• Aucun résultat trouvé

K´ evin Le Balc’h and Marius Tucsnak

N/A
N/A
Protected

Academic year: 2022

Partager "K´ evin Le Balc’h and Marius Tucsnak"

Copied!
30
0
0

Texte intégral

(1)

ESAIM: COCV 27 (2021) 17 ESAIM: Control, Optimisation and Calculus of Variations

https://doi.org/10.1051/cocv/2021008 www.esaim-cocv.org

A PENALTY APPROACH TO THE INFINITE HORIZON LQR OPTIMAL CONTROL PROBLEM FOR THE LINEARIZED

BOUSSINESQ SYSTEM

K´ evin Le Balc’h and Marius Tucsnak

*

Abstract. In this paper, we consider the infinite time horizon LQR optimal control problem for the linearized Boussinesq system. The goal is to justify the approximation by penalization of the free divergence condition in this context. We establish convergence results for optimal controls, optimal solutions and Riccati operators when the penalization parameter goes to zero. These results are obtained under two different assumptions. The first one treats the linearization around a sufficiently small stationary state and an arbitrary control operator (possibly of finite rank), while the second one does no longer require the smallness of the stationary state but requires to consider controls distributed in a subdomain and depending on the space variable.

Mathematics Subject Classification.49N10, 93B05.

Received November 29, 2020. Accepted January 13, 2021.

Dedicated to Enrique Zuazua for his 60th birthday.

1. Introduction

The optimal control of the Boussinesq system and of its linearization around a stationary state are matters of great interest in various applications fields, such as designing and exploiting energy efficient buildings, (see, for instance, [9, 10, 23, 33]). A difficulty which has to be handled in solving this type of problems is that, due to the free divergence condition for the velocity field and to the presence of the pressure, the governing equations cannot be written as a well-posed control system in the sense of Salomon-Weiss (as described, for instance, in Curtain and Weiss [12]), but merely as an infinite dimensional descriptor system (see, for instance, Reis [26]). This feature implies several difficulties when simulating or tackling the associated optimal control problems, namely in developing efficient methods to solve the Riccati equations appearing in LQR problems.

More precisely, after semidiscretization with respect to the space variable, the original problem becomes a LQR problem for a finite dimensional differential algebraic system, which, although projection based efficient numerical methods have been recently developed in the literature, see (for instance, [8, 18, 25]), is in general very difficult from the computational viewpoint.

In this paper we study theinfinite horizonLQR problem for a system described by the linearized Boussinesq equations with a bounded control operator, by adopting a different methodology, based on the penalization

Keywords and phrases:Linear quadratic optimal control, Riccati theory, Boussinesq system, penalty method, controllability.

Institut de Math´ematiques de Bordeaux, 351 Cours de la Lib´eration, 33400 Bordeaux, France.

* Corresponding author:marius.tucsnak@u-bordeaux.fr

c

The authors. Published by EDP Sciences, SMAI 2021

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

(2)

of the free divergence condition. Within this approach, denoting the velocity field of the fluid by v and the pressure field by p, the free divergence condition div v = 0 is replaced by divv+εp= 0. As described in the next section, this procedure yields a standard well-posed linear system (governed by a PDE of parabolic type with distributed control). For simulation purposes this approach has been widely used for Stokes, Navier-Stokes or Boussinesq type system. We refer to Temam [31] for the first analysis of the method (for the Navier-Stokes equations) and to Hebeker et al.[17] and Shen [27, 28] for error estimates in the case of non stationary Stokes and Navier-Stokes systems. From an optimal control perspective, the penalty method has been applied in Badra, Buchot and Thevenet [2] for the finite horizon LQR problem where the authors prove convergence results for the penalized Oseen system with boundary control.

To give a more precise view on the questions in the present work, consider the Boussinesq system in a bounded domain Ω⊂Rd, withd∈ {2,3}, whose boundary is denoted by Γ :=∂Ω. It can be written as









tv−ν∆v+ (v· ∇)v+∇p=yed+f +PB˜u˜ in (0,∞)×Ω,

ty−α∆y+v· ∇y=g+Bd+1ud+1 in (0,∞)×Ω,

divv= 0 in (0,∞)×Ω,

v= 0, y= 0 on (0,∞)×Γ,

v(0,·) =v0, y(0,·) =y0 in Ω.

(1.1)

In the above equations,v is the velocity field of the fluid,pis the pressure field,y stands for the temperature field and {e1, e2, . . . , ed} is the standard basis ofRd. Moreover,ν >0 is the viscosity parameter, α >0 is the heat conductivity in the fluid. The terms f : Ω→Rd and g: Ω→R describe the influence of field forces and possible internal heat sources. The operatorP :

L2(Ω)d

→V0(Ω), with (denoting bynthe unit normal to∂Ω oriented towards the exterior of Ω)

V0(Ω) ={v∈L2(Ω)d ; divv= 0, v·n= 0 on∂Ω} (1.2) is the Leray projector (see, for instance Sohr [29], Sect. II.2.5). The (bounded) control operatorB := ( ˜B, Bd+1) will be required to satisfy various assumptions, which will be described below, whereas u:= (˜u, ud+1) is an input function, taking its values in a Hilbert space U, called the space of controls.

In this work we consider the linearization of the system (1.1) around a stationary solution (vs, ps, ys). More precisely, (vs, ps, ys) is supposed to satisfy the equations





−ν∆vs+ (vs· ∇)vs+∇ps=ysed+f in Ω,

−α∆ys+vs· ∇ys=g in Ω,

div vs= 0 in Ω,

vs= 0, ys= 0 on Γ,

(1.3)

where the source terms f andg coincide with those in (1.1). It is not difficult to see that the linearization of (1.1) around a stationary state satisfying (1.3) is









tv−ν∆v+ (v· ∇)vs+ (vs· ∇)v+∇p=yed+PB˜u˜ in (0,∞)×Ω,

ty−α∆y+vs· ∇y+v· ∇ys=Bd+1ud+1 in (0,∞)×Ω,

div v= 0 in (0,∞)×Ω,

v= 0, y= 0 on (0,∞)×Γ,

v(0,·) =v0, y(0,·) =y0 in Ω.

