• Aucun résultat trouvé

Feedback stabilization of the 2-D and 3-D Navier-Stokes equations based on an extended system

N/A
N/A
Protected

Academic year: 2022

Partager "Feedback stabilization of the 2-D and 3-D Navier-Stokes equations based on an extended system"

Copied!
35
0
0

Texte intégral

(1)

DOI:10.1051/cocv:2008059 www.esaim-cocv.org

FEEDBACK STABILIZATION OF THE 2-D AND 3-D NAVIER-STOKES EQUATIONS BASED ON AN EXTENDED SYSTEM

Mehdi Badra

1

Abstract. We study the local exponential stabilization of the 2D and 3D Navier-Stokes equations in a bounded domain, around a given steady-state flow, by means of a boundary control. We look for a control so that the solution to the Navier-Stokes equations be a strong solution. In the 3D case, such solutions may exist if the Dirichlet control satisfies a compatibility condition with the initial condition.

In order to determine a feedback law satisfying such a compatibility condition, we consider an extended system coupling the Navier-Stokes equations with an equation satisfied by the control on the boundary of the domain. We determine a linear feedback law by solving a linear quadratic control problem for the linearized extended system. We show that this feedback law also stabilizes the nonlinear extended system.

Mathematics Subject Classification.35Q30, 76D05, 76D07, 76D55, 93B52, 93C20, 93D15.

Received March 23, 2006. Revised July 11, 2007 and April 8, 2008.

Published online October 21, 2008.

1. Introduction

Let Ω be a bounded and connected domain inRdford= 2 ord= 3, with a boundary Γ =∂Ω of classC4, and composed ofN connected components Γ(1), . . . ,Γ(N). Let us consider a stationary motion of an incompressible fluid in Ω which is described by the pair (zs, ps), the velocity and the pressure, solution to the stationary Navier-Stokes equations:

−νΔzs+ (zs· ∇)zs+∇ps=f, ∇ ·zs= 0 in Ω and zs=vb on Γ. (1.1) In the above setting,ν >0 is the viscosity,f is a function inL2(Ω),vbbelongs toH3/2(Γ) and obeys

Γ(j)vb·n= 0, for allj = 1. . . N, wherendenotes the unit normal vector to Γ, exterior to Ω. Notice that here and in the following, we write in bold the spaces of vector fields: L2(Ω) = (L2(Ω))d,H3/2(Γ) = (H3/2(Γ))d, etc. We recall that a solution to (1.1) is known to exist inH2(Ω)×H1(Ω)/R[14], Chapter VIII, Theorems 4.1 and 5.2.

Ifzsis an unstable equilibrium state, and if we assume that at time t= 0 the velocity is equal to z0=zs, then even ifz0 is close tozs, the resulting unsteady velocity ¯z(t) whent >0 will not necessary stay close to zs. Hence, the question we address here is: how to obtain a controller localized on the boundary Γ, which makes

¯

z(t) go back tozsas t→ ∞?

Keywords and phrases. Navier-Stokes equation, feedback stabilization, Dirichlet control, Riccati equation.

1 Laboratoire LMA, UMR CNRS 5142, Universit´e de Pau et des Pays de l’Adour, 64013 Pau Cedex, France.

mehdi.badra@univ-pau.fr

Article published by EDP Sciences c EDP Sciences, SMAI 2008

(2)

More precisely, we consider a pair (¯z,p) solution to the instationary Navier-Stokes equations:¯

tz¯−νΔ¯z+ (¯z· ∇)¯z+∇p¯=f and ∇ ·¯z= 0 in Q, (1.2)

¯

z=vb+u on Σ, z(0) =¯ zs+z0, (1.3)

and we assume that zs is an unstable solution of (1.2)-(1.3) corresponding toz0= 0 and u= 0. In the above setting Q= Ω×(0,∞) and Σ = Γ×(0,∞). Thus, if we make the change of variable (¯z,p) = (z¯ s+z, ps+p), we have:

tz−νΔz+ (z· ∇)zs+ (zs· ∇)z+ (z· ∇)z+∇p= 0 in Q, (1.4)

∇ ·z= 0 in Q, z=u on Σ, z(0) =z0, (1.5)

and the question of making ¯z(t) go back tozsast→ ∞is equivalent to the one of makingz(t) go back to 0 as t→ ∞. The following questions may be addressed:

Can we find a set of initial conditionsWδ ={y∈X(Ω)| yX(Ω)< δ}, where δ >0 andX(Ω)⊂ {y∈ L2(Ω) | ∇ ·y = 0}, and can we find a space of controlsU(Γ) such that, for z0 ∈ Wδ, there exists a boundary controlu∈L2(0,∞;U(Γ)) for which the solution to (1.4)-(1.5) satisfies the exponential decay stated below?

z(t)X(Ω)≤Ce−ηtz0X(Ω) η >0. (1.6)

Can we express uin a feedback formulation? More precisely, we are interested in the existence of an operatorF∈ L(X(Ω), U(Γ)), independent of the time variablet≥0, and such that

u(t) =F z(t), t≥0. (1.7)

Is there a practical way to computeF?

First, let us mention some results partially answering those questions. In the two and three dimensional case, the existence of a pair (z, u), which satisfies (1.4)-(1.5) and (1.6), is stated in [11,12] withX(Ω) ={y∈H1(Ω)|

∇ ·y= 0, y|Γ0 = 0, tΓy·n= 0}andU(Γ) ={y∈H3/2(Γ)|y|Γ0 = 0,

Γy·n= 0}, Γ = ΓΓ0and ΓΓ0=∅.

