• Aucun résultat trouvé

2.1 The case of a cylindrical arch

N/A
N/A
Protected

Academic year: 2022

Partager "2.1 The case of a cylindrical arch"

Copied!
29
0
0

Texte intégral

(1)

Arnaud M ¨ UNCH

October 26, 2009

Abstract

We consider a circular arch of thicknessεand curvaturer−1whose elastic deforma- tions are described by a 2×2 system of linear partial differential equation. The system - of the typey00+Aεy =0,y = (y1, y3) - involves the tangentialy1 and normaly3

component of the arch displacement. We analyze in this work the null controllability of these two components by acting on the boundary through a Dirichlet and a Neu- mann control simultaneously. Using the multiplier technic we show that, for anyε >0 fixed, the arch may be exactly controlled provided that the curvature be small enough.

Then, we consider the numerical approximation of the controllability problem, using aC0−C1 finite element method and analyze some experiments with respect to the curvature and the thickness of the arch. We also highlight and discuss numerically the loss of uniform controllability as the thicknessε goes to zero, due to apparition of an essential spectrum for the limit operatorA0.

Key Words. Arch equation, Null controllability, Dirichlet and Neumann control.

Mathematics Subject Classification. 35L05, 49J20, 65K10, 65M60, 93B05.

1 Introduction

Let ω be a domain in R with boundary ∂ω, ϕ : ω → R2 a smooth function and ε > 0.

We consider an arch clamped on ∂ω, made with homogeneous and isotropic, linear elastic material, with thickness ε and whose middle surface isϕ(ω). Submitted to applied forces with resultant f (expressed in a covariant basis attached toϕ(ω)) and to initial condition (y0,y1) at time t = 0 the arch undergoes a displacement field y = (y1, y3) (expressed in the associated contravariant basis) which satisfies a system of linear partial differential equations:

(1)





y00+Aεy=f in QT =ω×(0, T), y=0, ∂νy3= 0 on ΣT =∂ω×(0, T), (y(·,0),y0(·,0)) = (y0,y1) in ω.

Laboratoire de Math´ematiques de Clermont-Ferrand, UMR CNRS 6620, Universit´e Blaise-Pascal, Cam- pus des C´ezeaux, 63177 Aubi`ere France, Email: arnaud.munch@math.univ-bpclermont.fr. Phone: (+33) - 473 407 054. Fax: (+33) - 473 407 064. Partially supported by grants ANR-07-JCJC-0139-01 (Agence national de la recherche, France) and 08720/PI/08 from Fundac´ıon S´eneca (Gobernio regional de Murcia, Spain).

1

(2)

The symbol 0 denotes the partial differentiation with respect to time and Aε a partial differential linear operator of the fourth order (depending on the mapϕand specified in the sequel). ∂νy3 designates the normal derivative ofy3.

We address in this work, on a theoretical and numerical viewpoint, the null controllability of the dynamical system (1): given anyT >0 large enough and any initial condition (y0,y1) in a suitable space, do there exists a functionv= (v1, v3) acting on the system only through the boundary as follows:

y1=v1, y3= 0, ∂νy3=v3 on ΣT

such that

(2) y(·, T) =y0(·, T) =0 in ω ?

y1andy3denotes respectively the tangential and normal components of the displacement of the arch. v1 acts on the tangential component and may be qualified as a Dirichlet control.

v3 acts on the derivative of the normal component and may be qualified as a Neumann control. Following linear arch and shell theory (we refer to [6], chapter 7), the operator Aε is decomposed as follows:

Aε =AM2 3AF

where AM andAF designate the membrane and flexural operator respectively.

In spite of several works in the eighties about the controllability of beam and plate (see [17] and the references therein), much less is known in the case of the shell and rod. In this case, due to the coupling provided by the curvature, the study is more involved and most of the existing works deal with specific examples of the mapϕ. Thus, in [9, 11], Geymonat, Loreti and Valente show the exact controllability of an axi-symmetrical spherical cap for a small enough curvature. In [10], the same authors consider also the case of an hemispher- ical shell and get controllability using harmonic analysis. For these two cases the analysis highlights the loss of exact controllability in the transition shell-membrane, i.e. when the thickness ε tends to zero. From the analytical expression of the eigenvalues of Aε for the hemi-spherical, it appears that the spectrum ofAε does not fulfill a uniform gap condition with respect toεand exhibit a minimal time of controllabilityTmin of orderO(ε−1). More- over, the operator AM has a non empty essential spectrum σess(AM) = {a}, a > 0. This non-controllability result is extended in [12] to more general operators. According to this property, relaxed exact spectral [24] or partial [16] controllability of membrane shells have been studied. Forε >0, we also mention [20] where a controllability result is obtained for a shallow Koiter’s shell using the Hilbert Uniqueness Method [18] and more recently [15]

in the context of so-called intrinsic shell model. Furthermore, on the related thematic of stabilization, let us mention [4, 14] where dissipative terms are added on the shell model in order to obtain a specific decay of the corresponding energy. The boundary stabilization of a nonlinear arch is studied in [22].

From a numerical viewpoint, few works are available for mainly two reasons. First,

(3)

independently of the controllability problem, the numerical approximation of arch and shell (by finite element method) remains difficult because of the locking phenomenon that may appear in the flexural case (we refer to [5], chapter 3). Secondly, since the pioneering work of Glowinski et al [13] (see [25] for a recent review), the exact controllability for hyperbolic systems is known to be sensitive to numerical approximation.