(1.4)

(3)

The goal of this article is to study the approximation of an infinite horizon LQR optimal control problem for (1.4) by the corresponding optimal control problem for the penalized system









tvε−ν∆vε+ (vε· ∇)vs+ (vs· ∇)vε+∇pε=yεed+ ˜Bu˜ in (0,∞)×Ω,

tyε−α∆yε+vs· ∇yε+vε· ∇ys=Bd+1ud+1 in (0,∞)×Ω,

div vε+εpε= 0 in (0,∞)×Ω,

vε= 0, yε= 0 on (0,∞)×Γ,

vε(0,·) =v0, yε(0,·) =y0 in Ω,

(1.5)

whereε >0 is a penalization parameter. Note that the Leray projectorP does not appear in (1.5) because the velocity field is not divergence free. By eliminating the pressure in (1.5), we can reformulate the above equations as the following parabolic system:





tvε−ν∆vε+ (vε· ∇)vs+ (vs· ∇)vε1ε∇div vε=yεed+ ˜Bu˜ in (0,∞)×Ω,

tyε−α∆yε+vs· ∇yε+vε· ∇ys=Bd+1ud+1, in (0,∞)×Ω,

vε= 0, yε= 0 on (0,∞)×Γ,

vε(0,·) =v0, yε(0,·) =y0 in Ω.

(1.6)

For v0

y0

∈H :=V0(Ω)×L2(Ω) and u∈L2([0,∞);U), we introduce the quadratic cost functional

J

u;

v0

y0

= Z +∞

0

ku(t)k2U+

v(t,·) y(t,·)

2

H

!

dt, withv, p, ysubject to (1.4). (1.7)

Under suitable assumptions, to be described later, the functional J admits a unique minimum for u=uopt∈ L2([0,∞);U). The corresponding solution of (1.4) is denoted by

 vopt popt yopt

 and we have vopt

yopt

∈C([0,∞);H).

Similarly, for v0

y0

∈X :=L2(Ω)d+1 andu∈L2([0,∞);U), we introduce the cost quadratic functional

Jε

u;

v0 y0

= Z +∞

0

ku(t)k2U+

vε(t,·) yε(t,·)

2

X

!

dt, with vε

yε

subject to (1.6). (1.8)

Under suitable assumptions, the functional Jε admits a unique minimum for u=uopt,ε ∈L2([0,∞);U), with the corresponding state trajectory

vopt,ε

yopt,ε

∈C([0,∞);X).

The main results of this work assert that, under appropriate assumptions, for every v0

y0

∈H we have

ε→0limuopt,ε=uopt and lim

ε→0

vopt,ε

yopt,ε

= vopt

yopt

, (1.9)

in the sense which will be made precise in Theorems 3.1and3.2.

Control problems with an approximation by penalization of the free divergence condition have been studied in different ways. From a controllability point of view, the null-controllability of the Stokes system, respectively Navier-Stokes, is established in [19], respectively in [1], with as many controls as equations. More recently, [6]

proves the null-controllability of the Stokes system in 2D with only one control, this result having been extended

(4)

in [7] for the (nonlinear) Boussinesq system in 2D and 3D for cylindrical geometries. For LQR optimal control problems, up to our knowledge, the only article in the literature seems to be [2] where the authors prove results in the spirit of (1.9) for the Oseen system, with boundary control, and for a finite time horizon. We emphasize that the error estimates provided in [2] explode when the horizon time goes to +∞.

The main novelty brought in by this work consists in obtaining convergence results for infinite time horizon LQR problems. An essential ingredient to accomplish this goal is a version of the strategy proposed in Banks and Kunisch [5] for Galerkin approximations, see also Banks and Ito [4]. More precisely, the method essentially consists in the three main steps described below.

1. A question which is common to the various situations we consider is proving that the Riccati opera- tor associated to (1.8) is bounded, uniformly with respect to ε and that the optimal trajectories decay exponentially, uniformly with respect toε. This will be shown to hold under two different assumptions.

The first one considers an arbitrary bounded control operatorB (possibly of finite rank) and assumes a smallness condition for the stationary state. This is first done for the case of a vanishing stationary state and then by a perturbation argument.

The second situation considered in this work treats the case of a control distributed in a subdomainO ⊂Ω.

In this case the control space is infinite dimensional, which enables us to derive uniform null-controllability results for (1.4) and (1.6). This task is accomplished by combining Carleman estimates for the penalized Stokes system due to [1] and Carleman estimates for the heat equation due to [14].

2. Secondly, we prove (1.9) for finite time horizon problems. In order to do this, we first establish convergence results for solutions of (1.6) to solutions of (1.4), in the limitε→0 and in compact time intervals. This is obtained through an adaptation of the Trotter-Kato theorem and the use of Lax-Phillips semigroups.

Then, using the well-known formulas for optimal controls and optimal trajectories, given for instance in [15], we can pass to the limit to deduce (1.9).

3. Finally, by using the uniform bounds and decays that are established in the first point of the strategy, we can approximate the infinite time horizon problem in the interval [0,∞) by finite time horizon problems in intervals [−s,0] with s >0. By gathering this approximation and the fact that (1.9) holds in [−s,0], we obtain the expected convergence results (1.9) for the infinite time horizon problem.

The paper is organized as follows. In Section 2, we introduce a functional analytic framework to establish the well-posedness of the control systems (1.4) and (1.6) and we recall some basic facts about LQR problems in infinite horizon. In Section 3 we state the main results of the paper, which basically formalize (1.9) under appropriate assumptions. Section 4 mainly relies on convergence results (in the limit ε→0+) and uniform bounds (inε >0) for semigroups associated to penalized problems. In Section5, we prove uniform stabilizability results and null-controllability results for penalized systems. With all these preliminary results at hand, Section6 is dedicated to the proof of the main results of the paper. Finally, Section 7contains some final comments and open questions.