The key idea in [11,12] relies in an adequate extension operator which maps an initial condition, defined in Ω, to an extended and stable initial condition, defined in an open setGwhich contains Ω. By this way, the author obtains an operatorF0∈ L(X(Ω), L2(0,∞;U(Γ))) such thatu=F0z0, but he does not obtain a control in the pointwise (in time) formulation (1.7).

In the two dimensional case, the existence of a pair (z, u) satisfying (1.4)-(1.5) and (1.6)-(1.7), and such that uis localized in a part of Γ and has a non vanishing normal component, is proved in [21]. In this paper, X(Ω) = {y H1/2−(Ω) | ∇ ·y = 0, y·n = 0 on Γ} and U(Γ) = {my | y L2(Γ),

Γmy·n = 0}, where ]0,1/4[ and m C2(Γ) is an adequate localization function. The feedback controller is determined by an algebraic Riccati equation which is obtained by solving an optimal control problem. The key point of this approach relies in a reformulation of system (1.4)-(1.5), which only involvesP z, where P is the orthogonal Leray projection operator (see Sect. 2.1). We point out the fact that, since the three dimensional case is more demanding in terms of velocity regularity, and in particular we will see that it requires the compatibility condition u(0) = z0|Γ, it cannot be treated in the same way. In [22] the author overcomes this difficulty, by introducing a time dependent feedback law.

The three dimensional case is treated in [4]. The existence of a pair (z, u), which satisfies (1.4)-(1.5) and (1.6)-(1.7), is stated. In this paper, X(Ω) ={y H1/2+(Ω) | ∇ ·y = 0, y·n = 0 on Γ} and U(Γ) = {y L2(Γ)|y·n= 0 on Γ}. The authors follow the ideas developed in [3,5], where the case of pointwise feedback stabilization of the 3D Navier-Stokes equations, by means of a distributed control, is investigated. However, the boundary feedback law which is proposed in [4] cannot be numerically calculated. This difficulty is closely linked to the high degree of regularity for the velocity which is necessary to obtain the exponential decrease of the solution of the Navier-Stokes system in the three dimensional case. To obtain the required smoothness degree for

(3)

the state, the authors solve an optimal control problem involving the velocity normL2(0,∞;H3/2+(Ω)) in the cost functional, and it does not allow to define a feedback law from a well posed Riccati equation. Indeed, the Riccati equation is only defined inD(A2R), whereARis the infinitesimal generator of the associated closed-loop system, which itself depends on the unknownR of the Riccati equation.

In fact, when d= 3, obtaining a well-posed closed-loop system (1.4)-(1.5)-(1.7) is not an easy task. Let us give some explanations about the difficulties linked to the three dimensional analysis, which requires a particular compatibility condition between the state and the control. After we compare the maximal order derivatives in time and in space appearing in (1.4), we introduce the function space

Hα,β(Q) =L2(0,∞;Hα(Ω))∩Hβ(0,∞;L2(Ω)), α≥0, β0,

and we postulate that a strong solution to (1.4)-(1.5) should be searched in H1+s,1/2+s/2(Q) for s≥0. This framework is used in [21] to define solutions to the two dimensional closed-loop Navier-Stokes system and in [23]

to obtain optimal regularity results for the Oseen system with a nonhomogeneous boundary condition. Hence, z obeys:

tz∈H−1/2+s/2(0,∞;L2(Ω)) and −νΔz+ (z· ∇)zs+ (zs· ∇)z∈L2(0,∞;Hs−1(Ω)). (1.8) In order to get rid of the pressure term in (1.4) and to simplify our analysis, it is more convenient to evaluate the state equation (1.4) in a space of distributions which is orthogonal to the space of gradient pressures. By projecting equation (1.4) ontoV0s−1(Ω), the dual space ofV01−s(Ω) ={y∈H1−s0 (Ω)| ∇ ·y= 0, y·n= 0 on Γ}

(see Sect. 2.1), we obtain the following necessary condition from (1.8):

−P(z· ∇)z=P(∂tz−νΔz+ (z· ∇)zs+ (zs· ∇)z)∈Ys(Q), whereP is the Leray orthogonal projection operator (see Sect.2.1), and where

Ys(Q) =H−1/2+s/2(0,∞;L2(Ω)) +L2(0,∞;V0s−1(Ω)). (1.9) Thus, by remarking that the free divergence condition∇ ·z= 0 yields (z· ∇)z=∇ ·(z⊗z), we deduce that we shall look for a velocityz solution to (1.4)-(1.5)-(1.7) which obeys:

z∈H1+s,1/2+s/2(Q) and P∇ ·(z⊗z)∈Ys(Q). (1.10) A brief check of the regularity of∇ ·(z⊗z) which can be obtained fromz∈H1+s,1/2+s/2(Q), shows that when d= 3, the valuesshould be chosen greater than 1/2. Indeed, from the continuous embeddingH1+s,1/2+s/2(Q) H1/4(0,∞;Hs+1/2(Ω)) and with

uv|(u, v)∈H1/4(0,∞;Hs+1/2(Ω))×H1/4(0,∞;Hs+1/2(Ω))

