−bubble eigenfunctions per edge

Absolute error

H1 semi−norm A posteriori error indicator Line y = CI−2/3

FIG. 4.7. Convergence curve for a varying number of edge-bubble eigenfunctions (L-shaped domain,f= 1, h=14,IK= 256).

Finally, in the next experiment, the number of edge-bubble eigenfunctions is varied while the mesh sizehis set to 14 and the number of bubble eigenfunctions to 256. Figure4.7il- lustrates the convergence when the number of edge-bubble eigenfunctions is uniformly in- creased. The semi-norm of the error and the a posteriori error indicator decrease proportion- ally toI

2

e 3 before reaching a plateau set by the constant number of bubble eigenfunctions.

Bound (4.2) suggests only a decrease proportional toI

1

e 2. This discrepancy is due to relation (A.2) which does not exploit smoothness beyondH1(Γ).

4.3. Problem with varying coefficient. Finally, consider the problem

−∇(c(x)∇u(x)) =−1 inΩ = [0,1]×[0,1], u= 0 on∂Ω,

(4.3)

(15)

TABLE4.1

Error evolution for problem (4.3) as the mesh sizehis reduced.

Mesh size E(v)− E ηint ηedge

h=14 6.81×102 1.80×101 1.8×103 h=18 2.04×102 4.24×102 3.6×104 h=161 6.94×103 1.31×102 6.34×105 h=321 1.35×103 3.58×103 6.98×106 where the coefficientcis

c(x, y) = 2 + 1.8 sin 2πxε

2 + 1.8 cos 2πyε + 2 + sin 2πyε 2 + 1.8 sin 2πxε

withε = 18. The domain Ωis partitioned by square elements. This problem was initially studied in [21]. The exact solution belongs to H2(Ω)∩H01(Ω). For this problem, the approximate value forEis

E=−4.826726636113407×103.

The objective of this subsection is to assess the quality of the error indicator in Proposi- tion3.6. Denote

ηint= X

K∈Th

kf− PIK(f)k2L2(K)

λIK,K

and

ηedge= X

K∈Th

kf− PIK(f)k2L2(K)

X

e∂KΓ

1 λ2Ie,e

!

+X

e⊂Γ

Je νTeA∇uACMS2

L2(e)

λIe,e

.

Table4.1describes the reduction of errors and error indicators as the mesh size is refined.

One edge-bubble eigenfunction for each edge and no bubble eigenfunctions are used. The energy differences and the indicatorηint exhibit a reduction proportional toh2. As can be seen in Figure4.4, a superconvergence phenomenon for the edge part of errors is possible;

see Bourquin [9, p. 45]. Here, the edge indicatorηedgeis decreasing slightly faster thanh3 for this range of mesh sizes.

Table4.2illustrates the same information when the number of edge-bubble eigenfunc- tions is uniformly increased. The mesh size is set toh= 18 and no bubble eigenfunction is used. For this setup, the energy differences reach a plateau while the edge indicatorηedgeis decreasing slightly faster than(maxeIe)1, the prediction in (3.12).

5. Conclusion. This paper derives a priori error estimates for a special finite element discretization based on component mode synthesis. The a priori error bounds state the explicit dependency of constants with respect to the mesh size and the first neglected eigenvalues. A residual-based a posteriori error indicator is also presented. Numerical experiments illustrate that the error indicator is reliable.

Such indicator could guide the adaptive selection for the number of bubble and edge- bubble eigenfunctions. In practice, the basis functions and eigenfunctions used in this special finite element method are computed numerically by introducing a nested finer grid. To en- hance the practicality of these special finite elements, future works will study error estimates and a posteriori error indicators for the resulting two-grid scheme.

(16)

TABLE4.2

Error evolution for problem (4.3) as the number of edge-bubble eigenfunctions is increased andh=18.

Edge-bubble eigenfunctions E(v)− E ηint ηedge

1 2.04×102 4.24×102 3.65×104 2 1.81×102 4.24×102 1.69×104 4 1.62×102 4.24×102 5.25×105 8 1.59×102 4.24×102 1.63×105

FIG. A.1. Example of domainD.

Acknowledgments. U. Hetmaniuk acknowledges the partial support by the National Science Foundation under grant DMS-0914876.

Appendix A. Review of properties of the Steklov-Poincar´e operator.

In this section, properties of the Steklov-Poincar´e operator are compiled. Further details and references are included in Bourquin [10] and Khoromskij and Wittum [25].

Consider a bounded polygonal domain D ⊂ R2 partitioned into two regions, D = D1∪D2. The subdomainsD1 andD2 are bounded convex polygons with straight edges. The interfaceS=D1∩D2is illustrated in FigureA.1.

For anyτ ∈H

1 2

00(S), the energy-minimizing extensionE1(τ)∈H1(D1)is defined as the unique solution to the problem

−∇ ·(A∇E1(τ)) = 0 inD1, E1(τ) =τ onS, E1(τ) = 0 on∂D1∩∂D.

The energy-minimizing E2(τ) ∈ H1(D2) is defined similarly in D2. The matrix A is uniformly symmetric positive definite onDas described by (2.1).