2. Semigroup formulation and background on LQR problems

For the remaining part of this work we assume that the stationary solution

 vs ps ys

of (1.3) verifies

vs

ys

∈W1,∞(Ω)d+1, ps∈L(Ω). (2.1)

(5)

2.1. Well-posedness of the controlled systems

We introduce below a semigroup formulation of (1.4) and (1.6). To this aim we introduce the spaces

V1(Ω) ={v∈H01(Ω)d ; divv= 0}, (2.2)

D(A0) =

H2(Ω)d+1

∩[V1(Ω)×H01(Ω)]. (2.3)

Recalling thatH =V0(Ω)×L2(Ω) (withV0(Ω) introduced in (1.2)), we consider the operatorA0:D(A0)→H defined by

A0

ϕ ζ

=

νP∆ϕ−(vs· ∇)ϕ−(ϕ· ∇)vs+ζed α∆ζ−(vs· ∇ζ)−(ϕ· ∇ys)

((ϕ, ζ)∈ D(A0)), (2.4) where P is the Leray projector. Equivalently,A0 can be defined by duality, by setting:

A0 ϕ

ζ

, ψ

η

H

=ν Z

∇ϕ:∇ψdx+ Z

[(vs· ∇)ϕ+ (ϕ· ∇)vs]·ψdx− Z

ζψddx +α

Z

∇ζ· ∇ηdx+ Z

[(vs· ∇)ζ+ϕ· ∇ys]·ηdx

ϕ ζ

, ψ

η

∈ D(A0)

. (2.5) We will need the result below, for which we sketch the proof, for the sake of completeness, with no claim of originality.

Proposition 2.1. The operator A0 defined above generates an analytic semigroup on H, denoted by T0 = T0t

t>0.

Proof. By integration by parts and Young’s inequalities, it is not difficult to see that there exists a positive constant cdepending on

vs

ys

W1,∞(Ω)d+1

such that for every ϕ

ψ

∈ D(A0),

A0 ϕ

ζ

, ϕ

ζ

H

>νk∇ϕk2L2(Ω)d+αk∇ζk2L2(Ω)d−ck(ϕ, ζ)k2X,

where we recall the notationX =

L2(Ω)d+1

. Thus the corresponding sesquilinear form associated to−A0+cI (this form is simply defined by the right hand side of (2.5)) is coercive onH. We also readily check that this form is densely defined, closed and continuous on H. By Theorem 1.52 of [21], A0−cI generates an analytic semigroup onH and finally, by a perturbation argument (see, for instance, [22], Sect. 3.2, Cor. 2.2)A0generates an analytic semigroup onH.

For every ε >0, consider the operatorAε:D(Aε)→X defined by D(Aε) =

H2(Ω)∩H01(Ω)d+1

, (2.6)

Aε

ϕ ζ

=

ν∆ϕ−(vs· ∇)ϕ−(ϕ· ∇)vs+1ε∇(div ϕ) +ζed

α∆ζ−(vs· ∇ζ)−(ϕ· ∇ys)

ϕ ζ

∈ D(Aε)

. (2.7)

(6)

Alternatively, Aεcan be defined by

Aε

ϕ ζ

, ψ

η

X

=ν Z

∇ϕ:∇ψdx+ Z

[(vs· ∇)ϕ+ (ϕ· ∇)vs]·ψdx +1

ε Z

(divϕ)(divψ) dx− Z

ζψddx+α Z

∇ζ· ∇ηdx +

Z

[vs· ∇ζ+ϕ· ∇ys]·ηdx

ϕ ζ

, ψ

η

∈ D(Aε)

. (2.8)

We have the following result.

Proposition 2.2. For every ε >0 the operator Aε defined above generates an analytic semigroup on X = L2(Ω)d+1

, denoted byTε= (Tεt)t>0. Moreover, there existM >1 andω∈R such that

kTεtkL(X;X)6Meωt (ε >0, t>0). (2.9) Proof. By integration by parts and Young’s inequalities, it is not difficult to see that there exists a positive constant cdepending onk(vs, ys)kW1,∞(Ω)d+1 (independent ofε >0), such that for every

ϕ ψ

∈ D(Aε),

Aε ϕ

ζ

, ϕ

ζ

X

>νk∇ϕk2L2(Ω)d+1

εkdivϕk2L2(Ω)+αk∇ζk2L2(Ω)d−ck(ϕ, ζ)k2X. (2.10) Thus the sesquilinear form associated to −Aε+cI is coercive on X. We also readily check that this form is densely defined, closed and continuous on X. By Proposition 5.1 of [21] and Theorem 1.52 of [21], Aε−cI generates an analytic semigroupTfε satisfying

Tfεt

L(X;X)6Meωt (ε >0, t>0).

Since the constant cis independent of εwe can apply ([22], Sect. 3.2, Cor. 2.2), to obtain that for every ε >0 the operatorAεgenerates an analytic semigroup onX and that we have (2.9).

Let us introduce the projectorPH∈ L(X;H) defined by PH

v y

= P v

y

v y

∈X

,

where P is the Leray projector. Let U be a Hilbert space (the space of controls). We introduce the control operators

B∈ L(U;X) andB0:=PHB∈ L(U;H). (2.11) We have the following well-posedness result for controlled systems.