⊂L2(0,∞;H2s−1/2(Ω)),

it yields z⊗z ∈L2(0,∞;H2s−1/2(Ω)). Then we obtain ∇ ·(z⊗z)∈ L2(0,∞;H2s−3/2(Ω)), and for s≥ 1/2 the second statement in (1.10) follows from L2(0,∞;H2s−3/2(Ω)) →L2(0,∞;H−1+s(Ω)). As a consequence, when d = 3 the nonlinearity of the Navier-Stokes system imposes to define a solution z which belongs to H1+s,1/2+s/2(Q) for s 1/2. Hence, the trace theorem yields z|Σ H1/2+s,1/4+s/2(Σ), and the feedback controller has to obey:

z|Σ=F z∈H1/4+s/2(0,∞;L2(Γ)) s≥1/2. (1.11) Since 1/4 +s/2≥1/2, some kind of continuity is required for the control. In the particular case wheres >1/2, the spaceH1/4+s/2(0,∞;L2(Γ)) is a subspace ofC([0,∞[ ;L2(Γ)) (the space of time continuous functions with value inL2(Γ)) and we deduce from (1.11) that the velocity z must satisfy the initial compatibility condition

(4)

z0|Γ=F z0. This explains why the feedback law which is given in [21] cannot be used in the three dimensional case, and why the author overcomes this difficulty in [22] by introducing a feedback law which is time dependent in an initial transitory time interval. Notice that spaces of initial conditions, for which a stabilization result can be obtain with the Riccati approach, are precisely given in [2].

In fact, finding a feedback controller independent of the time variable and which satisfy F z0 = z0|Γ for a sufficiently large class of initial conditionsz0 is not obvious. That is the reason why in the present paper, we propose to search another type of pointwise (in time) feedback law. We searchuas a solution to the following evolution system:

tu−Δbu−σ n=K(z, u), u(0) =z0|Γ, (1.12) where the feedback controller K now acts on the pair (z, u). Here Δb is the vector-valued Laplace Beltrami operator (see Sect.5). Formulation (1.12) involves the time derivative ofu, so we can fix the initial condition:

the initial boundary value u(0) now fits the initial trace z0|Γ. We underline that we had a large degree of freedom in the choice of the boundary system. We have chosen (1.12) for its simplicity and for numerical computational conveniences in view of future implementations (umust be numerically calculated). But one may imagine another boundary system which would have a physical interpretation and which could be concretely constructed. The present paper is a complete and detailed version of [1]. Its main objectives are:

(i) to show the existence of a pair (z, u) satisfying (1.4)-(1.5)-(1.6) and (1.12) in the two and three dimen- sional case;

(ii) to find an operatorK which is provided by a well-posed Riccati equation;

(iii) to find a way to obtain a controlulocalized on an arbitrary small part of Γ.

The Navier-Stokes equations coupled with a boundary system. We define the space of initial conditions:

X(Ω) =Vs(Ω) :=

y∈Hs(Ω)| ∇ ·y= 0,

Γ

y·n= 0

, s∈]1/2,1], (1.13) and we assume that z0 ∈X(Ω). In order to impose the compatibility conditionu(0) = z0|Γ and to obtain a sufficient time regularity level forz, we choose to searchuas the first component of (u, σ) which satisfies:

tu−Δbu−σ n=K(z, u) in Σ, u(0) =z0|Γ and

Γ

u(t)·n= 0. (1.14) Recall that Δb is the vector-valued Laplace Beltrami operator andσ∈L2(0, T) plays the role of the Lagrange multiplier associated with the constraint

Γu·n= 0. The feedback lawKis a linear operator, independent of t≥0, and it couples (1.14) with (1.4)-(1.5). The state (z, u) now satisfies the following coupled system:

tz−νΔz+ (z· ∇)zs+ (zs· ∇)z+ (z· ∇)z+∇p= 0, ∇ ·z= 0 in Q, (1.15)

tu−Δbu−σ n=K(z, u), z=u on Σ,

Γ

u(t)·n= 0, (1.16)

z(0) =z0∈Vs(Ω), u(0) =z0|Γ. (1.17)

We are going to show that we can choose K so that (1.15)-(1.16)-(1.17) is well defined when z0 is small in Vs(Ω), and so thatzobeys (1.6). The operatorKcan be considered as a pointwise feedback controller which is acting on (z, u) solution to the extended system (1.15)-(1.16)-(1.17). That is the reason why we shall say that our approach is a compromise between the formulation of a control in the form (1.7), and the treatment of the 3 dimensional case which requires a high regularity level for the control.

Calculation of the feedback controller K. In a first step, we shall simplify our problem and consider the question of stabilizing the linear system obtained from (1.15)-(1.16)-(1.17) by linearizing this one around (0,0).

In other words, we want to find a controlg∈L2(0,∞;L2(Γ)), which can be expressed in a feedback form, and

(5)

such that the solution (z, u) to the following linear system is stable:

tz−νΔz+ (z· ∇)zs+ (zs· ∇)z+∇p= 0, ∇ ·z= 0 in Q, z(0) =z0, (1.18)

tu−Δbu−σ n=g, z|Σ=u on Σ,

Γ

u(t)·n= 0, u(0) =u0. (1.19) In the above setting, (z0, u0) is an arbitrary initial pair satisfyingz0·n=u0·non Γ. The question of constructing a linear feedback controller stabilizing a linear dynamical system can be answered with the so-called “Riccati approach”. It consists in solving an auxiliary optimal control problem, defined over an infinite time horizon, and which involves a linear quadratic cost functional. It provides an optimal control in a feedback form, with a feedback law depending on the solution to an algebraic Riccati equation. Such optimal control theory is developed in [18], Chapter 2. We shall underline the fact that to use the Riccati theory, we shall work with an abstract dynamical model representing equations (1.18)-(1.19) [18], Chapter 2, Section 2.1. Hence, we need to rewrite the system (1.18)-(1.19) as an evolution equation of the type:

