• Keine Ergebnisse gefunden

and Raviart–Thomas-type mixed finite elements in elastostatics

N/A
N/A
Protected

Academic year: 2023

Aktie "and Raviart–Thomas-type mixed finite elements in elastostatics"

Copied!
22
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Numerische Mathematik

c Springer-Verlag 1996

Electronic Edition

Symmetric coupling of boundary elements

and Raviart–Thomas-type mixed finite elements in elastostatics

Ulrich Brink1, Carsten Carstensen2, Erwin Stein1

1 Institut f¨ur Baumechanik und Numerische Mechanik, Universit¨at Hannover, D-30167 Hannover, Germany

2 Mathematisches Seminar II, Christian-Albrechts-Universit¨at zu Kiel, D-24098 Kiel, Germany

Received February 21, 1995 / Revised version received December 21, 1995

Dedicated to Professor G. C. Hsiao on occasion of his 60th birthday.

Summary. Both mixed finite element methods and boundary integral methods are important tools in computational mechanics according to a good stress ap- proximation. Recently, even low order mixed methods of Raviart–Thomas-type became available for problems in elasticity. Since either methods are robust for critical Poisson ratios, it appears natural to couple the two methods as proposed in this paper. The symmetric coupling changes the elliptic part of the bilinear form only. Hence the convergence analysis of mixed finite element methods is applicable to the coupled problem as well. Specifically, we couple boundary el- ements with a family of mixed elements analyzed by Stenberg. The locking-free implementation is performed via Lagrange multipliers, numerical examples are included.

Mathematics Subject Classification (1991): 65N30, 65R20

1. Introduction

In the classical finite element approach, the displacements are the unknowns while the stresses are computed afterwards in lower accuracy. In many applications, the stresses rather than the displacements are of primary interest. In the boundary element method (BEM), displacements and stresses in the interior of the domain are approximated with the same order.

Regarding finite elements, the approximation of the stresses can be improved by mixed methods. Here we consider Raviart–Thomas-type finite elements due to Stenberg [11], which are an improvement of Arnold–Brezzi–Douglas’s PEERS element (plane elasticity element with reduced symmetry) [1]. The unknowns are three independent fields, namely the stress tensor, which is not a priori assumed to be symmetric, the rotation, which acts as a Lagrange multiplier to enforce the symmetry of the stress tensor in a weak form, and the displacement vector.

(2)

The analysis is based on the theory of mixed methods but refined by using mesh-dependent norms (Pitk¨aranta and Stenberg [9], [10]), by stability proofs on patches of elements and by weakening of the symmetry condition.

Another advantage of these elements is the capability of modeling nearly incompressible elasticity; more precisely, the relative error is independent of Poisson’s ratioν: there is no locking.

In many applications, for example, if nonlinearities in a bounded domain ΩF are present as well as homogeneous, isotropic linear elastic material in an unbounded, e.g., exterior, domain ΩB, one might combine the finite element method (inΩF) with the boundary integral method (acting on∂ΩBbut treating the problem inΩB). The symmetric coupling with boundary elements was proposed and analyzed by Costabel in [6] where the displacements inside the FEM domain are sought in the Sobolev space H1(ΩF); traces on the interface are inserted into the boundary integral equations, while the equilibrium of tractions across the interface is satisfied in a weak sense only.

In contrast, in the mixed FEM under consideration here, the stresses σ are required to satisfyσ∈L2(ΩF) and divσ∈L2(ΩF), while the displacements are sought in L2(ΩF) only. Hence, the discrete tractions across interelement sides are continuous and, consequently, our coupled scheme is designed to yield continuous tractions across the interface between FEM and BEM while continuity of the displacements across the interelement sides and the interface is satisfied in a weak sense.

The construction of finite element spaces with continuous interelement trac- tions may be cumbersome and hence is enforced by using Lagrange multipliers.

Then all unknowns, except the Lagrange multipliers, are discontinuous across the interelement sides and can be eliminated on each element before assembling the global linear system. We extended this scheme to the coupled system.

A similar coupling is possible with other mixed finite element methods than those considered in this paper. The weakening of the symmetry of the stress tensor is a particular way to construct stable finite element spaces, but is by no means necessary for the coupling.

The paper is organized as follows: A model problem (as depicted in Fig. 1) is described in its strong form in Sect. 2 and rewritten in a weak form. We prove existence and uniqueness of solutions and a norm estimate by using Brezzi’s theory of mixed problems. The discretization is described in Sect. 3 where we state convergence estimates under some hypotheses on the mixed finite element methods. Proofs are given in Sect. 4 while in Sect. 5 Stenberg’s locking-free fam- ily of methods is discussed. We describe the implementation by using Lagrange multipliers in Sect. 6 and give some numerical examples in Sect. 7.