In this numerically oriented work, we address the exact controllability issue in the simple but rich case of a circular arch with the aim of producing some numerical experiments and discuss the sensitivity of the control with respect to the parameters, in particular the curvature and the thickness. In Section 2, we define precisely the operatorAεand state some existence results for the corresponding static formulation (Lemma 2.2). Similar qualitative results are given for the homogeneous dynamic problem (i.e. without control) in Section 3.1 (Lemma 3.1). In Section 3.2, we use the multiplier technic to obtain respectively the so-called hidden -regularity inequality (Theorem 3.1) and the observability inequality (Theorem 3.2), both involving the energy of the homogeneous solution. Then, Section 4, using the Hilbert Uniqueness Method introduced by J.-L. Lions in [18], these estimates permit to construct explicitly a control, assuming the curvature of the elastic arch small enough (see Theorem 4.1). In Section 5, we then address the numerical approximation of this control and study the influence of the curvature and thickness. The assumption on the curvature appears weak in practice (Section 5.2). Moreover, we numerically check, Sectionsec-influence-epsilon, the loss of uniform controllability as the thickness ε goes to zero. As explained, this phenomenon is due to the fact that the limit operator AM posseses an essential spectrum: precisely, 0∈σess(AM). Then, we observe that the uniform controllability is not recovered for initial datum (y0,y1) ∈ (KerAM). This is related to the non convergence of σ(Aε) toward σ(AM), enhanced by the degeneracy of the boundary conditions ony3. Section 6 concludes this work with some perspectives and Appendix 7 analyzes the spectrum ofAεand exhibits its densification overR+ asε→0, responsible of the lack of uniform spectral gap property.

2 Static problem

We explicit in this section the operatorAεin the case of a circular arch and then state some important existence and regularity results in the static case.

2.1 The case of a cylindrical arch

Without loss of generality, let us assume that the mid-surface of the arch is ω= (0,1) and note by ξ ∈ ω the curvilinear abscissa. For a cylindrical arch, the map ϕ : ω → R2 may takes the following expression

ϕ(ξ) =

rsin(r−1ξ), rcos(r−1ξ)

=x= (x1, x3)

wherer >0 denotes the constant radius of curvature of the arch. The curvature is thenr−1. Remark that the length of the arch is equal to 1 for allr >0. The tangent vectorτ =∂ξϕ

(4)

and the normal vector νtake the expression τ=

cos(r−1ξ),−sin(r−1ξ)

, ν=

sin(r−1ξ)),cos(r−1ξ)

and any displacement vector y(x) of component (ˆy1,yˆ3) in the orthonormal basis (e1,e3) ofR2 so thaty(x) = ˆy1(x)e1+ ˆy3(x)e3may then be expressed as follows

y(x) =y1(ξ)τ(ξ) +y3(ξ)ν(ξ), ξ∈ω

in the contravariant basis (τ,ν) (which coincides here with the covariant basis). The oper- atorAε is then defined byAε=AM2/3AF where

(3) AMy=a −(y1,1+r−1y3),1

r−1(y1,1+r−1y3)

!

and

(4) AFy=a 2r−1(y3,11−2r−1y1,1−r−2y3),1

y3,1111−2r−1y1,111−2r−2y3,11+ 2r−3y1,1+r−4y3

! .

adesignates the longitudinal elasticity coefficient of the arch. For simplicity, we takea= 1 in the sequel. Moreover,yi,1denotes the partial derivative ofyiwith respect to the curvilinear abcissa ξ: yi,1(ξ, t) =∂yi(ξ, t)/∂ξ,i∈ {1,3}.

Remark 1 When the curvature is assumed to be small (r−1 < 1), one may simplify the operator AF as follows:

A?Fy=a r−1(y3,1−r−1y1),11

(y3,1−r−1y1),111

! .

When the curvature is assumed to be very small (r−1<<1), one may consider simply

(5) A??F y=a 0

y3,1111

! .

The operatorAM remains unchanged in both cases. We refer for instance to [8], chapter 4.

Remark 2 The curvaturer−1is the coupling parameter between the two componentsy1and y3, so that when r−1 = 0, the components are respectively solution of the wave and beam equation

y100−y1,11= 0, y0032

3y3,1111= 0 in QT.

The first equation is null controllable for any T >1 and(y10, y11)∈L2(ω)×H−1(ω)through y1 = v1 ∈ L2T) (see [18], chapter 3). For any ε > 0, the second equation is null controllable for any T > 0 and (y03, y31) ∈ L2(ω)×H−2(ω) through ∂νy3 =v3 ∈ L2T),

(5)

taking y3= 0 onΣT (see [18], chapter 5).

Before to proceed to the analysis of the null controllability of system 1, let us insist on the following point. In the two-dimensional shell theory as presented in detail in [6, 7], the mid-surface ω belongs to R2 and systems involves three components of the displacement (u1, u2, u3). The operatorAM andAF given in that section are derived from the 2−dcase by simply eliminating the variableu2. Moreover, the operatorA=AM+2/3AF appears in the well-admitted Koiter shell model (we refer to [6, 7] for the analysis of the Koiter shell model).

2.2 Some results for the static problem

In order to fix the notations, we summarize here some important results about the static problem : findu= (u1, u3) such that

(6)

(Aεu=f in ω, u= 0, ∂νu3= 0 on ∂ω.