Y=AY + ΛG on D(A), Y(0) =Y0, (1.20) where Ais a linear free dynamic operator, Λ is a linear control operator, andY and Gare the new state and control variables. This can be achieve with the new variableY = (y, u)T = (P z, u)T, whereP is the orthogonal Leray projection operator (see Sect. 2.1), with Λ as the canonical projection (y, u)T −→ (0, u)T and with an operatorAdefined from the free dynamic operators related to each equations (1.18) and (1.19) separately. More precisely, the dynamical system (1.20) can be obtain from the dynamical system related toy=P zandu:

y = Ay+Bu on D(A), y(0) =P z0, u = Abu+g on D(Ab), u(0) =u0,

where A is the free dynamic Oseen operator (see Def. 6.1), Ab is the free dynamic operator defined from Δb (see Sect.5), andBis an input operator which allows to represent the trace condition linkingyandu. We shall insist on the fact that the main difficulty relies in the definition of the operatorB, which can be done by using the theory developed in [23]. Hence, we look for the control Gand the associated state Y solution to (1.20) which minimize the cost functional:

J(Y, G) =

0

CY2Z+

0

G2L2(Ω)×L2(Γ),

whereC is an observation operator with value in the spaceZ. We define the optimal control problem:

inf

J(Y, G),(Y, G) satisfies (1.20)

(1.21) and the resolution of (1.21) provides a feedback control G= ΛΠY where Π is the solution of an algebraic Riccati equation which can be formally written as follows:

ΠA+AΠΠΛΠ +CC= 0. (1.22)

The precise definition of such a Riccati equation is given in Theorem7.3. Thus, we apply this feedback control to the nonlinear system, and it yields the following expression ofK in (1.15)-(1.16)-(1.17):

K(z, u) =−Π2P z−Π3u where Π =

Π1 Π2 Π2 Π3

is the solution to (1.22).

(6)

Notice that (1.22) gives a practical way to calculate the operatorK. Finally, we show thatK stabilizes (1.15)- (1.16)-(1.17) in a neighborhood of the origin: there is δ >0 such that, if z0Vs(Ω) < δ, then there exists a unique pair (z, u) satisfying (1.15)-(1.16)-(1.17), which obeys (1.6). We recall thatVs(Ω) is defined in (1.13).

Localization of the control in a part of the boundary. To treat the case of a boundary control which is localized in an open part of Γ, we replace the boundary condition z|Σ =u byz|Σ =m(u−σm(u)n), where σm(u) = (

Γm)−1

Γmu·n andm∈C2(Γ) is an adequate cut-off function with values in [0,1]. By this way, the action ofuis localized on Γm = Supp(m). Thus, we define the corresponding operator (Am,D(Am)) and Cm, and the resolution of

inf

0

CmY2Ξ+

0

G2L2(Ω)×L2(Γ)| Y =AmY + ΛG, Y(0) =Y0

, (1.23)

provides an operator Πm satisfying

ΠmAm+AmΠmΠmΛΠm+CmCm= 0, Πm=

Π1,m Π2,m Π2,m Π3,m

.

Hence, we obtain a local stabilization result with the feedback control K(z, u) =−Π2,mP z−Π3,mu. Such a treatment of localized control only adds technical difficulties in the statement of the finite cost condition which guarantees the well-posedness of (1.23). For readability convenience, the main parts of this paper deals with the non localized case, and we postpone the treatment of a localized control to Sections9 and10.

The paper is organized as follows. In Section2 we recall some background material needed throughout the paper and we define spaces of initial conditions. Section3is dedicated to the statement of the local stabilization result. We write the Oseen system in the form of an evolution equation in Section4, and we write the differential boundary system in the form of an evolution equation in Section 5. Next, we define the operator A and we study the linear system (1.20) in Section6. Section7is dedicated to the study of the optimal control problem (1.21) which provides a feedback controllerK. In Section8, we apply this feedback law to the nonlinear system and we give a proof of the local stabilization result. Finally, we deal with the localization of the control on a part of the boundary in Section 9, and we postpone in a appendix the proof of a finite cost condition ensuring the well-posedness of (1.21) and (1.23).

2. Functional framework 2.1. Notations

LetX and Y be two Hilbert spaces. If Ais a closed linear mapping inX, we denote its domain byD(A).

We denote by L(X, Y) the space of all bounded operators from X to Y, and we use the shorter expression L(X) = L(X, X). For 0 < T ≤ ∞, the space L2(0, T;X) is the usual vector-valued Lebesgue space and Hα(0, T;X) for α 0 is the usual vector-valued Sobolev space. If C0(]0, T[ ;X) is the space of infinitely differentiable and compactly supported functions oft∈]0, T[ with values inX, we denote byH0α(0, T;X) the closure ofC0(]0, T[ ;X) inHα(0, T;X), and byH−α(0, T;X) the dual space ofH0α(0, T;X), whereXdenotes the dual space ofX. We also define:

W(0, T;X, Y) =

y∈L2(0, T;X) | dy

dt ∈L2(0, T;Y)

.

It is well known that ifX is continuously and densely embedded inY, then the spaceW(0, T;X, Y) is continu- ously embedded inC([0, T]; [X, Y]1/2) ifT <∞or inCb([0,[ ; [X, Y]1/2) ifT =, the space of bounded and time continuous functions with values in the interpolation space [X, Y]1/2[19].