2. The coupled problem

Let Ω be a bounded polygonal (resp. polyhedral) domain in Rd, d = 2 (resp.

d = 3) with boundary ∂Ω = Γu ∪Γt with disjointΓu and Γt, either having a positive surface measure.

(3)

Γu F

FEM

Γt

-n

n B

BEM

Γ

Fig. 1. Notation

Throughout this paper, we consider the linear elasticity problem divσ = −f inΩ

σ = Cε(u) inΩ

u = 0 onΓu

σn = 0 onΓt. (2.1)

The displacement field is denoted by u, the related (linear Green) strain tensor is ε(u) = 12( grad u + ( grad u)T). The elasticity tensor C describes the stress–

strain relationship; in the simplest case we have σ= Cε= λtrεI + 2µε where λwhereµ are the Lam´e coefficients and I denotes the identity matrix. (In the two-dimensional case, this is the constitutive equation of plane strain).

As depicted in Fig. 1, the domain Ω is partitioned into Ω= ΩF∪Γ ∪ΩB; the outward unit normal ofΩFis denoted by n.

OnΩFa mixed finite element method based on the Hellinger–Reissner prin- ciple is applied. This means the displacement u :F → Rd and the stress σ : ΩF → Rd×d are independent unknowns. Furthermore, the stress tensor σ is not a priori assumed to be symmetric. Symmetry will be enforced in a weak form by a Lagrange multiplier technique. To obtain a variational formulation, we choose test functions τ : ΩF → Rd×d satisfying τn = 0 on Γt, and gain from

(2.1)b Z

F

τ:C1σdΩ− Z

F

τ :ε(u) dΩ= 0 so that integration by parts gives

Z

F

τ:C1σdΩ+ Z

F

divτ·u dΩ+ Z

F

τ:γdΩ= Z

Γ

τn·u ds

(4)

where the rotation γ := 12( grad u( grad u)T) is a new variable for the skew- symmetric part of the displacement gradient,

γ∈W :={η∈L2(ΩF)d×d :η+ηT= 0}. The stresses and displacements are sought in

H := {τ∈L2(ΩF)d×d : divτ∈L2(ΩF)d, τn = 0 onΓt}, L := L2(ΩF)d.

H is a Hilbert space when endowed with the norm kτkdiv := (kτk20,ΩF+ kdivτk20,ΩF)1/2.

In summary, the variational form of the problem inΩFreads: Given a body load f and a displacementϕonΓ, find (u, σ, γ)∈L ×H ×W satisfying

R

Fτ :C1σdΩ+R

F divτ·u dΩ +R

Fτ:γdΩ = R

Γτn·ϕds ∀τ ∈H R

F divσ·vdΩ = −R

Ff ·vdΩ ∀v∈L R

Fσ:ηdΩ = 0 ∀η∈W .

(2.2)

The identity (2.2)a is derived above, (2.2)b is a weak form of (2.1)a, and (2.2)c is a weak form of the symmetry ofσ.

We assume that the body load fL2(Ω)d vanishes onΩB for simplicity and that the Lam´e coefficients are constant onΩB. Then, at any point x ∈ΩB, the displacement field can be represented by the Betti formula

u(x ) =− Z

Γ

G(x,y) Tyu(y) dsy+ Z

Γ

TyG(x,y)T

u(y) dsy.

Here Tyu(y) =Cε(u(y)) n(y) is the traction corresponding to u at a point y ∈ Γ, and TyG(x,y) are the columnwise tractions of G(x,y) at y. G(x,y) is the fundamental solution and equals

λ+3µ 4πµ(λ+2µ)

n

log|x1y|I +λλ+3+µµ(x|xy)(xy|2y)T

o if d = 2,

λ+3µ 8πµ(λ+2µ)

n 1

|xy|I +λλ+3+µµ(x|xy)(xy|3y)T

o if d = 3.

Letting x →Γ we obtain with the classical jump relations the boundary integral equation

1

2u =Vt + Ku (2.3)

with t (y) = Tyu(y) and the integral operators (Vt )(x ) =

Z

Γ

G(x,y) t (y) dsy, x ∈Γ,

(Ku)(x ) = Z

Γ

TyG(x,y)T

u(y) dsy, x∈Γ .

(5)

Applying the traction operator Tx we get another boundary integral equation

1

2t =K0tWu (2.4)

where

(K0t )(x ) = Z

Γ

TxG(x,y) t (y) dsy, x ∈Γ , (Wu)(x ) = −Tx

Z

Γ

TyG(x,y)T

u(y) dsy, x∈Γ .