Proposition 2.3. For every v0

y0

∈H,u∈L2([0,∞);U), the Cauchy problem v˙

˙ y

=A0

v y

+B0u,

v(0) y(0)

= v0

y0

, (2.12)

(7)

admits an unique mild solution v

y

defined by

v(t) y(t)

=T0t

v0

y0

+

Z t 0

T0t−σB0u(σ) dσ (t>0). (2.13)

Similarly, for every v0

y0

∈X,u∈L2([0,∞);U), the Cauchy problem v˙

˙ y

=Aε

v y

+Bu,

v(0) y(0)

= v0

y0

, (2.14)

admits an unique mild solution (v, y), defined by v(t)

y(t)

=Tεt

v0 y0

+ Z t

0

Tεt−σBu(σ) dσ (t>0). (2.15)

Proof. The proof readily follows from Propositions 2.1 and2.2 and standard semigroup theoretic results (see, for instance, [32], Prop. 4.2.5).

2.2. Some background on LQR optimal control problems

We recall below some basic facts on LQR problems in an infinite dimensional context.

Let A be the generator of a strongly continuous semigroup on a Hilbert space X and B ∈ L(U;X) be a bounded control operator on a Hilbert spaceU. The LQR theory in infinite horizon focus on trajectories of the linear control system x0 =Ax+Bu, x(0) =x0, which minimize the quadratic cost functional

J(u;x0) = Z +∞

0

ku(t)k2U+kx(t)k2X

dt (u∈ U). (2.16)

We recall below the well-known notion of stabilizability.

Definition 2.4. The pair (A,B) is stabilizable if there exists K ∈ L(X,U) such that A − BK generates an exponentially stable semigroup FtK

t>0 onX,i.e.there existM >0 andω >0 such that FtK

L(X;X)6M e−ωt (t>0).

One of the main results of the LQR theory states that the optimal control associated to (2.16) is given by a feedback law, see, for instance, Theorem 2.1 of [5] and Theorem 4.2 of [15].

Theorem 2.5. If the pair(A,B)is stabilizable then for everyx0∈ X,J(·;x0)admits a unique minimum uopt

given by

uopt(t) =−BPFtx0 (t∈[0,∞)). (2.17)

where P∈ L(X;X), is the unique nonnegative self-adjoint solution of the Riccati equation

AP+PA − PBBP+IX = 0, (2.18)

(8)

and F = (Ft)t>0 is the exponentially stable semigroup on X generated by A − BBP. Moreover, for every x0∈ X,

min

u∈L2([0,+∞);U)J(u;x0) =hPx0, x0i. (2.19) We recall that P ∈ L(X;X) is a solution to the algebraic Riccati equation (2.18) ifP mapsD(A) toD(A) and (2.18) is satisfied when the left hand side is applied to an arbitraryx∈ D(A).

3. Main results

In this section we continue to use the notation introduced in Section2 for the spaces H, X, U and for the operators (Aε)ε>0, B0 and B (we recall, in particular, from (2.11) that B0 =PHB). Moreover, we introduce some new notation and we state the main results of the paper.

Assume that (A0, B0), withA0defined in (2.4) andB0defined in (2.11), is stabilizable. By Theorem2.5, let us denote by Π0∈ L(H) the unique nonnegative self-adjoint solution of

A0Π0+ Π0A0−Π0B0B0Π0+IH = 0. (3.1) In (3.1), IH denotes the identity operator on H. The feedback semigroup, i.e.generated by A0−B0B0Π0 is denoted by F0t

t>0, whereas the optimal control associated to an initial data v0

y0

∈H is given byuopt,0(t) =

−B0Π0F0t

v0 y0

.

In the same way, assume that (Aε, B) is stabilizable with Aεdefined in (2.7) andB defined in (2.11). Then let us denote by Πε∈ L(X) the unique nonnegative self-adjoint solution of

AεΠε+ ΠεAε−ΠεBBΠε+IX= 0. (3.2) In (3.2), IX denotes the identity operator on X. The closed loop semigroup, i.e.generated by Aε−BBΠε, is denoted by (Fεt)t>0 and the optimal control, associated to an initial data

v0

y0

∈X is given byuopt,ε(t) =

−BΠεFεt

v0 y0

.

The main results of the article focus on asymptotic properties of Πε, (Fεt)t>0 anduopt,ε in the limitε→0.

The first main result of the paper treats the case of a stationary state vs

ys

“sufficiently small”, with an arbitrary control spaceU and an arbitrary bounded control operator (possibly of finite rank).

Theorem 3.1. There existsδ >0such that if

vs ys

W1,∞(Ω)d+16δ, (3.3)

then A0 is exponentially stable and Aε is uniformly (with respect to ε) exponentially stable. Moreover, there exist M >0 andω >0, independent of εsuch that the solution

εkL(X;X)6M (ε>0), (3.4)

kFεtkL(X;X)6M e−ωt (ε, t>0), (3.5)

(9)

In addition, for every v0

y0

∈H,

ε→0+lim

Πε

v0

y0

−Π0

v0

y0

H

= 0 (t>0), (3.6)

ε→0+lim Fεt

v0

y0

−F0t

v0

y0

H

= 0 (t>0), (3.7)

ε→0+lim kuopt,ε(t)−uopt,0(t)kU = 0 (t>0), (3.8) and the last two convergences are uniform with respect to ton compact intervals.

Our second main result does no longer require the smallness of (vs, ys), considering a particular class of control operators.

Theorem 3.2. Assume thatU =X and letB∈ L(U, X) be defined by

Bu=u1O (u∈U), (3.9)

where O is an open subset of Ω and 1O is the characteristic function of O. Then (A0, B) is exponentially stabilizable and there exists ε0 >0 such that (Aε, B) is uniformly, with respect to ε∈(0, ε0), exponentially stabilizable. The uniform bounds (3.4) and (3.5) still hold true for ε∈ (0, ε0). Moreover, the corresponding Riccati operators, closed loop generators and optimal LQR controls still satisfy (3.6),(3.7)and (3.8).