Definition 2.1 For anyv= (v1, v3)andu= (u1, u3), we introduce the following notation:









γ(v) =v1,1+r−1v3,

ρ(v) =v3,11−r−1v1,1−r−1γ(v) =v3,11−2r−1v1,1−r−2v3,

bM(u,v) =γ(u)γ(v), bF(u,v) =ρ(u)ρ(v), bε(u,v) =bM(u,v) +ε2

3 bF(u,v).

γ is the linearized change of metric function associated with the map ϕ. ρis the linearized change of curvature function.

From these definitions, we may rewrite the operator AM and AF defined in (3) and (4) respectively in term ofγ andρ:

(7) AMy= −γ(y),1

r−1γ(y)

!

, AFy= 2r−1ρ(y),1

ρ(y),11−r−2ρ(y)

!

Then, computations based on integrations by parts, lead to the following relations (we refer to [20] for similar developments for shell operator).

Lemma 2.1 For allu∈H3(ω)×H4(ω), we have,

• For allv∈H1(ω)×H2(ω) Z

ω

AMu·vdξ= Z

ω

bM(u,v)dξ−[γ(u)v1]10, Z

ω

AFu·vdξ= Z

ω

bF(u,v)dξ+ [ρ(u),1v3]10−[ρ(u)v3,1]10+ 2r−1[ρ(u)v1]10,

(6)

• For allv∈H01(ω)×H02(ω) Z

ω

AMu·vdξ= Z

ω

bM(u,v)dξ, Z

ω

AFu·vdξ= Z

ω

bF(u,v)dξ, therefore

Z

ω

Aεu·vdξ= Z

ω

bε(u,v)dξ.

We now introduce the space V = H01(ω)×H02(ω). As a consequence of the Poincar´e’s inequality on ϕ(ω) (see [6], chapter 4 for the counterpart Korn’s inequality for shell), the energy norm

kvkV = Z

ω

bε(v,v)dξ 1/2

is a norm over V. Therefore, the symmetric bilinear form (u,v) ∈V → R

ωbε(u,v)dξ is continuous and V-coercive so that Lax-Milgram theorem leads to the following result (see for instance [6]).

Lemma 2.2 (i) Iff ∈(L2(ω))2, there exists a unique weak solutionu∈V to the variational problem

Z

ω

bε(u,v)dξ = Z

ω

f·vdξ, ∀v∈V.

(ii) LetE= (H3(ω)∩H01(ω))×(H4(ω)∩H02(ω)). Iff ∈H1(ω)×L2(ω), the unique solution u of (6) belongs toE and satisfies the estimates

∃C >0 :kukE ≤C(kfkH1(ω)×L2(ω)+kukL2(ω)×L2(ω)).

3 Homogeneous evolution problem

The homogeneous problem plays an important role in the Hilbert Uniqueness Method. We first recall some basic results and then derive several technical equalities related to the multiplier method.

We introduce the spaceH=L2(ω)×L2(ω) endowed with the usual norm and identified to its dual space. The Hilbert space V = H01(ω)×H02(ω) is dense in H with compact imbedding. LetV0=H−1(ω)×H−2(ω) denote the dual space ofV, so thatV ⊂H⊂V0.

3.1 Existence - Uniqueness

LetT >0 be given. We consider in this section the uncontrolled system

(8)





u00+Aεu=f in QT =ω×(0, T), u=0, ∂νu3= 0 on ΣT =∂ω×(0, T), (u(·,0),u0(·,0)) = (u0,u1) in ω.

The existence, uniqueness and regularity of the solution of (8) is given by the next lemma,

(7)

direct consequence of Lions’s ”abstract” theorem (we refer to [19]) and of the regularity given by Lemma 2.2.

Lemma 3.1 (i) For all (u0,u1,f) ∈ V ×H×L2(0, T;H), there exists a unique weak solution u∈C(0, T;V)∩C1(0, T;H)that satisfies the variational problem

<u00,v>V0,V + Z

ω

bε(u,v)dξ = Z

ω

f·vdξ, a.e.t∈(0, T), ∀v∈V. Moreover, the mapping (u0,u1,f)→(u(t),u0(t))is continuous :

∃C >0;ku(t)k2V +ku0(t)k2H ≤C

ku0k2V +|u1k2H+ Z t

0

|fk2Hds

, t∈(0, T).

(ii) For all (u0,u1,f) ∈ E×V ×L2(0, T;V), there exists a unique strong solution u ∈ C(0, T;E)∩C1(0, T;V)and

∃C >0 : ku(t)k2E+ku0(t)k2V ≤C

ku0k2E+ku1k2V + Z t

0

kfk2Vds

, t∈(0, T).

Whenf =0, we denote byE the ”natural” energy of the arch E(t,u) = 1

2 Z

ω

(u0·u0+bε(u,u))dξ.

Lemma 3.2 Whenf =0, the energy of the weak solution to (8) is constant in time: E(t) = E(0), for all t∈(0, T).

Proof. First, when uis the strong solution to (8), multiplying (8) by u0 and integrating overω, we haveR

ω(u00+Aεu)·u0dξ= 1/2d/dtR

ω(u0·u0+Aεu·u)dξ= 1/2d/dtR

ω(u0·u0+ bε(u,u))dξusing the step (ii) of Lemma 2.1. Secondly, whenuis a weak solution associated with initial conditions (u0,u1) ∈ V ×H, we get the result by a density argument: let (u0k,u1k)(k>0) ∈ E×V be a sequence of Cauchy conditions and let uk be the associated strong solution such that (u0k,u1k)→(u0,u1) inV×Hask→ ∞. From the continuous de- pendence of the solution with respect to the initial data,uk→uinC(0, T;V)∩C1(0, T;H).