The symmetric coupling of (2.2) with the integral equations (2.3) and (2.4) is performed as follows: Introduce a new variable ϕ:= u|Γ belonging to the trace space

H1/2:= H1/2(Γ)d :={u|Γ : uH1(Ω)d},

and use continuity of the tractions onΓ, i.e., t =σn. Then, (2.3) reads ϕ=−V (σn) + (12I + K )ϕ

and this is inserted in the right-hand side of (2.2)a while (2.4) reads Wϕ+ (12I + K0)(σn) = 0

and this is added in a weak form to (2.2). (This coupling is in a sense ‘dual’ to the more classical approach, see e.g. [4, 6, 7], where a new variable is introduced for σn on the interface while for u|Γ continuity is used.) The resulting weak formulation is rewritten in a saddle point structure: Find (σ, ϕ,u, γ) ∈ H × H1/2×L ×W such that

a(σ, ϕ;τ, ψ) + b(τ; u, γ) = 0 b(σ;v, η) = −R

Ff ·vdΩ (2.5)

for all (τ, ψ , v, η)∈H×H1/2×L ×W . Here, a(σ, ϕ;τ, ψ) :=

Z

F

τ:C1σdΩ +hτn, V (σn)i −

τn,(12I + K )ϕ

− hψ ,Wϕi −

ψ ,(12I + K0)(σn) b(σ;v, η) :=

Z

F

divσ·vdΩ+ Z

F

σ:ηdΩ.

Throughout this paper, hϕ , ψi denotes the extension of the L2-scalar product R

Γϕ·ψds to the duality in H1/2×H1/2; Hs := Hs(Γ)d.

(6)

Remark 2.1. It is known that the mappings V : H1/2H1/2, K : H1/2H1/2, K0 : H1/2H1/2, W : H1/2H1/2 are well defined, linear and continuous (cf., e.g., [5]). V and W are symmetric; K0 is dual to K and W is positive semi-definite and ker W = kerε|Γ, i.e., the kernel of W consists of the (linearized) rigid body motions. W is positive definite on (H1/2/kerε)2. For d = 3, V is positive definite. For d = 2, V is positive definite when restricted to H01/2 where

H0s:={w∈Hs : Z

Γ

wds = 0} ≡Hs/Rd.

We refer to [7] for proofs in case d = 3 and mention that the proofs work verbatim in case d = 2 (provided the radiation condition gives a sufficiently strong decay which is guaranteed owing to the restriction on H01/2).

Remark 2.2. We note thatσn is defined in H1/2 via Green’s formula even if σ ∈ H as follows. Given v ∈ H1/2 extend it to some v ∈ H1(ΩF)d with v|Γu = 0. Then, let

hv , σni:=

Z

F

σ: gradvdΩ+ Z

F

v·divσdΩ.

The right-hand side is well defined and depends linearly and continuously on v andσ. Furthermore,

kσnk1/2Ckσkdiv. (2.6)

In view of the remarks, a is a symmetric and continuous bilinear form on (H ×H1/2)2, and b is continuous onH ×(L ×W ).

Theorem 2.1. For every fL2(ΩF)dthe saddle point problem (2.5) has a unique solution satisfying

kσk0,ΩF+ kdivσk0,ΩF+ kϕk1/2+ kuk0,ΩF+ kγk0,ΩFCkfk0,ΩF

(2.7)

with a positive constant C which is independent of f .

Proof. We apply the theory of saddle point problems, cf., e.g., [3, Sect. II.1]. It is sufficient to verify surjectivity of b and the inf–sup condition on a. As it is well known, the bilinear form b has the following surjectivity property: For all (v, η)∈L ×W there is someτ∈H satisfying

divτ=v and asτ :=12(τ−τT) =η.

(2.8)

Therefore, it remains to prove that, for a constantα >0, inf

(σ,ϕ) sup

(τ,ψ)

a(σ, ϕ;τ, ψ)

(kσk0,ΩF+ kϕk1/2)(kτk0,ΩF+ kψk1/2)≥α (2.9)

where in inf(σ,ϕ) and sup(τ,ψ) the nonzero arguments run through ker B , ker B :={τ ∈H : divτ= 0 and asτ= 0} ×H1/2.

(7)

We will prove (2.9) in two steps partly arguing as in [4]: Given (σ, ϕ)∈ker B letϕ=ϕ0+ r0 whereϕ0H1/2/kerεand r0∈kerε, kerεthe (linearized) rigid body motions. Let tc∈Rd be defined by

Z

Γ

(σn−tc) ds = 0.