Introduce the symmetric bilinear form b(τ, η) =

Z

D1

∇E1(τ)TA∇E1(η) + Z

D2

∇E2(τ)TA∇E2(η),

for any functionτandηinH

1 2

00(S). The continuity and coerciveness ofbare consequences of the continuity of the energy-minimizing extension, of the trace operator on S, and of

(17)

properties of A. Given that the injection of H

1 2

00(S) into L2(S) is compact (see Bour- quin [10, p. 390–391]), there exists a self-adjoint unbounded linear operatorB onL2(S) with compact inverse such that

b(τ, η) = Z

S

(Bτ)η, ∀η∈L2(S) and for any arbitraryτin the domain of the operatorB,

D(B) =n τ ∈H

1 2

00(S) ; Bτ =νT1A∇E1(τ) +νT2A∇E2(τ)∈L2(S)o , whereν1, respectivelyν2, is the unit outer normal vector to∂D1, respectively∂D2. Note that the operatorBcan be decomposed as follows

Bτ=B1τ+B2τ with B1τ=νT1A∇E1(τ) and B2τ=νT2A∇E2(τ) for any elementτinD(B).

Whenηbelongs toH

1 2

00(S)∩H1(S), the compatibility conditions for traces on a poly- gon [2, Theorem 6.1] indicate thatηsatisfies

˜

η|∂D1 ∈H1(∂D1) and η˜|∂D2 ∈H1(∂D2). Then we have

kBηkL2(S)≤ kB1η˜kL2(∂D1)+kB2η˜kL2(∂D2)

≤CAkη˜kH1(∂D1)+CAkη˜kH1(∂D2)≤CAkηkH1(S), (A.1)

where CA denotes a generic constant that may depend on the coefficient matrix A. The constantCA does not depend on the length of S or on the diameter ofD; see Neˇcas [27, Theorem 1] for the bound betweenkBkη˜kL2(∂Dk)andkη˜kH1(∂Dk), wherek= 1,2.

Spectral decomposition. Spectral theory yields a familyn)+n=1 forming an orthogo- nal basis ofH

1 2

00(S)andL2(S)and a sequence of real numbers(θn)+n=1 such that b(φn, η) =θn

Z

S

φnη, ∀η∈H

1 2

00(S), and

Z

S

φ2n= 1 and 0< θ1≤θ2≤ · · ·. The eigenfunctions also satisfyBφnnφn; see Bourquin [10, p. 392].

Forη∈L2(S), define the projection ΠL(η) =

L−X1 n=1

Z

S

ηφn

φn.

WhenBηbelongs toL2(S), we write Z

S

ηφn= 1 θn

Z

S

η(Bφn) = 1 θn

Z

S

(Bη)φn.

(18)

Forη∈H

1 2

00(S)withBη∈L2(S), it holds that b(η−ΠL(η), η−ΠL(η)) =

+

X

n=L

θn

Z

S

ηφn

2

=

+

X

n=L

1 θn

Z

S

(Bη)φn

2

≤ 1

θLkBη−ΠL(Bη)k2L2(S)≤ 1

θL kBηk2L2(S). In particular, whenη ∈H

1 2

00(S)∩H1(S), relation (A.1) implies thatBηbelongs toL2(S).

In this case, the projection error satisfies

(A.2) b(η−ΠL(η), η−ΠL(η))≤ CA

θL kηk2H1(S). Bounds in dual spaces will also be needed. Forη∈H

1 2

00(S), we write kη−ΠL(η)k2L2(S)=

+

X

n=L

Z

S

ηφn

2

=

+

X

n=L

1 θ2sn θn2s

Z

S

ηφn

2

≤ 1 θ2sL

+

X

n=L

θn2s Z

S

ηφn

2

for0≤s <12. Using the equivalence between the norms vu

ut+X

n=1

(1 +θ2sn ) Z

S

ηφn

2

and kηkHs(S) for 0≤s < 1 2 (see, for example, Khoromskij and Wittum [25, Section 1.7]), we obtain (A.3) kη−ΠL(η)k2L2(S)≤Cs,A

θsL kηk2Hs(S)

for0< s <12, whereCs,Adoes not depend on the length ofS.

After continuously extending the projection ΠL toH12(S) = H

1 2

00(S)

, similar estimates hold inH12(S),

(A.4) kη−ΠL(η)k2H1

2(S)≤ 1

θL kη−ΠL(η)k2L2(S)≤ Cs,A

θ1+2sL kηk2Hs(S)

for0≤s <12, whereCs,Adoes not depend on the length ofS.

Appendix B. Proof of Proposition3.6.

Proof. Recall that the exact solutionusatisfies a(u, v) =

Z

f v, ∀v∈H01(Ω)

and thatPIKis the projection operator defined by (3.5). The functionf can be decomposed as follows

f = X

K∈Th

PIK(f) + X

K∈Th

[f− PIK(f)]

参照

関連したドキュメント

The problem is modelled by the Stefan problem with a modified Gibbs-Thomson law, which includes the anisotropic mean curvature corresponding to a surface energy that depends on

Let X be a smooth projective variety defined over an algebraically closed field k of positive characteristic.. By our assumption the image of f contains

Thus, we use the results both to prove existence and uniqueness of exponentially asymptotically stable periodic orbits and to determine a part of their basin of attraction.. Let

Next, we prove bounds for the dimensions of p-adic MLV-spaces in Section 3, assuming results in Section 4, and make a conjecture about a special element in the motivic Galois group

It turns out that the symbol which is defined in a probabilistic way coincides with the analytic (in the sense of pseudo-differential operators) symbol for the class of Feller

Then it follows immediately from a suitable version of “Hensel’s Lemma” [cf., e.g., the argument of [4], Lemma 2.1] that S may be obtained, as the notation suggests, as the m A

We give a Dehn–Nielsen type theorem for the homology cobordism group of homol- ogy cylinders by considering its action on the acyclic closure, which was defined by Levine in [12]

In this paper we focus on the relation existing between a (singular) projective hypersurface and the 0-th local cohomology of its jacobian ring.. Most of the results we will present