Hence,Ek(t,uk) = 1/2(|u0k|2H +|uk|2V)→E(t,u) = 1/2(|u0|2H +|u|2V) ask→ ∞.

In the sequel, we takef = 0 and note E0(u) =E(t= 0,u) = 1

2 Z

ω

(u1·u1+bε(u0,u0))dξ.

3.2 Observability and direct inequalities on the homogeneous prob- lem

In view of the null controllability, we establish in this section that for any T large enough, r−1small enough and any initial data (u0,u1)∈V ×H, there exist two positive constants

(8)

C1 andC2 such that

(9) C2E0(u)≤

Z T 0

bε(u,u) 1

0

dt≤C1E0(u).

Those inequalities, obtained using the multiplier technic, show that (RT

0 [bε(u,u)]10dt)1/2 defines a norm on the set of initial data. The left part of (9) is called the observability inequality associated with the adjoint system (8) and allows to show that ifusolution of (8) is such that u≡0 on ΣT, thenu≡0 inQT. In that sense, the whole solution is observed from the boundary. This inequality requires that the time be large enough, due the finite speed of propagation of the solution alongω. The right part of (9), easier to obtain, is called the direct inequality, and does not require any condition on T. The right inequality is also referred as the hidden inequality because implies extra regularity on the boundary terms.

3.2.1 Some technical lemma

The multiplier method is based on the multiplication of the conservative system (8) by suitable functions and allows to express the solutionuonQT in term of the solution on ΣT. For the sake of clarity, we collect in this section such computations leading to Proposition 3.1.

Lemma 3.3 Let ube the strong solution of (8) andq∈C1(ω). We have Z

QT

u00·qu,1dξdt=1 2

Z

QT

q,1u0·u0dξdt+ Z

ω

u0·qu,1 T

0

dξ.

Proof. Integrating by part, we have Z

QT

u00·qu,1dξdt=− Z

QT

u0·qu0,1dξdt+ Z

ω

u0·qu,1 T

0

=−1 2

Z

QT

q(u0·u0),1dξdt+ Z

ω

u0·qu,1

T

0

= 1 2 Z

QT

q,1u0·u0dξdt−1 2

Z T 0

qu0·u0

1 0

dt+ Z

ω

u0·qu,1

T 0

dξ and the result follows sinceu= 0 on ΣT =∂ω×(0, T).

We also have the following relations.

Lemma 3.4 For allq∈C2(ω)andv∈C2(ω)×C3(ω) we have

γ(qv,1) =qγ(v),1+gM(q,v), ρ(qv,1) =qρ(v),1+gF(q,v) with

gM(q,v) =q,1v1,1, gF(q,v) =q,11v3,1+ 2q,1v3,11−2r−1q,1v1,1.

(9)

Proof. From Definition 2.1, we write for allq∈C1(ω) andv∈C2(ω)×C1(ω) that γ(qv,1) = (qv1,1),1+r−1qv3,1=q,1v1,1+qv1,11+r−1qv3,1

=q(v1,1+r−1v3),1+q,1v1,1=qγ(v),1+gM(q,v) Similarly, for allq∈C2(ω) andv∈C2(ω)×C3(ω), we write

ρ(qv,1) = (qv3,1),11−2r−1(qv1,1),1−r−2qv3,1

=q,11v3,1+qv3,111+ 2q,1v3,11−2r−1(q,1v1,1+qv1,11)−r−2qv3,1

=q(v3,11−2r−1v1,1−r−2v3),1+q,11v3,1+ 2q,1v3,11−2r−1q,1v1,1

=qρ(v),1+gF(q,v).

Lemma 3.4 then implies the following one.

Lemma 3.5 For allq∈C2(w)andv∈H3(ω)×H4(ω), we have Z

ω

bε(v, qv,1)dξ=−1 2

Z

ω

q,1bε(v,v)dξ+1 2

qbε(v,v) 1

0

+ Z

ω

gε(q,v)dξ where

gε(q,v) =γ(v)gM(q,v) +ε2

3ρ(v)gF(q,v).

Proof. From the definition ofbM (see Definition 2.1) and using Lemma 3.4, we write Z

ω

bM(v, qv,1)dξ= Z

ω

γ(v)γ(qv,1)dξ = Z

ω

γ(v)(qγ(v),1+gM(q,v))dξ

=1 2

Z

ω

qbM(v,v),1dξ+ Z

ω

γ(v)gM(q,v)dξ

=−1 2 Z

ω

q,1bM(v,v)dξ+1 2

qbM(v,v) 1

0

+ Z

ω

γ(v)gM(q,v)dξ.

A similar computation for the termR

ωbF(v, qv,1)dξ then leads to the result.

Lemma 3.6 For all(u0,u1)∈E×V and allq∈C2(w), we have Z

ω

Aεu·qu,1dξ= Z

ω

bε(u, qu,1)dξ−

qbε(u,u) 1

0

.

Proof. Sinceuis a strong solution to (8), we use the step (i) of Lemma 2.1 withv=qu,1. We write

Z

ω

AMu·qu,1dξ= Z

ω

bM(u, qu,1)dξ−

γ(u)qu1,1

1

0

. Sinceu3= 0 on∂ω×(0, T), the last integral is

γ(u)q(u1,1+r−1u3) 1

0

=−

γ(u)γ(u)q 1

0

=−

bM(u,u)q 1

0

.