Then, letσc:=Cε(uc) where ucH1(ΩF)d is the solution of divσc= 0 onΩF whileσcn = tconΓ,σcn = 0 onΓt and uc= 0 onΓu. Furthermore, letϕc∈Rd

satisfy

V (σntc) + (12I + K )ϕ ,tc

− hσn,uci=htc, ϕci. (2.10)

(If tc= 0, every ϕc ∈Rd solves (2.10).) Note thatϕc can be chosen such that kϕck1/2C (kσk0,ΩF+ kϕk1/2). Thus,

kσ−σck0,ΩF+ kϕ−ϕck1/2C (kσk0,ΩF+ kϕk1/2) (2.11)

where C >0 is independent ofσandϕ. Integration by parts shows Z

F

σc:C1σdΩ=hσn,uci. (2.12)

Then, using (2.10) and Kϕc= 12ϕc, a(σ, ϕ;σ−σc,−ϕ+ϕc) =

Z

F

σ:C1σd

+hσn−tc,V (σntc)i+hϕ , Wϕi. Since W defines a positive definite bilinear form on (H1/2/kerε)2 and since V is positive definite on (H01/2)2 (recallσn−tcH01/2 by definition of tc) we obtain

a(σ, ϕ;σ−σc,−ϕ+ϕc)≥C1(kσk20,ΩF+ kϕ0k21/2) (2.13)

with C1>0 depending only on C and W .

Now we show that for every rigid body motion r0 there is a stress field τ0

with (τ0,0)∈ker B and

a(0,r00,0) =hr0,r0i. (2.14)

For example, let τ0H satisfy divτ0 = 0 in ΩF and τ0n =r0 on Γ. Since Kr0= 12r0, we conclude (2.14).

In the second step we assume for contradiction that (2.9) is false. Hence we may find a sequence (σj, ϕj) in ker B withjk0,ΩF+ kϕjk1/2 = 1 and

sup

(τ,ψ)ker B\{0}

a(σj, ϕj;τ, ψ)/(kτk0,ΩF+ kψk1/2)<1/j (2.15)

for all j . Letϕj0j + rj whereϕ0jH1/2/kerεand rj ∈kerε.

According to (2.11), (2.13) and (2.15) we get

(8)

0 = lim

j→∞(kσjk0,ΩF+ kϕ0jk1/2)

and hence, since rj is a bounded sequence in a finite dimensional space, (σj, ϕj)→(0,r0) as j → ∞

(2.16)

for a subsequence (which is not relabeled) and a rigid body motion r0. In particu- lar, kr0k1/2 = 1. Letτ0satisfy (2.14) with r0as in (2.16). Since a is continuous, (2.16) shows

0 = lim

j→∞

a(σj, ϕj0,0) kτ0k0,ΩF

= a(0,r00,0) kτ0k0,ΩF

>0

owing to (2.14) and r0/= 0. This contradiction verifies (2.9); the proof is finished.

u t

3. Approximation and convergence

For the finite element method we consider a regular family of triangulationsTh of Ω¯F. Two different triangles (resp. tetrahedrons) inTh are either disjoint or have one common side or edge or vertex. Let the sides of finite elements in Th on the interfaceΓ define a partitionEh ofΓ and takeEh as boundary elements for simplicity. These partitions give rise to finite-dimensional subspacesHh, Hh1/2, Lh andWh which are assumed to satisfy (H1)–(H3); examples are considered in Sect. 5.

(H1) (Conformity and approximation property) There holds Hh×Hh1/2×Lh ×WhH×H1/2×L ×W

andRdHh1/2. For all (τ, ψ , v, η)H ×H1/2×L ×W , 0 = lim

h0

τhinf∈Hh

kτ−τhkdiv + inf

ψhHh1/2

kψ −ψhk1/2

+ inf

vh∈Lh

kv−vhk0,ΩF+ inf

ηh∈Wh

kη−ηhk0,ΩF

.

(H2) (Equilibrium condition) For eachτhHh, the condition b(τh;vh, ηh) = 0 ∀(vh, ηh)∈Lh ×Wh

implies divτh = 0.

Remark 3.1. Note that a piecewise polynomial function τh satisfies divτhL2(ΩF)d if and only if the tractionsτhn are continuous across interelement sides (which is thus implied byHhH).