4. Preliminary results on the (penalized) Boussinesq semigroup

In this section and the following ones we continue to use the notation H := V0(Ω)×L2(Ω) and X = L2(Ω)d+1

for the state space of the system described by the Boussinesq equations and their penalized version, respectively. We recall that the operatorA0:D(A0)→H has been defined in (2.4), whereas for everyε >0 the operator Aε has been defined in (2.7). As in the previous sections, the analytic semigroup on H generated by A0is denoted by T0, whereas, for everyε >0,Tεstands for the analytic semigroup on X generated byAε. 4.1. Convergence results for the semigroups describing the free dynamics

In this subsection we prove that whenε→0+ the family of semigroups (Tε)ε>0strongly converges toT0(and the same for adjoint semigroups). The proof relies on an adaptation of the Trotter-Kato’s theorem, described below.

We begin by noticing that the adjointA0 ofA0 is defined byD(A0) =D(A0) and A0

ϕ ζ

=

ν∆ϕ+ (∇ϕ)trvs+ (∇ϕ)vs−ζ∇ys

α∆ζ+ (vs· ∇ζ) +ϕd

ϕ ζ

∈ D(A0)

. (4.1)

In the same way, for every ε >0 the adjointAε ofAεis defined byD(Aε) =D(Aε) and Aε

ϕ ζ

=

ν∆ϕ+ (∇ϕ)trvs+ (∇ϕ)vs−ζ∇ys+1ε∇(div ϕ) α∆ζ+ (vs· ∇ζ) +ϕd

ϕ

ζ

∈ D(Aε)

. (4.2)

In (4.1), (4.2), the notation (∇ϕ)tr means the transpose of the matrix∇ϕ.

(10)

We first prove the convergence of the resolvents of Aε (respectively Aε) towards the resolvents of A0

(respectively A0).

Proposition 4.1. There existsλ0>0 such that for everyλ∈CwithReλ>λ0, we have

ε→0lim

[λI−Aε]−1 f

g

−[λI−A0]−1PH

f g

X

= 0

f g

∈X

, (4.3)

ε→0lim

[λI−Aε]−1 f

g

−[λI−A0]−1PH

f g

X

= 0

f g

∈X

. (4.4)

Proof. We only prove (4.3), since the proof of (4.4) is fully similar.

Let f

g

∈X and

λ0> c >0, (4.5)

where the constant cis the one appearing in (2.10). Then, forλ∈Cwith Reλ>λ0, by setting ϕε

ζε

= (λI−Aε)−1 f

g

,

and by using (2.8) we have λ

Z

ϕε·ψdx+ν Z

∇ϕε:∇ψdx+ Z

[(vs· ∇)ϕε+ (ϕε· ∇)vs]·ψdx +1

ε Z

(divϕε)(divψ) dx− Z

ζεψddx+λ Z

ζεηdx+α Z

∇ζε· ∇ηdx +

Z

[vs· ∇ζεε· ∇ys]·ηdx= Z

f·ψdx+ Z

gηdx

ψ η

∈ D(Aε)

. (4.6)

Taking ψ

η

= ϕε

ζε

in (4.6) and using (2.10) we obtain

k∇ϕεk2L2(Ω)d+1

εkdiv ϕεk2L2(Ω)+k∇ζεk2L2(Ω)d 6c

f g

2

X

(ε >0),

wherec >0 is some another constant. From this and the Poincar´e inequality it follows that there exists (another) c >0 with

εk2H1

0(Ω)d+1

εkdiv ϕεk2L2(Ω)+kζεk2H1 0(Ω)6c

f g

2

X

(ε >0).

The above estimate implies that there exists ϕ

ζ

∈H01(Ω)d+1 such that ϕε

ζε

→ ϕ

ζ

asε→0 inH01(Ω)d+1 weakly,

(11)

ϕε ζε

→ ϕ

ζ

as ε→0 inL2(Ω)d+1 strongly,

divϕ= 0.

For ψ

η

∈ D(A0), note that divψ= 0, we can thus pass to the limit in (4.6) to obtain that

λ Z

ϕ·ψdx+ν Z

∇ϕ:∇ψdx+ Z

[(vs· ∇)ϕ+ (ϕ· ∇)vs]·ψdx

− Z

ζψddx+λ Z

ζηdx+α Z

∇ζ· ∇ηdx +

Z

[vs· ∇ζ+ϕ· ∇ys]·ηdx= Z

f·ψdx+ Z

gηdx,

ψ η

∈ D(A0)

, (4.7)

then using

Z

f·ψdx+ Z

gηdx= Z

PH

f g

· ψ

η

dx

ψ η

∈ D(A0)

, (4.8)

we deduce (4.3) from (4.7).

Note that, arguing as in Proposition 2.1 and in Proposition 2.2, the operator A0 generates an analytic semigroup on H, denoted by (T0) = (T0t)

t>0 and for every ε >0 the operator Aε generates an analytic semigroup on X, denoted by (Tε)= ((Tεt))t>0. Moreover, there existM >1 andω∈Rsuch that

k(Tεt)kL(X;X)6Meωt (ε >0, t>0). (4.9) An important consequence of Proposition4.1is the following result:

Proposition 4.2. For every v0

y0

∈H, we have

ε→0lim Tεt

v0

y0

−T0t

v0

y0

X

= 0 (t>0), (4.10)

ε→0lim

(Tεt) v0

y0

−(T0t) v0

y0

X

= 0 (t>0), (4.11)

uniformly with respect to ton compact intervals.

Proof. We only prove (4.10). The proof of (4.11) follows the same scheme, using (4.4) instead of (4.3).

We follow Section 3.4, Theorem 4.2 of [22]. We takeλ>λ0 where λ0 is defined in Proposition 4.1 and we note

Rελ= [λI−Aε]−1, R0λ= [λI−A0]−1.

(12)

By using that the resolvent commutes with the associated semigroup, see Section 1.2, Theorem 2.4, c) of [22], we have

(Tεt−T0t)R0λz0

X6

Tεt(R0λ−Rλε)z0

X+

Rελ(Tεt−T0t)z0

X+

(Rελ−R0λ)T0tz0

X

=:Dε1z0+Dε2z0+Dε3z0 (ε >0, t∈[0, T], z0∈H).