Next, let us recall that Ω is a bounded and connected domain inRd, ford= 2 or d= 3, with a boundary Γ =∂Ω of class C4, and composed ofN connected components Γ(1), . . . ,Γ(N). We will use the usual function

(7)

spacesL2(Ω), Hs(Ω),H0s(Ω) andH−s(Ω) = (H0s(Ω)), and we write in bold the spaces of vector fieldsL2(Ω) = (L2(Ω))d, Hs(Ω) = (Hs(Ω))d, Hs0(Ω) = (H0s(Ω))d and H−s(Ω) = (H−s(Ω))d. The norms are denoted by · Z(Ω), where the subscriptZ(Ω) refers to the space which is considered, and we denote the scalar product in L2(Ω) by (·|·). We denote by Δ the vector-valued Laplace operator, with domainD(Δ) =H2(Ω)∩H10(Ω), which is known to be selfadjoint. For alls∈[0,2], its fractional power (−Δ)s/2is well defined and obeysD((−Δ)s/2) = [H2(Ω)H10(Ω),L2(Ω)]1−s/2, where [·,·] denotes the complex interpolation method. LetL2−1/2(Ω) be the space of functions y L2(Ω) such that

Ωρ(x)−1|y|2 < +, where ρ(x) is the distance from x to Γ. From [15], Theorem 8.1, it can be deduced thatD((Δ)s/2) =Hs(Ω) if 0≤s <1/2, thatD((Δ)1/2) ={y∈H1/2(Ω) | y∈L2−1/2(Ω)} and thatD((−Δ)s/2) ={y∈Hs(Ω) | y|Γ= 0}if 1/2< s≤2.

Next, ify∈L2(Ω) is such that∇·y∈L2(Ω), we can define its normal tracey·ninH−1/2(Γ) [13], Section III.3, and we introduce the spaces of free divergence functions:

Vs(Ω) =

y∈Hs(Ω)| ∇ ·y= 0 in Ω,

Γ

y·n= 0

, s∈[0,2], Vns(Ω) =

y∈Hs(Ω)| ∇ ·y= 0 in Ω, y·n= 0 on Γ

, s∈[0,2].

Moreover, we denote by P the so-called Leray projector which is the orthogonal projector from L2(Ω) onto Vn0(Ω) [13], Chapter III, Theorem 1.1, we define the self-adjoint operator A0 = PΔ with domain D(A0) = H2(Ω)H10(Ω)∩Vn0(Ω), and fors∈[0,2] we introduce the spaceV0s(Ω) =D((−A0)s/2). According to [10], we have

V0s(Ω) =D((−Δ)s/2)∩Vn0(Ω) =

H2(Ω)H10(Ω),L2(Ω) 1−s/2∩Vn0(Ω) for alls∈[0,2], which yields the following equalities:

V0s(Ω) = Vns(Ω), s∈[0,1/2[, V01/2(Ω) =

y∈Vn1/2(Ω) | y∈L2−1/2(Ω)

, V0s(Ω) =

y∈Vns(Ω) | y= 0 on Γ

, s∈]1/2,2].

Notice that the subscript 0 in V0s(Ω) only means that one may have a vanishing Dirichlet boundary condi- tion. The above characterization can also be obtain from the equality D((−A0)s/2) = [H2(Ω)H10(Ω) Vn0(Ω), Vn0(Ω)]1−s/2 for s [0,2] (which can be obtain from [6], Chap. 1, Cor. 6.1), by remarking that K0=A−10 PΔ defines a bounded projector from L2(Ω) intoVn0(Ω) which allows to invoke [26], Section 1.17.1, Theorem 1. Finally, for s >2 we define V0s(Ω) =V02(Ω)Hs(Ω) and fors <0 we defineV0s(Ω) = (V0−s(Ω)), the dual space of V0−s(Ω) with respect to the pivot space Vn0(Ω). It is equipped with the duality pairing

·|·V−s

0 (Ω),V0s(Ω). We also recall that P can be extended to a bounded linear operator from H−1(Ω) onto V0−1(Ω) by

P y:w∈V01(Ω)−→

yw

H−1(Ω),H10(Ω), [4], Appendix A.

Next, we define the spaces of pressures with free mean L20(Ω) =

p∈L2(Ω)|

Ω

p= 0

and Hs(Ω) =Hs(Ω)∩L20(Ω), s≥0, and we recall that the followingHelmholtz decomposition holds:

L2(Ω) =Vn0(Ω)⊕ ∇H1(Ω).

(8)

Next, we define the following trace spaces with a free mean normal component Vs(Γ) =

y∈Hs(Γ)|

Γ

y·n= 0

, s∈[0,3], V−s(Γ) =

y∈H−s(Γ)| y|nH−s(Γ),Hs(Γ)= 0

, s∈[0,3].

We make the identification (Vs(Γ)) =H−s(Γ)/(Vs(Γ)), where (Vs(Γ)) denotes the dual space ofVs(Γ) with respect to the pivot space V0(Γ) and (Vs(Γ)) = {y H−s(Γ) | y|wH−s(Γ),Hs(Γ) = 0, ∀w Vs(Γ)}. We verify that (Vs(Γ))=Rn, and that:

V−s(Γ) =H−s(Γ)/Rn= (Vs(Γ)).

Moreover, we introduce the orthogonal projectorPb fromL2(Γ) ontoV0(Γ), which is explicitly given by Pbv=v− 1

|Γ| Γv·n

n. (2.1)

Next, we recall that the normal trace operator γn ∈ L(L2(Γ)) is defined by γn(u) = (u·n)n∈L2(Γ), and we extend its definition toV−s(Γ) fors >0 with the formula

γn(u) =u|γn(·)V−s(Γ),Vs(Γ) for allu∈V−s(Γ).

The boundary normal derivative on Γ of a vector fieldv∈H2(Ω) is defined bynv= (∇v)n.

Finally, we shall underline that we will also need the spaces

H0=Vn0(Ω)×V−1/2(Γ) and H0=Vn0(Ω)×V1/2(Γ),

and the spacesHandH , for allθ∈[−1,1], which will be precisely defined later on in Definition6.4.