(10)

Similarly, Z

ω

AFu·qu,1dξ = Z

ω

bF(u, qu,1)dξ+

q

ρ(u),1u3,1−ρ(u)u3,11+ 2r−1u1,1

1 0

. Once again, sinceu3=u3,1= 0 on∂ω, the last term is [qρ(u)(−u3,11+2r−1u1,1+r−2u3)]10=

−[qρ(u)ρ(u)]10=−[qbF(u,u)]10. The result follows frombε(u,u) =bM(u,u)+ε2/3bF(u,u).

These lemmata lead to the following result.

Proposition 3.1 Let q ∈ C2(ω). For all (u0,u1) ∈ V ×H, the weak solution to (8) satisfies the identity

(10)

1 2

Z T 0

qbε(u,u) 1

0

dt=1 2

Z

QT

q,1(u0·u0−bε(u,u))dξdt+ Z

ω

u0·qu,1

T 0

dξ +

Z

QT

gε(q,u)dξdt.

Proof. As in Proposition We first consider the strong solution associated with (u0,u1)∈ E×V. Multiplying (8) by the multiplierqu,1and integrating overQT, we obtainR

QT(u00· qu,1+Aεu·qu,1)dξ = 0 and the result follows using Lemmas 3.3,3.5 and 3.6. For weak solution of (8) with the data (u0,u1)∈V×H, the result is obtained using a classical density argument. (see [18] for the wave and beam equation and [20] for the general situation of

shell operator.)

3.2.2 Direct and inverse inequality

We are now in a position to prove the right inequality of (9), the so-called direct (or hidden- regularity) inequality.

Theorem 3.1 Let ε >0.There exists a constant C1, which depends only on ω and on the map ϕ such that for all (u0,u1)∈V ×H, the weak solution to (8) satisfies the inequality

Z T 0

bε(u,u) 1

0

dt≤C1(1 +T)E0(u).

Therefore, the linear mapping (u0,u1)∈V ×H→bε(u,u)∈L1T)is continuous.

Proof. Let us takeq∈C2(ω) such thatq(0) =−1 andq(1) = 1. It follows from (10) that Z T

0

bε(u,u) 1

0

dt= Z

QT

q,1(u0·u0−bε(u,u))dξdt+ 2 Z

ω

u0·qu,1

T 0

dξ + 2

Z

QT

gε(q,u)dξdt.

The first integral is lower than kq,1kL(ω)T E0(u). Similarly, using that (R

ωbε(u,u)dξ)1/2 is a norm equivalent to kukV, the third term is lower than 2CT R

ωbε(u,u)dξ≤2CT E0(u) for some C >0. At last, the second term is less than 4CE0(u). The majoration follows.

(11)

The left inequality in (9) - the observability inequality - is obtained with the new multi- plierp= 2(ξ−ξ0),ξ0∈R, in this case Proposition 3.1 takes the following form.

Proposition 3.2 Let p= 2(ξ−ξ0), ξ0∈R. For all(u0,u1)∈V ×H, the weak solution to (8) satisfies the identity

1 2

Z T 0

pbε(u,u) 1

0

dt=T E0(u) + Z

QT

(γ(u)γ(u) +ε2ρ(u)ρ(u))dξdt +

Z

QT

hε(u)dξ+ Z

ω

u0·

pu,1+u 2

T 0

dξ.

Proof. As in Proposition 3.1, the identity is first established for Cauchy conditions (u0,u1)∈E×V corresponding to a strong solution and by the same density argument we show that it is still valid for (u0,u1)∈V ×H.

Hence, Lemma 3.3 reads Z

QT

u00·pu,1dξdt=1 2

Z

QT

2u0·u0dξdt+ Z

ω

u0·pu,1

T

0

dξ.

From the relation Z

QT

u0·u0dξdt= Z

ω

u0·u

T

0

dξ+ Z

QT

bε(u,u)dξdt, we obtain that

Z

QT

u00·pu,1dξdt=1 2

Z

QT

u0·u0dξdt+1 2

Z

QT

bε(u,u)dξdt+ Z

ω

u0·

pu,1+u 2

T

0

=T E0(u) + Z

ω

u0·

pu,1+u 2

T

0

dξ.

Lemma 3.4 now reads

γ(pv,1) =pγ(v),1+ 2v1,1=pγ(v),1+ 2γ(v) +hM(v),

ρ(pv,1) =pρ(v),1+ 4v3,11−4r−1v1,1≡pρ(v),1+ 4ρ(v) +hF(v) where hM(v) =−2r−1v3andhF(v) = 4r−1γ(v). At last Lemma 3.5 becomes

Z

ω

bε(u, pu,1)dξ= 1 2

pbε(u,u) 1

0

+ Z

ω

(γ(u)γ(u) +ε2ρ(u)ρ(u))dξ+ Z

ω

hε(u)dξ where

hε(u) =γ(u)hM(u) +ε2

3ρ(u)hF(u).

Summarizing these equalities yields the result.

Before showing the left part of (9), we need the following intermediate result :

(12)

Proposition 3.3 Let p= 2(ξ−ξ0), ξ0 ∈R, ε > 0 and r−1 >0. There exist a constant C >0 which depends onω,εand on the mapϕand a constantD which depends onεsuch that for all (u0,u1)∈V ×H, the weak solution to (8) satisfies the inequality

(11) 1

2 Z T

0

pbε(u,u) 1

0

dt≥

T(1−r−1D(ε))−2C

E0(u).