It follows from (4.3) and (2.9) that D1εz0 →0 as ε→0, uniformly on [0, T]. Also, since t∈[0, T]7→T0tz0 is continuous becauseT0is a strongly continuous semigroup, we deduce that the set

T0tz0 ; t∈[0, T] is compact in H and thereforeDε3z0→0 asε→0 uniformly on [0, T].

For the term Dε2z0, we use the fact, proved in Section 3.4, Lemma 4.1 of [22], that Rελ(Tεt−T0t)R0λz0=−

Z t 0

Tεt−s(R0λ−Rελ)T0sz0ds (ε >0, t∈[0, T], z0∈H). (4.12) From (4.12), we have

Rελ(Tεt−T0t)R0λz0

X6

Z t 0

Tεt−s(Rλ0−Rελ)T0sz0

X ds 6

Z t 0

M eωT

(R0λ−Rελ)T0sz0

X ds (ε >0, t∈[0, T], z0∈H).

The integrand in the right hand side of the above estimate is bounded by an independent function of s and tends to 0 asε→0 by using (4.3). So, by Lebesgue’s convergence theorem, we have

Rελ(Tεt−T0t)R0λz0

X→0 asε→0 (t∈[0, T], z0∈H), and the limit is uniform in [0, T].

So we have thatDε2R0λz0→0 as ε→0 uniformly in [0, T] for everyz0∈H. Moreover, we know that every z0∈ D(A0) can be written asz0=R0λz1withz1∈H so we deduce thatDε2z0→0 asε→0 uniformly in [0, T] for every z0∈ D(A0). Then, we have actually proved that

(Tεt−T0t)Rλ0z0

X →0 asε→0 (t∈[0, T], z0∈ D(A0)), so we deduce that

(Tεt−T0t)z0

X→0 asε→0 t∈[0, T], z0∈ D(A20)

. (4.13)

Moreover the above convergence is uniform on [0, T]. By using that

Tεt−T0t

is uniformly bounded on [0, T] and since D(A20) is dense in H, see Theorem 1.2.7 of [22], we have that (4.13) holds for every z0 ∈H and uniformly fort∈[0, T], which concludes the proof.

4.2. Lax-Phillips semigroups

In this section we recall, following Staffans and Weiss [30], the concept of Lax-Phillips semigroups associated to a control system which is then used, combined with the Trotter-Kato theorem, to prove the convergence of the penalized Boussinesq control systems towards the usual one.

LetA:D(A)→ X be the generator of a strongly continuous semigroupF = (Ft)t>0. We denote byωF the growth bound of F and we fixω > ωF. LetB ∈ L(U,X) be a bounded control operator. We denote

Uω=L2ω([0,∞);U) := eωL2([0,∞);U), (4.14)

(13)

where (eωv)(t) =eωtv(t) andkeωvkL2ω =kvkL2. For later use, we also introduce the notation Wωm,2([0,∞);U) := eωWm,2([0,∞);U),

withkeωvkWm,2

ω =kvkWm,2. Let

V =D(A)× Uω.

It is known (see Grabowski and Callier [16] for the first derivation of this result and Staffans and Weiss [30] for a presentation of these results in a much more general context), that the operatorC:D(A)→ X × Uω defined by

D(C) =D(A)×Wω1,2((0,∞);U), (4.15)

C=

A Bδ0

0 d

, (4.16)

generates a strongly continuous semigroup ΞonX × Uω. More precisely, we have Ξt=

Ft Φt 0 St

(t>0), where (Φt)t>0 are the output maps defined by

Φtu= Z t

0

Ft−σBu(σ) dσ (t>0, u∈ Uω), andSt is the left shift by tonL2loc([0,∞);U),i.e.,

Stu(τ) =u(τ+t) (t, τ >0, u∈L2loc([0,∞);U)).

Remark 4.3. Note that, as mentioned in [30], the growth bound of Ξ equals toω.

Proposition 4.4. Let C be the operator defined in (4.15)and (4.16). Then its resolvent set ρ(C)contains the half plane

Cω={λ∈C | Reλ > ω}, and we have

(λI− C)−1 z

u

=

(λI− A)−1(BR

0 e−ληu(η) dη+z) ξ7→ −R

ξ eλ(ξ−η)u(η) dη

(λ∈Cω, z∈ X, u∈ Uω). (4.17) Proof. Given λ∈Cω, the equation

(λI− C) z0

u0

= z

u

, (4.18)

(14)

writes

(λI− A)z0+Bu0(0) =z, (4.19)

λu0−du0

dξ =u. (4.20)

Denotingu−ω= e−ωuand u0,−ω= e−ωu0, we note that u−ω∈L2([0,∞);U) and thatu0∈Wω1,2([0,∞);U) if and only ifu0,−ω∈W1,2([0,∞);U). Moreover, equation (4.20) can be rephrased in terms ofu−ω andu0,−ω as

du0,−ω

dξ (ξ) = (λ−ω)u0,−ω(ξ) +u−ω(ξ) (Re λ > ω, ξ∈(0,∞)).