2.2. Space of initial conditions

Definition (1.13) is unnecessary restrictive. We have fixeds >1/2 in the introduction for readability conve- nience but we can also assume that s∈[0,1/2[ if d= 2. The limit cases = 1/2 may involve some technical difficulties so we choose to avoid it (see Rem. 3.4). In the whole following, we choose X(Ω) = Vs(Ω) for s∈[d−22 ,1]\{1/2} as the space of initial condition, and we introduce the operatorγs:Vs(Ω)→Vs−1/2(Γ) as follows.

Definition 2.1. For alls∈[0,1]\{1/2}, we define the linear operatorγs:Vs(Ω)→Vs−1/2(Γ) by γs(y) =

(y·n)n, if s∈[0,1/2[, y|Γ if s∈]1/2,1].

Proposition 2.2. The linear operator γs satisfies the following regularity properties:

γs∈ L(Vs(Ω), Vs−1/2(Γ)), s∈[0,1]\{1/2}. (2.2) Proof. This proposition is a straightforward consequence of the well known trace and normal trace

properties.

(9)

3. Main result

In this article, we prove the following local stabilization result.

Theorem 3.1. Let s∈ [d−22 ,1]\{1/2}. There exists two linear operators Π2 ∈ L(Vn0(Ω), V1/2(Γ)) and Π3 L(V−1/2(Γ), V1/2(Γ))such that, if we consider the following coupled system:

tz−νΔz+ (z· ∇)zs+ (zs· ∇)z+ (z· ∇)z+∇p= 0 and ∇ ·z= 0 in Q, (3.1)

tu−Δbu+ Π3u−σ n=−Π2P z in Σ, z=u on Σ,

Γ

u(t)·n= 0, t0, (3.2)

z(0) =z0∈Vs(Ω), u(0) =γs(z0), (3.3)

then the following results hold. There existc >0 andμ0>0 such that, if δ∈(0, μ0)and z0∈ Wδs=

z0∈Vs(Ω) | z0Vs(Ω)≤cδ

, (3.4)

then, (3.1)-(3.2)-(3.3)admits a unique solution in the set Dsδ =

(z, p, u, σ)∈W(0,∞;Vs+1(Ω), V0s−1(Ω))×Hs/2−1/2(0,∞;Hs(Ω)) (3.5)

×W(0,∞;Vs+1/2(Γ), Vs−3/2(Γ))×L2(0,∞) |

zW(0,∞;Vs+1(Ω),V0s−1(Ω))+uW(0,∞;Vs+1/2(Γ),Vs−3/2(Γ))+σL2(0,∞)≤δ, pHs/2−1/2(0,∞;Hs(Ω)) ≤δ(1 +δ)

. Moreover, there exist C >0 andη >0 such that(z, u)obeys

z(t)Vs(Ω)+u(t)Vs−1/2(Γ)≤Cz0Vs(Ω)e−ηt ∀t≥0. (3.6) Remark 3.2. The linear operators Π2 and Π3 are components of Π which is the solution to the Riccati equation (7.7) given later on in Section7, see Remark7.4.

Remark 3.3. In fact, assuming Ω of classC3 [16], Chapter 1, Definition 1.2.1.2, is sufficient to obtain Theo- rem3.1. Indeed, the assumption Ω of classC4 is only needed in the second step of the proof of Theorem10.2 of the appendix, to treat the case of a control localized on a part of the boundary (see Rem.10.3). Notice that with Ω of class C3 the trace space Vs(Γ) is well defined fors [0,3] [16], Chapter 1, Definition 1.3.3.2. See also Remark5.3.

Remark 3.4. We decide to avoid the limit cases= 1/2 only because it involves technical difficulties. In fact, Theorem3.1remains valid for initial condition (z(0), u(0)) = (z0, u0) belonging to

(z0, u0)∈V1/2(Ω)×V0(Γ)|z0−Du0∈V01/2(Ω), z0V1/2(Ω)+u0V0(Γ)≤cδ

, whereD is the lifting operator given in Section4. Hence, we have to replace (3.6) by

z(t)V1/2(Ω)+u(t)V0(Γ)≤C(z0V1/2(Ω)+u0V0(Γ)) e−ηt, t≥0.

Remark 3.5. Fors∈[d−22 ,1] andz0∈V0s(Ω), Theorem3.1holds with u(0) = 0.

(10)

As explained in the introduction, the case of a boundary control localized in an open part of Γ can be treated by introducing an adequate cut-off functionm∈C2(Γ), with values in [0,1]. We assume thatmis supported in ΓmΓ, and is equal to 1 in Γ1, where Γ1 is an open subset of Γm. We introduce the space of initial conditions

Vms(Ω) =

y∈Vs(Ω) | (1−m)γs(y) = 0

, (3.7)

and we prove the following localized version of Theorem3.1.

Theorem 3.6. Let s∈ [d−22 ,1]\{1/2}. There is two linear operators Πm,2 ∈ L(Vn0(Ω), V1/2(Γ)) and Πm,3 L(V−1/2(Γ), V1/2(Γ))such that, if we consider the following coupled system:

tz−νΔz+ (z· ∇)zs+ (zs· ∇)z+ (z· ∇)z+∇p= 0 and ∇ ·z= 0 in Q, (3.8) z=m(u−σm(u)n) on Σ, σm(u) =

Γ

m −1

Γ

mu·n, (3.9)

tu−Δbu+ Πm,3u−σ n=−Πm,2P z in Σ, (3.10)

z(0) =z0∈Vms(Ω), u(0) =γs(z0), (3.11)

then the following result holds. There exist c > 0 and μ0 > 0 such that, if δ (0, μ0) and z0 ∈ Wm,δs = Wδs∩Vms(Ω), then, (3.8)-(3.9)-(3.10)-(3.11)admits a unique solution in the set Dδs, which is defined by (3.5).