Proof. We first prove that there exists a constantC such that (12)

Z

ω

u0·(pu,1+u

2)dξ ≤CE0(u), ∀t≥0.

For any constant B >0, we have 2

Z

ω

u0·

pu,1+u 2

≤ Z

ω

Bu0·u0+ 1

B(pu,1+u

2)·(pu,1+u 2)

dξ.

Using that R

ωpu,1·udξ=−R

ωu·udξ, we obtain

Z

ω

pu,1+u 2

2

dξ≤ Z

ω

pu,1·pu,1dξ.

Moreover, from Poincar´e’s inequality, there exists a constantCK such that Z

ω

(u,1·u,1+u·u)dξ≤CK

Z

ω

bε(u,u)dξ.

Hence,

Z

ω

u0·

pu,1+u 2

≤1 2max

B, CK

kpk2L(ω)

B

E0(u).

(12) then follows with B =√

CKkpkL(ω) andC =B/2. From Proposition 3.2, the inter- mediate result is therefore that

1 2

Z T 0

pbε(u,u) 1

0

dt≥(T−2C)E0(u) + Z

QT

hε(u)dξdt.

From hε(u) = r−1γ(u)(−2u3+ 4ε32ρ(u)) we first notice that the last term vanishes for r−1 = 0 (corresponding to a flat arch) and then (11) follows with D = 0. For r−1 > 0, using the equivalence of the norm inV and the elastic energy norm (R

ωbε(u,u)dξ)1/2, there exists a constantD(ε) independent ofr such that

Z

ω

hε(u)dξ

≤r−1D(ε) Z

ω

bε(u,u)dξ.

(11) then follows.

(13)

Let us now take ξ0 ∈ ω so that p defined by p(ξ) = 2(ξ−ξ0) satisfies p(0) < 0 and p(1)>0. Then, defining

T?(r−1, ε) = 2C 1−r−1D(ε)

and assuming rlarge enough so that 1−r−1D(ε)>0, we deduce from (11) the following result:

Theorem 3.2 Letε >0 andr−1<1/D(ε). For allT > T?(r−1, ε), there exists a constant C2 = (1−r−1D(ε)) such that for all(u0,u1)∈V ×H, the weak solution to (8) satisfies the inequality

Z T 0

bε(u,u) 1

0

dt≥C2(r−1, ε)(T −T?)E0(u).

We then conclude from Theorem 3.1 and Theorem 3.2 that for all (u0,u1) ∈ V ×H, if r−1<1/D(ε) andT > T?, then the semi-norm (RT

0 [bε(u,u)]10dt)1/2is a norm overV ×H.

Similarly, the observability from the right extremityξ= 1 is obtained with the choiceξ0= 0 in p.

4 The control problem

Let V0 the dual space of V defined by V0 =H−1(ω)×H−2(ω). We show in this section that for any (y0,y1) in H×V0, there exists a control function v ∈(L2T))2 such that the solution yof

(13)





y00+Aεy=0 in QT, y1=v1, y3= 0, ∂νy3=v3 on ΣT, (y(·,0),y0(·,0)) = (y0,y1) in ω,

is at rest at time T, i.e. satisfy (2). Following [20], we first establish an existence theorem for (13) and then construct the control v.

Proposition 4.1 For any(y0,y1,v)∈H×V0×(L2T))2, there exists a unique solution y∈C(0, T;H)∩C1(0, T;V0)and the mapping(y0,y1,v)→(y(t),y0(t))is continuous:

∃C >0 :ky(t)k2H+ky0(t)k2V0≤C(ky0k2H+ky1k2V0+kvk2(L2T))2), t∈(0, T).

Proof. We refer to a detailed proof for hyperbolic equation to [18] and to [20] for shell system. For clarity, let us simply precise how the estimate is obtained. Let us first define a weak solution of (13). Let D(ω) =Cc(ω). For all (u0,u1)∈ D(ω)× D(ω), letube the unique solution of (8). Multiplying (13) by u and (8) byy, and integrating over QT, we

(14)

get, for allt∈(0, T) Z

ω

(u0(t)·y(t)−u(t)·y0(t))dξ = Z

ω

(u1·y0−y1·u0)dξ+ Z t

0

Z

ω

(Aεy·u−Aεu·y)dtdξ

= Z

ω

(u1·y0−y1·u0)dξ+ Z t

0

Bε(u,v) 1

0

dt where, from Lemma 2.1, the boundary operatorBεreads

Bε(u,v) =γ(u)v12

3ρ(u)(v3,1−2r−1v1).

Noting <u,v >=R

ωu·vdξ the duality product betweenH and H0(= H) or V and V0, the previous equality becomes

(14) <u0(t),y(t)>−<u(t),y0(t)>=<u1,y0>−<y1,u0>+ Z t

0

Bε(u,v) 1

0

dt.

Therefore,yis a weak solution to (13) if it satisfies (14). Now, according to Lemma 3.1, for (u0,u1)∈V ×H, there existsC1>0 such that

ku(t)kV +ku0(t)kH+ Z t

0

bε(u(s),u(s)) 1

0

ds 1/2

≤C1(ku0kV +ku1kH).

Moreover, from the definition of Bε, there exists a constant C2>0 such that Z t

0

Bε(u,v) 1

0

ds≤C2|v|(L2T))2

Z t 0

bε(u(s),u(s))ds 1

0

1/2

.