In the following, we use some basic facts about Laplace transformation, see for instance Section 12.4 of [32]. By applying the Laplace transform to all the terms in the above equation we obtain

sbu0,−ω(s)−u0,−ω(0) = (λ−ω)ub0,−ω(s) +ub−ω(s) (Re s >0,Reλ > ω). (4.21) Choosing u0,−ω(0) =−ub−ω(λ−ω) we obtain

ub0,−ω(s) = bu−ω(s)−ub−ω(λ−ω)

s−(λ−ω) (Res >0,Reλ > ω). (4.22) Since, by the Paley-Wiener theorem, see Theorem 12.4.2 of [32], we have that ub−ω lies in the Hardy space H2(C0;U), the above formula implies that ub0,−ω ∈ H2(C0;U). Applying again the Paley-Wiener theorem it follows that ub0,−ω is the unique solution of (4.21) with u0,−ω∈L2([0,∞);U). Moreover, applying the inverse Laplace transform to (4.22), we obtain that

u0,−ω(ξ) =−e(λ−ω)ξ Z

ξ

e(ω−λ)ηu(η) dη˜ (Reλ > ω, ξ∈(0,∞)),

so that

u0(ξ) =− Z

ξ

eλ(ξ−η)u(η) dη (Reλ > ω, ξ∈(0,∞)). (4.23) Going back to (4.19), it follows that

z0= (λI− A)−1(−u0(0) +z) (Reλ > ω). (4.24) Finally, putting together (4.23) and (4.24) we obtain the conclusion (4.17).

Let us now retrieve our (penalized) Boussinesq context, with the notation for the operatorsAεand semigroups Tεrecalled at the beginning of this section. We know from Proposition2.1and Proposition2.2that there exist M >1 and ˜ω∈Rsuch that the corresponding semigroups satisfy

T0t

L(X;X)6M eωt˜ and kTεtkL(X;X)6M eωt˜ , (ε >0, t>0). (4.25) Forω >ω, we consider the space˜ Uω introduced in (4.14) takingU =U.

(15)

We can introduce the Lax-Phillips semigroup Ξ0= (Ξ0t)t>0onH× Uω, associated to the pair (A0, B0=PHB) and, for everyε >0, the Lax-Phillips semigroup Ξε= (Ξεt)t>0onX× Uω, associated to the pair (Aε, B).

The main result in this section is given just below and it will be used several times in the remaining part of this work.

Proposition 4.5. With the above notation, we have

ε→0+lim

Ξεt z

u

−Ξ0t z

u

X×U

= 0

t>0,

z u

∈H× Uω

, (4.26)

ε→0+lim

εt) z

u

−(Ξ0t) z

u

X×U

= 0

t>0,

z u

∈H× Uω

, (4.27)

Moreover, the convergence above is uniform for t in compact intervals.

Proof. We only prove (4.26).

The generator of Ξ0is the operatorC0:D(C0)→H× Uω defined by

D(C0) =D(A0)×Wω1,2((0,∞);U), (4.28)

C0=

A0 B0δ0

0 d

. (4.29)

Similarly, for everyε >0, the generator of Ξεis the operatorCε:D(Cε)→X× U defined by

D(Cε) =D(Aε)×Wω1,2((0,∞);U), (4.30)

Cε=

Aε0

0 d

. (4.31)

By Proposition4.4, we have for everyλ∈Cω, (λI− C0)−1

z u

=

(λI−A0)−1(B0

R

0 e−ληu(η) dη+z) ξ7→ −R

ξ eλ(ξ−η)u(η) dη

z u

∈H× Uω

, (4.32)

(λI− Cε)−1 z

u

=

(λI−Aε)−1(BR

0 e−ληu(η) dη+z) ξ7→ −R

ξ eλ(ξ−η)u(η) dη

z u

∈X× Uω

. (4.33)

By combining Proposition4.1, recalling that B0=PHB and takingz∈H, it follows that

ε→0+lim

(λI− Cε)−1 z

u

−(λI− C0)−1 z

u

X×U

= 0

z u

∈H× Uω

. (4.34)

Then the conclusion (4.26) follows from (4.34) and (4.25) by a slight variation of the Trotter-Kato theorem, in the spirit of the proof of Proposition 4.2.

(16)

Remark 4.6. Note that for every ω > 0, L2([0,∞);U) is contained in Uω so one can apply in particular Proposition 4.5for controlsu∈L2([0,∞);U).

4.3. Uniform constants in Datko’s theorem

The goal of this subsection is to provide a version of a well-known theorem of Datko, in which we make explicit the fact that the exponential decay rate of the semigroup depends only on the constants involved in assumptions (4.35) and (4.36) below.

Proposition 4.7. Let A be a generator of a strongly continuous semigroup (Ft)t>0 on a Hilbert space X. Assume that there exists a positive constantC >0 such that

kFtkL(X;X)6CeCt (t>0), (4.35) and

Z 0

kFthk2Xdt6Ckhk2X (h∈ X). (4.36)

Then there existM >0 andω >0, depending only onC, such that

kFtkL(X;X)6M e−ωt (t>0). (4.37) Proof. We closely follow the methodology used, for instance, in Tucsnak and Weiss ([32], Sect. 6.1), in proving Datko’s theorem.

First, from (4.35), we have that for every τ >0 there exists mτ >1 (depending also on C) such that kFtkL(X;X)6mτ for allt∈[0, τ]. Then

kFτhk2= 1 τ

Z τ 0

kFτ−tFthk2dt6m2τ τ

Z τ 0

kFthk2dt (τ >0, h∈ X). (4.38)

By combining (4.38) and (4.36) it follows that

c20khk2>X

k∈N

Z (k−1)τ

kFthk2Xdt>κ2τX

k∈N

kFhk2 (τ >0, h∈ X), (4.39)

where kτ =

τ

mτ. In particular, we see from the above thatkFkL(X;X)6 κc0τ for everyτ >0 and everyk∈N (and this holds also fork= 0). Hence, for everyτ >0, everyn∈Nand everyh∈ X,

kFhk2= 1 n

n

X

k=1

kF(n−k)τFhk26 c202τ

n

X

k=1

kFhk2.

By (4.39) we get that

kFhk26 c202τ · c20

κ2τkhk2 (τ >0, n∈N, h∈ X),

(17)

so that

kFkL(X;X)6 c20

√nκ2τ (τ >0, n∈N).

Denotingnτ =h4c4 0

κ4τ

i

+ 1 it follows that

n∈infN

1

nτlogkFkL(X;X)6 1

nττlogkFnττkL(X;X)6 −log 2

nττ (τ >0).