(9)

(H3) (inf–sup condition) The bilinear form b satisfies the inf–sup condition in some mesh-dependent norm. That is there exist a norm k·kH,hon some space H(h)H withHhH(h)and

hk0,ΩF ≤ kτhkH,hChkdiv

(3.1)

for allτhHh and a norm k(·,·)kh inL(h)×Wh withLhL(h)L such that a and b are uniformly continuous (i.e., with h-independent bounds), and there is a positive constantβ such that for all (vh, ηh)∈Lh ×Wh

sup

τh∈H0h\{0}

b(τh;vh, ηh) kτhkH,h

≥βk(vh, ηh)kh. (3.2)

where H0h :={τhHh : τhn|Γ = 0}. Both C andβ are assumed to be independent of h.

Remark 3.2. To ensure the continuity of a, we need the estimatenk1/2CkτkH,h ∀τ∈H(h)

(3.3)

Note that (3.3) does not affect (3.2) owing toH0h.

Example 3.1. Let Th be a finite element triangulation ofF and letSh denote the set of sides in the interior ofFandSh¯the set of all finite element sides. The value of the jump ofvacross an interelement side S is denoted by [v], while the diameter of S is hS. We follow Stenberg [10, 11]: The first mesh-dependent norm

kτk2H,h := kτk20,ΩF+ X

SSh¯

hS Z

S

n|2ds

is defined on the ‘intermediate’ space

H(h):={τ∈HnL2(S )dSSh¯}, and the second

k(v, η)k2h := X

T∈Th

(kε(v)k20,T + kη−ω(v)k20,T)

+ X

S∈Sh

hS1 Z

S

|[v]|2ds + X

S⊂Γu

hS1 Z

S

|v|2ds

onL(h)×Wh withω(v) := 12( gradv−( gradv)T) and

L(h):={v∈L2(Ω)d :v|TH1(T )dTTh}.

On Hh, the norms k·kH,h and k·k0,ΩF are uniformly equivalent, i.e., for all τhHh

hk0,ΩF ≤ kτhkH,hChk0,ΩF. (3.4)

(10)

In these norms, the bilinear forms a and b are continuous, i.e., there exists C >0 such that for all (σ, ϕ), (τ, ψ)H(h)×H1/2

a(σ, ϕ;τ, ψ)≤C (kσkH,h+ kϕk1/2)(kτkH,h+ kψk1/2) (3.5)

and for all (τ, v, η)H(h)×L(h)×W

b(τ;v, η)≤CkτkH,hk(v, η)kh. (3.6)

To show continuity of a, it remains to verify (3.3). Recall kτnk1/2 = sup

v∈H1/2(Γ)d

hτn, vi kvk1/2.

Eachv∈H1/2(Γ)dcan be extended tov∈H1(ΩF)dwithv|Γu = 0 andkvk1,ΩFCkvk1/2. Integration by parts yields

hτn, vi= Z

F

τ: gradvd+ b(τ;v,0).

Further, k(v,0)kh ≤ kvk1,ΩFsince the jumps [v] on the interelement sides vanish.

Thus, using the continuity of b in the mesh-dependent norms, we obtain hτn, vi ≤CkτkH,hkvk1,ΩF.

This proves (3.3).

Remark 3.3. We assume throughout this paper that the solution to (2.5) satisfies σ ∈ H(h) and uL(h) for all fL2(ΩF)d, which is a regularity condition.

In the situation of Example 3.1, the condition uHs(ΩF)d with s > 3/2 is sufficient.

The discretized saddle point problem consists in finding (σh, ϕh,uh, γh) ∈ Hh×Hh1/2×Lh×Whsuch that for all (τh, ψh, vh, ηh)∈Hh×Hh1/2×Lh×Wh

a(σh, ϕhh, ψh) + b(τh; uh, γh) = 0 b(σh;vh, ηh) = −R

Ff ·vhdΩ (3.7)

Theorem 3.1. Assuming (H1)–(H3), the discrete problem (3.7) has exactly one solution. There exists some h-independent constant C >0 such that

kσ−σhkH,h+ kϕ−ϕhk1/2

C

τhinf∈Hh

kσ−τhkH,h+ inf

ψhHh1/2

kϕ−ψhk1/2+ inf

ηh∈Wh

kγ−ηhk0,ΩF

.

To derive L2-estimates for the displacements, we need a regularity–approxi- mation estimate and a mapping Ph :LLh×Wh.

(11)

(H4) (Regularity–approximation property) In the mesh-dependent norms we have the following estimate where%(h) is an h-dependent constant such that

inf

(τhh,vhh)∈Hh×Hh1/2×Lh×Wh

k(σ−τh, ϕ−ψh,u−vh, γ−ηh)kh

≤ %(h)kfk0,ΩF

(3.8)

for all fL2(ΩF)d and (σ, ϕ,u, γ) solving (2.5) (cf. Theorem 2.1). The mesh- dependent norm k·kh in (3.8) is defined by

k(τ, ψ , v, η)kh := kτkH,h+ kψk1/2+ k(v, η)kh. (3.9)

Lemma 3.1. There is a linear mapping Ph :LLh×Wh which satisfies b(τh;v,0) = b(τh; Ph(v))

for allτhHh and allv∈L.