Moreover, there exist C >0 andη >0 such that(z, u)obeys(3.6).

Remark 3.7. The linear operators Πm,2and Πm,3are components of Πmwhich is the solution to the Riccati equation (9.24) given later on in Section9, see (9.25).

4. The Oseen system

The main objective of this section is to give a precise definition for the solution of the system:

tz−νΔz+ (z· ∇)zs+ (zs· ∇)z+∇p=f in QT, (4.1)

∇ ·z= 0 in QT, z=u on ΣT, z(0) =z0. (4.2) In the above setting, T (0,∞) is a fixed time horizon, QT = Ω×(0, T), ΣT = Γ×(0, T) and f L2(0, T;H−1(Ω)). By following the ideas introduced in [23], we will rewrite (4.1)-(4.2) as an evolution equation.

First, we introduce the following unbounded operators (D(A), A) and (D(A), A) inVn0(Ω):

D(A) = V02(Ω) and Ay=νPΔy−P(y· ∇)zs−P(zs· ∇)y, D(A) = V02(Ω) and Ay=νPΔy−P(∇zs)Ty+P(zs· ∇)y.

Here, (b·∇)a= (d

i=1bixiaj)1≤j≤dand (∇a)Tb= (d

i=1bixjai)1≤j≤d, and one can verify that (D(A), A) is the adjoint of (D(A), A) with respect to the pivot spaceVn0(Ω). Throughout the following we denote byλ0>0 an element in the resolvent set ofA satisfying:

0−A)y|yV−1

0 (Ω),V01(Ω)≥ν 2y2V1

0(Ω) for ally ∈V01(Ω). (4.3) Theorem 4.1. The unbounded operator (D(A), A) (resp. (D(A), A)) is the infinitesimal generator of an analytic semigroup on Vn0(Ω), and the characterization below holds:

D((λ0−A)θ) =D((λ0−A)θ) =V0(Ω) for allθ∈[0,1]. (4.4)

Proof. See [23], Lemma 4.1.

(11)

We now introduce the Dirichlet operatorD:V0(Γ)L2(Ω) defined as follows. Foru∈V0(Γ), set Du=w where (w, q) satisfies the following system:

λ0w−νΔw+ (w· ∇)zs+ (zs· ∇)w+∇q= 0, ∇ ·w= 0, w=u on Γ. (4.5) For rough datau∈V0(Γ), defining a solution to (4.5) can be done with the transposition method. It consists in looking for a velocityw∈L2(Ω) obeying:

Γ

(rn−ν∂nϕ) =

Ω

w·f for all f L2(Ω), (4.6)

where (ϕ, r)∈V02(Ω)×H1(Ω) is the unique pair satisfying

λ0ϕ−νΔϕ+ (∇zs)Tϕ−(zs· ∇)ϕ+∇r=f and ∇ ·ϕ= 0 in Ω, ϕ= 0 on Γ,

Γ

r= 0. (4.7) The existence and uniqueness of w L2(Ω) solution to (4.6) is a consequence of the Riesz representation theorem, and an integration by parts allows to prove that a smooth velocity (sayw∈H2(Ω) and u∈V3/2(Γ)) solution to (4.5) in a classical sense is also the solution to (4.6). Moreover, since a smooth solution satisfies w|Γ =u, a density argument ensures that this trace condition remains true when w∈Hs(Ω) fors >1/2:

(Du)|Γ =u if Du∈Hs(Ω), s >1/2. (4.8)

However, if we are only interested in rough solution w∈ L2(Ω), it is sufficient to consider a boundary value u∈V−1/2(Γ) in (4.6) (where the sign

Γ must be understood as a duality product), and we can verify that

∇ ·Du= 0 and (Du)|Γ·n=u·n. (4.9)

Here is the argument. By choosing f = ∇π in (4.6), successively forπ ∈H01(Ω) and for π ∈H1(Ω) obeying

Γπ= 0, we can deduce thatϕ= 0 and π=r from (4.7), and integrations by parts allow to recover the free divergence condition∇·w= 0 and the normal trace conditionw|Γ·n=u·n. About such a Dirichlet operatorD one may refer to [23], Appendix 2, from which the following proposition is taken.

Proposition 4.2. (i)The operator D is bounded fromV0(Γ)intoV0(Ω) and it satisfies

D∈ L(Vs−1/2(Γ), Vs(Ω)) for all s∈[0,2]. (4.10) (ii)The operator D∈ L(V0(Ω), V0(Γ)), the adjoint ofD, is defined by

Df =rn−ν∂nϕ, (ϕ, r)∈V02(Ω)×H1(Ω) satisfies(4.7). (4.11)

Proof. See [23], Appendix 2.

Remark 4.3. According to [23], Lemma 7.4, the operatorDbelongs toL(V0s(Ω), Vs+1/2(Γ)) for alls∈[0,2].

Hence, it allows to extendD by duality to an element ofL(Vs−1/2(Γ), Vs(Ω)) fors∈[2,0].

Remark 4.4. As in the proof of [4], Lemma 3.3.1, one can prove that every ϕ H2(Ω)H10(Ω) satisfies

nϕ·n=∇·ϕ|Γ. Hence, everyϕinV02(Ω) has a boundary normal derivativenϕ∈V1/2(Γ) which is tangential.