Finally, combining the above estimate and a classical result (see, for instance, Tucsnak and Weiss [32], formula (2.1.3)), it follows that

ω0(F) = inf

t∈(0,∞)

1

t logkFtkL(X;X)6 inf

n∈N

1

nτ logkFkL(X;X)6 −log 2

nττ (τ >0), which ends the proof.

5. Uniform stabilizability and null-controllability results

In the first part of this section we show that, assuming a smallness assumption on the stationary state vs

ys

(see Prop. 5.1), the semigroupsTε generated by the operatorsAεdefined in (2.7) are uniformly (with respect toε) exponentially stable. This will be a basic ingredient of the proof of Theorem3.1. In the second part of this section we derive the corresponding ingredient of the proof of Theorem 3.2, which is is a uniform (again with respect to ε) null-controllability result for the pair (Aε, B), whereB is the operator defined in (3.9).

5.1. Small stationary states and arbitrary control operator

In this subsection we show that if the stationary state around which we linearized the problem is small (in an appropriate sense) then the semigroups (Tε)ε>0are uniformly exponentially stable. More precisely, we have:

Proposition 5.1. There exists δ >0 such that if (3.3) holds then Aε generates a uniformly stable semigroup on X, i.e. there existM >1 andω >0 independent ofε such that

kTεtkL(X;X)6M e−ωt (ε >0, t>0). (5.1) Proof. We split the proof into three steps.

Step 1.For everyε >0, consider the operatorAeε:D(Aeε) =D(Aε)→X defined by A˜ε

ϕ ζ

=

ν∆ϕ+1ε∇(divϕ) +ζed

α∆ζ

ϕ ζ

∈ D(Aeε)

.

In other words, looking to (2.7), we have A˜ε

ϕ ζ

=Aε

ϕ ζ

+

(vs· ∇)ϕ+ (ϕ· ∇)vs

(vs· ∇ζ) + (ϕ· ∇ys)

ϕ ζ

∈ D(Aeε)

. (5.2)

(18)

To show that the resolvent set of Aeεcontains the half planeC−β :={z∈C; Rez >−β} for someβ >0, we note that the resolvent equation forAeε, that is

(sI−Aeε) ϕ

ζ

= f

g

∈X, (5.3)

can, according to (2.8) and (5.2), be equivalently written as

s Z

ϕ·ψdx+ν Z

∇ϕ:∇ψdx+1 ε

Z

divϕ·divψ dx= Z

ζ·ψddx+ Z

f·ψdx (5.4) s

Z

ζηdx+α Z

∇ζ· ∇ηdx= Z

gηdx, (5.5)

for every ψ

η

∈ D(Aε). Note that the second equation is decoupled from the first. By Lax-Milgram’s lemma and Poincar´e’s inequality, there existsβ1>0 such that for everys∈C−β1, (5.5) admits a unique solutionζ= (sI− α∆)−1g. Inserting this information in (5.4), we see that by Lax-Milgram’s lemma and Poincar´e’s inequality, there existsβ2>0 such that for everys∈C−β2, (5.4) admits a unique solutionϕ= sI−ν∆−1ε∇(div )−1

d+f).

This proves that indeed the resolvent set ofAeεcontainsC−β forβ= min(β1, β2)>0.

Step 2. We remark that there exists σ∈(0, β) such that σI+Aeε generates an exponentially stable analytic semigroup. Indeed, this is a direct application of Theorem 4.3 page 118 in [22], using Step 1 and Proposition2.2.

Step 3.We show that the operatorE, with domainD(E) =H01(Ω)d+1, defined by

E ϕ

ζ

, ψ

η

X

=− Z

[(vs· ∇)ϕ+ (ϕ· ∇)vs]·ψdx

− Z

[vs· ∇ζ+ϕ· ∇ys]·ηdx

ϕ ζ

, ψ

η

∈ D(E)

, (5.6)

is (σI+Aeε)-bounded in the usual sense, recalled below. To this aim, let ϕ, ψ

∈ D(E). We know from Step 1, that

ϕ ψ

= (−σI−Aeε)−1 f

g

for some f

g

∈X and there exists a constantC >0 such that

ϕ ψ

H1

0(Ω)d+1

6C

f g

X

. (5.7)

Assuming that (3.3) holds, we have by using (5.7) that

E

ϕ ψ

X

ϕ ψ

H1

0(Ω)d+1

6Cδ

f g

X6Cδ

(−σI−A˜ε) ϕ

ψ

X

.

Takingδ >0 sufficiently small, and using Corollary 2.3, page 81 of [22] we obtain that ˜Aε+σI+E=Aε+σI generates a bounded analytic semigroup onX, soAεgenerates a uniform stable semigroup on X (note thatσ is independent ofε >0). This concludes the proof.

Références

Documents relatifs

Receding horizon control, model predictive control, value function, optimality systems, Riccati equation, turnpike property.. AMS

In a nutshell, receding horizon control (also called model predictive control) consists in replacing the infinite- horizon control problem (1) by a sequence of finite-horizon

In this note, we propose a new method to approximate the minimal time crisis problem using an auxiliary control and a penalty function, and show its convergence to a solution to

The simulations results are presented in the form of surface thermographic images, to show the contrast in temperature due to the presence of the weld zone.To better quantify

The penalty method is a classical and widespread method for the numerical treatment of con- strained problems, in particular the unilateral contact problems arising in mechanics of

There is a huge literature dealing with the Monge-Kantorovich optimal transport problem; see e.g. In this note we concentrate on the simplest possible setting, namely a finite number

Using uniform global Carleman estimates for semi-discrete elliptic and hyperbolic equations, we study Lipschitz and logarithmic stability for the inverse problem of recovering

We establish some optimal a priori error estimate on some variants of the eXtended Finite Element Method (Xfem), namely the Xfem with a cut-off function and the stan- dard Xfem with