Modifications of this mapping Ph are frequently used in the literature for a proof of the inf–sup condition; we conversely may construct it from (H3). In particular cases, Ph is known explicitly (see Sect. 5).

Theorem 3.2. Assuming (H1)–(H4), there exists some h-independent constant C >0 such that, with%(h) as in (H4),

kPh(u)(uh, γh−γ)k0,ΩFC·%(h)·

τhinf∈Hh

kσ−τhkH,h

+ inf

ψhHh1/2

kϕ−ψhk1/2 + inf

ηh∈Wh

kγ−ηhk0,ΩF

.

Remark 3.4. To obtain an estimate for kuuhk0,ΩF, let ( ˜u,γ) := P˜ h(u) and observe that

kuuhk0,ΩF ≤ kuu˜k0,ΩF+ kPh(u)(uh, γh−γ)k0,ΩF. (3.10)

The last term is estimated by Theorem 3.2, and, according to the specific Ph, an a priori estimate of kuu˜k0,ΩF is available.

Proofs are given in Sect. 4 while examples are studied in Sect. 5.

4. Proofs

The proofs of Theorem 3.1 and 3.2 use several lemmas where Z := {τ∈H : divτ= 0 andτ =τT}

Zh := {τhHh : b(τh;vh, ηh) = 0 ∀(vh, ηh)∈Lh×Wh} ker Bh := Zh×Hh1/2.

Throughout this section, C >0 denotes a generic h-independent constant.

(12)

Lemma 4.1. [11] For allτhZh Z

F

τh :C1τhdΩ≥Chk20,ΩF. Lemma 4.2. For all ¯σ∈Z there exists ¯σhZh satisfying

kσ¯ −σ¯hkdivC inf

τh∈Hh

kσ¯−τhk0,ΩF.

Proof. The result follows as in [11]; for related results we refer to [3, Proposition II.2.5] and [2, Remark III.4.6]. We give a proof for completeness and to stress that (H1)–(H3) are sufficient (even for different situations on the boundary). Let σ˜ be the best approximant to ¯σ∈Z inHh with respect to the L2(ΩF)-norm. By Lemma 4.1 and (H3), the mixed finite element problem

Z

F

σ¯h :C1τhd+ b(τh; ¯uh,γ¯h) = L1h) b( ¯σh;vh, ηh) = L2(vh, ηh)

(with linear forms L1,L2) satisfies ellipticity and inf–sup conditions in mesh- dependent norms so that we have a unique solution ( ¯σh,u¯h,γ¯h)∈Hh×Lh×Wh

satisfying

Z

F

( ¯σh−σ) :¯ C1τhd+ b(τh; ¯uh,γ¯h) = 0 (4.1)

b( ¯σh−σ;˜ vh, ηh) = 0 (4.2)

for all (τh, vh, ηh)∈Hh×Lh×Wh. Note that ¯σh−σ˜ ∈Zh according to (4.2).

Hence, Lemma 4.1 yields Ckσ¯h−σ˜k20,ΩF

Z

F

( ¯σh−σ) :˜ C1( ¯σh−σ) dΩ˜

= Z

F

( ¯σ−σ) :˜ C1( ¯σh−σ) dΩ˜ owing to (4.1). Thus, by Cauchy’s inequality,

kσ¯h −σ˜k0,ΩFCkσ¯−σ˜k0,ΩF.

Then, the triangle inequality and (H2) finish the proof of the lemma. ut Lemma 4.3. There is a constantα >0 such that

inf

(σhh) sup

(τhh)

a(σh, ϕhh, ψh)

(kσhkH,h+ kϕhk1/2)(kτhkH,h+ kψhk1/2) ≥α (4.3)

where the nonzero arguments in inf(σhh)and sup(τ

hh) belong to ker Bh.

(13)

Proof. According to the equilibrium condition (H2) it is sufficient to consider the norm k·k0,ΩF instead of k·kH,h (because, by (3.1), kτhk0,ΩF ≤ kτhkH,hChkdiv = Chk0,ΩF for allτhZh). We proceed as in the proof of Theorem 2.1. Assuming that (4.3) is false we find a sequence of meshes, discrete spaces and discrete functions (σj, ϕj) in ker Bhj with kσjk0,ΩF+ kϕjk1/2 = 1 and

sup

(τ,ψ)ker Bhj\{0}

a(σj, ϕj;τ, ψ)/(kτk0,ΩF+ kψk1/2)<1/j (4.4)

for all j . Letϕj0j + rj whereϕ0jH1/2/kerεand rj ∈kerε.