The estimate of Proposition 4.1 then follows from the relation (14). Finally, the regularity

onyis obtained by density argument.

We can now apply the duality HUM method (see [18]) to get the following controllability result.

Theorem 4.1 (Null controllability) Let ε > 0, r > D(ε) and T > T?. For any (y0,y1) ∈ H×V0, there exists a control v = (v1, v3) ∈ (L2T))2 such that y(·, T) = y0(·, T) = 0 inω.

Proof. We introduce the following backward system :

(15)





z00+Aεz = 0 in QT, z1=∂νu1, z3= 0, ∂νz3=∂νν2 u3 on ΣT, (z(·, T),z0(·, T)) = (0,0) in ω.

(15)

From Proposition 4.1 and from the reversibility property of (15), there exists a unique solutionz∈C(0, T;H)∩C1(0, T;V0). Moreover, since

<z0(0),u0>−<u1,z(0)>=

Z T 0

γ(u)z1

1

0

dt +ε2

3 Z T

0

(−ρ(u),1z3+ρ(u)z3,1−2r−1ρ(u)z1) 1

0

dt we get

(16) <z0(0),u0>−<u1,z(0)>=

Z T 0

Bε(u,z) 1

0

dt= Z T

0

bε(u,u) 1

0

dt.

Then, as a consequence of (9), the linear mapping Λε: (u0,u1)∈V×H→(z0(0),−z(0))∈ V0×H verifies

ε(u0,u1),(u0,u1)>=

Z T 0

bε(u,u) 1

0

dt

and is an isomorphism. Thus, for any (y0,y1) such that (−y1,y0)∈V0×H, there exists a control function v= (u1,1ν, u3,11) on ΣT and the solution yof (13) isy=zand satisfies E(T,y) = 0. Finally, the hidden regularity implies that bε(ϕ,ϕ) ∈ L1T). Therefore v212(v3,1−2r−1v1)2belongs toL1T) and sov1, v3∈L2T).

Remark 3 From the characterization (16), we obtain that the controlv= (v1, v3)obtained from Theorem 4.1 is among all the admissible controls, the one which minimizes kv12+ ε2(v3,1−2r−1v1)/3kL1T).

As a summary, we obtain the exact null controllability of the arch by acting, on the extremities, on the tangential component y1 through a Dirichlet control v1 and on the normal component y3 through a Neumann controlv3. In the next section, we address the numerical resolution of this control problem and highlight the influence of the parameter ε andr−1.

5 Numerical experiments

According to Section 4, the controllability problem is reduced, via the duality HUM method, to the solution of the following linear problem: find (u0,u1)∈V ×H such that

ε(u0,u1),(ˆu0,uˆ1)>=<(y1,−y0),(ˆu0,uˆ1)>V0×H,V×H, ∀(uˆ0,uˆ1)∈V ×H

(16)

which is equivalent to minimize the following functionalJ :V ×H→R (17)

Jε(u0,u1) =1

2 <Λε(u0,u1),(u0,u1)>−<(y1,−y0),(u0,u1)>V0×H,V×H

=1 2

Z T 0

bε(u,u) 1

0

dt+ Z

y0·u1dξ−<y1,u0>V0,V

=1 2

Z T 0

u21,12

3(u3,11−2r−1u1,1)2 1

0

dt+ Z

y0·u1dξ−<y1,u0>V0,V . Actually,Jεis the conjugate function ofJε: (L2T))2→Rdefined by Jε(v) = 1/2kv12+ ε2(v3,1−2r−1v1)/3kL1T)(see Remark 3).

SinceJεis continuous, convexe and (V ×H)-elliptic forT andrsufficiently large, the minimization can be done by a conjugate gradient (CG) algorithm (we refer to [13] for the wave equation).

In this section, we present some numerical experiments of the minimization of Jε and highlight the influence of the curvature r−1 and above all of the thicknessε. For the wave equation (we refer to [3, 13, 21] and the references therein), the numerical approximation of the exact control is very sensitive to the choice of the approximation method and to the regularity of the initial data (y0,y1). This sensitivity is likely to remain for our 2×2 system (1), at least for the Dirichlet controlv1 which formally converges to the control of the wave equation ony1whenr−1vanishes (see Remark 2). Moreover, very little seems to be known on the quality of approximation of the Neumann control.

Consequently, in order to expect a convergent approximation, we consider only regular data involving first frequency components. Secondly, we regularize the control at timet= 0 and t = T: we introduce a compact support function ρ ∈ Cc1([0, T]) - typically ρ(t) = sin2(tπ/T) - and search for a control under the form (ρv1, ρv3). These two assumptions, meaningful from a mechanical viewpoint, allow to obtain convergent and robust numerical approximations (see Section 5.1). In this respect, since the weak adjoint solutionubelongs toC(0, T;H01(ω)×H02(ω)), a conformalC0−C1finite element method is used: introducing a uniform subdivision{xi}(i=1,..,J)ofω= (0,1), we approximate V by the finite dimensional space

Vh=

(v1h, v3h)∈C0(ω)×C1(ω),(v1h)|[xi,xi+1]∈P1,(v3h)|[xi,xi+1]∈P3,∀i= 1, ..., J−1

where Pk designates the space of the polynomials of degree ≤k. The time approximation is performed in a standard way using centered finite differences : in particular, we use a Newmark method to get an unconditionally stable scheme. The introduction of aC1-finite element as well as the presence of a Neumann control is a major difference and new ingredient with respect to the numerical literature mainly devoted to the wave equation with Dirichlet boundary condition.