As a consequence, in (4.11)rnand −ν∂nϕare respectively the normal and the tangential component ofDf. In fact, the set of tangential boundary values inV1/2(Γ) is totally described by the normal derivatives of vector fields inV02(Ω): for allu∈V1/2(Γ) such thatu·n= 0, there existsϕu∈V02(Ω) which obeys

∇ ·ϕu= 0 in Ω, nϕu=u, ϕu= 0 on Γ and ϕuH2(Ω)≤cuV1/2(Γ), (4.12)

(12)

wherec >0 only depends on the geometry. Such a vector fieldϕucan be obtained as follows. In a first step, using a continuous right inverse of the trace and the boundary normal derivative operators [16], Theorem 1.5.1.5, we constructφuH2(Ω)∩H10(Ω) such thatnφu=u. Hence, by recalling thatuis tangential we havenφu·n= 0, and sinceφu|Γ = 0 yields the equality∇ ·φu|Γ =nφu·n, we deduce that∇ ·φu∈H01(Ω). Thus, it allows to constructζuH20(Ω) such that∇·ζu=−∇·φu[13], Chapter III, Theorem 3.2, and the vector fieldϕu=φuu satisfies (4.12).

Remark 4.5. In fact, the trace equality in (4.8) is still valid fors∈[0,1/2]. Indeed, as in [19], Theorem 6.5, Chapter 2, it can be proved that the trace operator can be extended to a continuous operator from {y V0(Ω) | νΔy−(y· ∇)zs(zs· ∇)y∈V0−2(Ω)} into V−1/2(Γ). Here is the argument. For allu∈V1/2(Γ) we construct a pair (ϕu, ru) ∈V02(Ω)×H1(Ω), depending continuously on u, and which obeys ru|Γ =u·n and

−ν∂nϕ=u−γn(u) (the tangential component of u, see Rem.4.4). Thus, for ally ∈V2(Ω) an integration by parts yields

Γ

(run−ν∂nϕu)·y=

Ω

(νΔy(y· ∇)zs(zs· ∇)y)·ϕu+

Ω

(∇ru−νΔϕu+ (∇zs)Tϕu(zs· ∇)ϕu), and by taking the supremum over allu=run−ν∂nϕu∈V1/2(Γ), the following estimate can be obtained:

y|ΓV−1/2(Γ)≤C(yV0(Ω)+νΔy−(y· ∇)zs(zs· ∇)yV−2

0 (Ω)).

Finally, it remains to extend the trace operator with a density argument.

We are now in position to state the following corollary.

Corollary 4.6. Let s∈[0,1]\{1/2}. The linear operatorγs∈ L(Vs(Ω), Vs−1/2(Γ)), which is given by Defini- tion 2.1, satisfies the following compatibility condition:

y−Dγs(y)∈V0s(Ω) for ally∈Vs(Ω). (4.13) Next, let us define solutions to (4.1)-(4.2).

Definition 4.7. Let z0 V0(Ω), u L2(0, T;V−1/2(Γ)) and f L2(0, T;V0−2(Ω)). We shall say that z∈L2(0, T;V0(Ω)) is a weak solution to (4.1)-(4.2), if and only if,

(i) P zis a weak solution of the evolution equation:

(P z) = AP z+ (λ0−A)P Du+f ∈L2(0, T;V0−2(Ω)), (4.14)

P z(0) = P z0 ∈Vn0(Ω). (4.15)

(ii) (I−P)z is defined by:

(I−P)z= (I−P)Dγn(u) ∈L2(0, T;V0(Ω)). (4.16) Remark 4.8. Let us underline that (4.16) can be reduced to

(I−P)z= (I−P)Du ∈L2(0, T;V0(Ω)). (4.17) Indeed, by remarking thatγn(u)−uis the tangential component of u, we haveD(γn(u)−u)∈Vn0(Ω) which gives (I−P)D(γn(u)−u) = 0, or equivalently (I−P)Dγn(u) = (I−P)Du.

Theorem 4.9. Let f ∈L2(0, T;H−1(Ω)),z0∈V0(Ω) andu∈C([0, T];V−1/2(Γ))obeyingz0·n=u(0)·n.

(i) If z W(0, T;V1(Ω), V0−1(Ω)) is a weak solution in the sense of Definition 4.7, associated with z0 and u, then there is a unique p H−1/2(0, T;L20(Ω)) such that (z, p) satisfies (4.1)-(4.2). Moreover, if z W(0, T;V2(Ω), V0(Ω)), then we have p∈L2(0, T;H1(Ω)).

Références

Documents relatifs

Odasso, Markov solutions for the 3D stochastic Navier Stokes equations with state dependent noise, Journal of Evolution Equations, Birkhuser Basel, vol.. Sinai, Gibbsian dynamics

The very first work concerning NBC was done by Solonnikov and ˇ Sˇ cadilov [10] for α = 0 where the authors considered stationary Stokes system with Dirichlet condition on some part

We analyze the problem mainly using the Bhatnagar–Gross–Krook (also known as BGK) model, but some results based on the original Boltzmann equation will also be presented. Because of

In Sections 4 and 5, we study the mixed finite volume approximation respectively for Stokes and Navier-Stokes equations: we show that this method leads to systems (linear in the case

The stabilizability of the constant density (or homogeneous) incompressible Navier-Stokes equation (with Dirichlet or mixed boundary condition) by a finite dimensional

We first study the local stabilization of the one dimensional compressible Navier-Stokes system, with periodic boundary conditions, by a distributed control... Let ω be any

One then applies methods of finite-dimensional theory (e.g., the pole assignment theorem [21, Section 2.5]) to find a stabilizing feedback control for the first equation and proves

Of course, neglecting the horizontal viscosity and requiring a lot of regularity on the initial data, one can prove the local existence of strong solutions by working in the frame