Since (σj, ϕj) are bounded inH×H1/2we may extract a weakly convergent subsequence (not relabeled); so assume,

j, ϕ0j)*(σ, ϕ0) (weakly) inH ×H1/2/kerε

for some (σ, ϕ0) ∈ H ×H1/2/kerε. By (H2) we have divσj = 0 and so divσ= 0. Define tj ∈Rd by

Z

Γ

jntj) ds = 0.

The weak convergence of (σj) causes strong convergence of tj inRd so that lim

j→∞tj =: ¯t∈Rd. (4.5)

Let ¯σ ∈Z satisfy ¯σn = ¯t on Γ. Assume ¯t = 0 first. Then, for any t/ j /= 0, we choose ¯σjZhj as in Lemma 4.2. If ¯t = 0 we choose ¯σj = 0 = ¯σ. Note ¯σj →σ¯ strongly inH as j→ ∞ because of Lemma 4.2, (H1) and (4.5).

By RdHh1/2 (cf., (H1)) we have −ϕj + cjHh1/2 for each cj ∈Rd. We define cj ∈Rd with minimal Euclid norm in Rd such that

hcj, tji = htj,V (2σjntj)i (4.6)

σ¯jn,V (σjn)−(12I + K )ϕj

− Z

F

σ¯j :C1σjdΩ.

Note that ( ¯σj,cj) are bounded inH ×H1/2. Using (4.6) we compute a(σj, ϕjj−σ¯j,−ϕj+ cj) =

Z

F

σj :C1σjdΩ +hσjntj,V (σjntj)i+hϕj,Wϕji

C

jk20,ΩF+ kϕ0jk21/2

where C >0 depends on W and the constant in Lemma 4.1 only. Since (4.4) and kσj−σ¯jk0,ΩF+ k−ϕj + cjk1/2C the above estimate proves

j, ϕj)→(0,r0) (strongly) inH ×H1/2, (σ, ϕ0) = 0; r0∈kerε. Thus, kr0k1/2 = 1.

(14)

Let ˆσ∈Z satisfy ˆσn =−r0 onΓ. By Lemma 4.2 we find a discrete stress field ˆσjZhj which, by (H1), converges towards ˆσ in (H,k·kdiv). Letting (τ, ψ) := ( ˆσj,0)∈ker Bhj \ {0}in (4.4),

a(σj, ϕj; ˆσj,0)<1

j kσˆjk0,ΩF.

By strong convergence, for j → ∞, a(0,r0; ˆσ,0) ≤0. But, by construction of σˆ and because kr0k1/2 = 1, a(0,r0; ˆσ,0) = hr0,r0i > 0. This contradiction proves the lemma. ut

Proof of Lemma 3.1. Let Zh0denote the set of functionalsΦonHh withΦ(τh) = 0 for allτhZh. By (H3), the mapping

Lh ×WhZh0 (vh, ηh) 7→ b(·;vh, ηh)

is an isomorphism [2, 3] (assuming Lh ×Wh be endowed with the mesh- dependent norm). For all v ∈ L, (H2) implies b(·;v,0)|HhZh0. Hence, for eachv∈L, there exists Ph(v) := (vh, ηh)∈Lh×Wh satisfying

b(·;vh, ηh)|Hh = b(·;v,0)|Hh. ut

Proof of Theorem 3.1. We follow [11, Proof of Theorem 3.1] and let ( ˜σ,ϕ,˜ γ) be˜ the best approximant to (σ, ϕ, γ) inHh×Hh1/2×Wh. LetHh×Hh1/2×Lh×Wh

be endowed with the norm (3.9). Because of (H3) and Lemma 4.3 we get the inf–

sup condition for the discrete spaces. With the theory of saddle point problems there exists (τh, ψh, vh, ηh) inHh×Hh1/2×Lh×Wh with

k(τh, ψh, vh, ηh)khC and

kσ˜−σhkH,h + kϕ˜−ϕhk1/2+ kPh(u)(uh, γh −γ)˜ kh

a( ˜σ−σh,ϕ˜−ϕhh, ψh)

+ b(τh; Ph(u)(uh, γh−γ)) + b( ˜˜ σ−σh;vh, ηh)

= a( ˜σ−σ,ϕ˜−ϕ;τh, ψh)

+ b( ˜σ−σ;vh, ηh)−b(τh; 0, γ−γ)˜

where we used the discrete equations and b(τh; Ph(u)(u,0)) = 0 according to Lemma 3.1. Since a and b are bounded (in the discrete norms)

kσ˜ −σhkH,h+ kϕ˜−ϕhk1/2 + kPh(u)(uh, γh−γ)˜ kh

C

kσ˜−σkH,h+ kϕ˜−ϕk1/2 + kγ˜−γk0,ΩF

.