Two stopping criteria for the CG algorithm may be used: the first one is related to the residualr(k)at the iterationk, namelyr(k)≤ηr(0),η >0, permits to analyze the quality of the numerical approximation. The second one is related to the numerical energy Eh of the

(17)

system, namelyEh(T)≤ηEh(0), permits to analyze the efficiency of the control.

Remark 4 A key point in the conjugate gradient algorithm is the choice of the scalar product on V ×H: in this respect, we have replaced the usual scalar product

(v0,v1),(˜v0,v˜1)

V×H

= Z

ω

(v1,10 ˜v01,1+v03,11˜v3,110 )dξ+ Z

ω

(v11˜v11+v13˜v31)dξ by the more natural one :

(v0,v1),(˜v0,v˜1)

V×H

= Z

ω

bε(v0,v˜0)dξ+ Z

ω

(v1111+v1331)dξ.

Remark 5 One may improve the robustness of the algorithm (in particular reduce the num- ber of iterations) by coupling the C0−C1 finite element scheme with a Bi-grid method: in- troduced in [13] for the wave equation, it consists in computing at each iteration the descent direction on a coarse mesh. We have observed that the Bi-grid method remains efficient in our context provided a careful extension (from the coarse to the fine mesh) of the nodal C1 finite element approximation of y3. The gain of robustness is very valuable when we are at the theoretical limit of controllability (see Section 5.3).

5.1 Analysis with respect to h

Before highlighting the influence of the parametersεandr−1on the controllability, it is es- sential to analyze our numerical approximation with respect to the discretization parameter handdt.

Let us first consider the following data : ε= 10−1, r−1= 10−1,T = 3. We assume that the control is acting on the left extremityξ= 0 ofω= (0,1) and we note Σ0T ={0} ×(0, T).

On the right extremityξ= 1, the arch is clamped. We start with the following simple initial condition (in the local curvilinear frame) (y0,y1)∈H×V0:

(18) y0= (y01, y30) = (sin(πξ),sin2(πξ)); y1= (y11, y31) = (0,0), ξ∈ω= (0,1).

We use the stopping criterion related to the residualr(k)withη = 10−8and takeh= 1/64 and ∆t=h/2. Figure 1 depicts the initial condition of the system (8), minimum inV ×H of the associated functional J (see 17). Figure 2-left represents the evolution with respect to time of the associated control v = (v1, v3) ∈ (L20T))2. As explained previously, the control and its first derivative are zero at time t= 0 andt =T. Figure 2-right represents the evolution of the energy (non monotonous). We check the (numerical) controllability by computing the ratio Eh(T)/Eh(0)≈4.75×10−5. Figure 3 represents the evolution of the controlled tangential y1 and normal y3 displacement on QT (attached to the local basis).

Finally, for r−1 = 1/5 and r−1 = 1/10, Figure 4 represents the evolution of the circular arch submitted to the initial position y0(x) =y01τ(ξ) +y30ν(ξ) and to the initial velocity y1(x) =y11τ(ξ) +y31ν(ξ).

We then collect in Table 1 the evolution of several relevant quantities with respect toh=

(18)

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

!0.015

!0.01

!0.005 0 0.005 0.01 0.015

x

u01 u03

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

!0.4

!0.35

!0.3

!0.25

!0.2

!0.15

!0.1

!0.05 0

x

u11 u13

Figure 1: T = 3 - Initial conditionsu0= (u01, u03) (Left) andu1 = (u11, u13) (Right) of the system (8) and minimum of the functional J (defined by eq. 17).

0 0.5 1 1.5 2 2.5 3

!5

!4

!3

!2

!1 0 1 2

t

v1 v3

0 0.5 1 1.5 2 2.5 3

0 0.5 1 1.5 2 2.5 3

t

E(t)

Figure 2: T = 3 -Left- Compact support control (v1, v3) = (∂νu1, ∂νν2 u3) on Σ0T; Right- Energy Eh(t,y) vst.

0 0.5

1 1.5

2 2.5

3

0 0.2 0.4 0.6 0.8 1

!0.8

!0.6

!0.4

!0.2 0 0.2 0.4 0.6 0.8 1

x T

0 0.5

1 1.5

2 2.5

3

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

x T

Figure 3: T = 3 - Controlled displacementy1(Left) andy3(Right) onQT.

Références

Documents relatifs

Thus each mathematical symbol (G, L,. .) has a single definite meaning inside each exercise, but might have different meanings from an exercise to another.. It is not required to

More pre- cisely, we are devoted to proving how standard arch models may be obtained as singular limit of a 1-D Marguerre-Vlasov system - with respect to a small param- eter ε -,

In section 4, we point out the sensibility of the controllability problem with respect to the numerical approximation, and then present an efficient and robust semi-discrete scheme

In the general case, the discrete Laplace operator applied to the standard interpolation of a regular function is not consistent with the continuous Laplace operator applied to

We first present the level set method, and then we derive from the previous analysis an iterative scheme which converges provided that the initial guess is not too far from

We show that the Maxwell equations in the low frequency limit, in a domain composed of insulating and conducting regions, has a saddle point structure, where the electric field in

We will focus our study on the approximation of Problem 1.1 with piecewise linear éléments, including numerical intégration, and obtain rates of convergence.. The layout of the paper

In this paper, we study the dependence of the buckling load of a non- shallow arch with respect to the shape of the midcurve.. The model we consider is the simple Eulerian buckling