This and the triangle inequality prove

kσ−σhkH,h+ kϕ−ϕhk1/2+ kPh(u)(uh, γh−γ)kh

(4.7)

C

τhinf∈Hh

kσ−τhkH,h+ inf

ψhHh1/2

kϕ−ψhk1/2+ inf

ηh∈Wh

kγ−ηhk0,ΩF

.

(15)

The proof of Theorem 3.1 is finished. ut

Proof of Theorem 3.2. We follow [11], assume (H4) and continue in the notations of the proof of Theorem 3.1. With Theorem 2.1 we find (Π, Ψ,z, µ) inH × H1/2×L ×W satisfying, for all (τ, ψ , v, η) inH×H1/2×L ×W ,

a(Π, Ψ;τ, ψ) + b(τ; z, µ) = 0 b(Π;v, η) =

Z

F

Ph(u)(uh, γh −γ)

·(v, η) dΩ.

Taking (τ, ψ , v, η) = (σ−σh, ϕ−ϕh,Ph(u)(uh, γh−γ)) we obtain kPh(u)(uh, γh−γ)k20,ΩF = b(Π; Ph(u)(uh, γh−γ))

= a(Π, Ψ;σ−σh, ϕ−ϕh)

+ b(σ−σh; z, µ) + b(Π; Ph(u)(uh, γh−γ))

= a(σ−σh, ϕ−ϕh;Π−Π, Ψ˜ −Ψ˜)

+ b(σ−σh; z˜z, µ−µ) + b(Π˜ −Π; P˜ h(u)(uh, γh−γ)) using Galerkin equations and the definition of Ph in Lemma 3.1 for the best approximants ( ˜Π,Ψ,˜ ˜z,µ) in˜ Hh×Hh1/2×Lh×Wh to (Π, Ψ,z, µ). Considering norms and using (4.7) forσ−σh,ϕ−ϕh and Ph(u)(uh, γh−γ) we gain

kPh(u)(uh, γh−γ)k20,ΩFC · k(Π−Π, Ψ˜ −Ψ,˜ z˜z, µ−µ)˜ kh

·(kσ˜−σkH,h+ kϕ˜−ϕk1/2+ kγ˜−γk0,ΩF).

By (H4), we have an a priori estimate of the form

k(Π−Π, Ψ˜ −Ψ,˜ z˜z, µ−µ)˜ kh ≤%(h)kPh(u)(uh, γh−γ)k0,ΩF. Using this in the former estimate and dividing by kPh(u)(uh, γh−γ)k0,ΩF we conclude the claimed estimate. ut

5. A family of elements

In the sequel we describe a family of finite element spaces due to Stenberg [11].

For each tetrahedron (resp. triangle) TTh we define a bubble function bT by bT(x ) =

Yd

i =0

λi(x ),

whereλ0, . . . , λd are the barycentric coordinates in T . By Pk(T ) we denote the space of polynomials of degree≤k on T . For d = 3 we define

Bl(T ) := {(τij) : (τi 1, . . . , τi 3) = curl (bTwi 1, . . . ,bTwi 3), wijPl(T ), i,j = 1,2,3}

Referenzen

ÄHNLICHE DOKUMENTE

So far we have shown in this section that the Gross Question (1.1) has actually a negative answer when it is reformulated for general quadratic forms, for totally singular

Figure 106: (a) Radial and (b) tangential residual stresses in the turbine disc parallel to the disc axis at a disc radius of R =100 mm determined by neutron diffraction (symbols)

Examples of analyses are shown considering a 1:1 slope reinforced by a triangular distribution of tree root blocks where the effect of both additional cohesion and block morphology

Key words: well-balanced schemes, steady states, systems of hyperbolic balance laws, shal- low water equations, evolution Galerkin schemes, finite element schemes,

Return of the exercise sheet: 14.Nov.2019 during the exercise

The analysis of the criticality of cracks in ice shelves is based on the evaluation of the stress intensity factor (SIF) at the crack tip using configurational forces.. Discretising

The main objective of this work is a comparison of a finite element (FEM) and a finite difference method (FDM) in terms of their computational efficiency when modeling shear

[r]