• Nenhum resultado encontrado

A special case with no boundary layer : the slip condition

No documento DE L’UNIVERSITÉ PIERRE ET MARIE CURIE (páginas 99-102)

Contrôle de Navier-Stokes avec des conditions au bord

4.3 A special case with no boundary layer : the slip condition

In this section, we prove Theorem8in the particular case where the friction coefficientAis the shape operatorM. In this setting, we can choose an Euler trajectory such that there is no boundary layer and the proof is thus much easier. This allows us to present some elements of our method in a simple setting before moving on to the general case which involves boundary layers. We consider the slip boundary condition on the uncontrolled boundary :

u·n= 0 and [∇ ×u]tan= 0 on Ω\Γ. (4.12)

As in [52], our strategy is to deduce the controllability of the Navier-Stokes equation in small time from the controllability of the Euler equation. In order to use this strategy, we are willing to trade small time against small viscosity using the usual fluid dynamics scaling. Note that, even in this easier context, Theorem 8 is new for multiply connected 2D domains and for all 3D domains since [52] only concerns simply connected 2D domains.

4.3.1 Small viscosity asymptotic expansion

The global controllability timeT is small but fixed. Let us introduce a positive parameterε1. We will be even more ambitious and try to control the system during the shorter time interval [0, εT]. We perform the scaling :uε(t, x) :=εu(εt, x) andpε(t, x) :=ε2p(εt, x). Now, (uε, pε) is the solution to the following system fort∈(0, T) :













tuε+ (uε· ∇)uεεuε+∇pε= 0 on (0, T)×Ω, divuε= 0 on (0, T)×Ω,

uε·n= 0 on (0, TΩ\Γ, [∇ ×uε]tan = 0 on (0, TΩ\Γ,

uε|t=0=εu on Ω.

(4.13)

Due to the scaling chosen, we need to prove that we can obtain|uε(T,·)|L2(Ω)=o(ε) if we want to achieve global approximate null controllability. Sinceε is small, we expectuε to converge to the solution of the Euler equation. Hence, we introduce the following asymptotic expansion foruε:

uε(t, x) =u0(t, x) +εu1(t, x) +εRε(t, x). (4.14)

The pressure is also expanded as :

pε(t, x) =p0(t, x) +εp1(t, x) +επε(t, x). (4.15) Let us provide some insight behind expansion (4.14)-(4.15). The first term (u0, p0) is the solution to an Euler equation. It models a smooth reference trajectory around which we are linearizing the Navier- Stokes equation. This trajectory will be chosen in such a way that it flushes the initial data out of the domain in timeT. The second term (u1, p1) takes into account the initial data u, which will be flushed out of the domain by the flowu0. Eventually,Rεcontains higher order residues. We need to prove

|Rε(T,·)|L2(Ω)=o(1) in order to be able to conclude with local results.

4.3.2 Euler’s equation

At orderO(1), the first part (u0, p0) of our expansion is a solution to the Euler equation. Hence, the pair (u0, p0) is a return-method-liketrajectory of the Euler equation on [0, T] :









tu0+ u0· ∇

u0+∇p0= 0, on (0, T)×Ω, divu0= 0 on (0, T)×Ω,

u0·n= 0 on (0, TΩ\Γ, u0(0,·) =u0(T,·) = 0 in Ω.

(4.16)

We want to use this reference trajectory to flush everything outside of our domain within the fixed time interval [0, T]. To make this statement precise, let us introduce the flow Φ associated withu0. This flow is defined by :

d

dtΦ(t, x) =u0(t,Φ(t, x)) and Φ(0, x) =x. (4.17) Hence, we look for trajectories satisfying :

x∈Ω, Φ(T, x)∈/. (4.18)

At this stage, note that equation (4.18) does not make much sense. Indeed, we need to start by introducing O, a smooth extension of Ω such that Ω\Γ ⊂O and Γ⊂ O. Outside of the physical domain Ω, we choose any smooth extension of the reference trajectory u0 (we do not assume that the extension is divergence-free in O \Ω since this set may have a smaller volume). However, we extend u0 such that u0·n= 0 on the whole ofO. This gives sense to (4.17) and (4.18) since the characteristics starting inside Ω can now leave Ω (but all characteristics remain withinO). This extension procedure is standard and can be justified by elementary methods. Indeed, we can assume that the restriction of Γ to each connected component ofΩ is simply connected (if this is not the case, we can shrink it until this condition is met).

∂Ω\Γ

u·n= 0

[D(u)n+Au]tan= 0 Γ

O

Figure 4.2 – Extension of the physical domain Ω.

Lemma 45. There exists a solution pair(u0, p0)∈ C([0, T]×Ω,¯ Rd×R)to system (4.16)such that the flowΦdefined in (4.17)satisfies (4.18). Moreover, u0 can be chosen such that :

∇ ×u0= 0 in [0, T]×Ω. (4.19)

Démonstration. This lemma is the key argument of multiple papers concerning the small time global exact controllability of Euler equations. We refer to the following references for detailed statements and construction of these reference trajectories. First, Coron used it in [51] for 2D simply connected domains, then in [53] for general 2D domains when Γ interesects all connected components ofΩ. Glass adapted the argument for 3D domains (when Γ intersects all connected components of the boundary), for simply connected domains in [92] then for general domains in [94]. He also used similar arguments to study the obstructions to approximate controllability in 2D when Γ does not intersect all connected components of the boundary in [95] for general 2D domains. Here, we use the assumption that our control domain Γ intersects all connected parts of the boundaryΩ. The fact that condition (4.19) can be achieved is a direct consequence of the construction of the reference profileu0as a potential flow :u0(t, x) =∇θ0(t, x), whereθ0is smooth.

4.3.3 Convective term and flushing of the initial data

We move on to order O(ε). Here, the initial datau comes into play. Letu1be the solution to :









tu1+ u0· ∇

u1+ u1· ∇

u0+∇p1= ∆u0, in Ω fort≥0, divu1= 0 in Ω fort≥0,

u1·n= 0 inΩ\Γ fort≥0, u1(0,·) =u in Ω att= 0.

(4.20)

In this paragraph, we exlain how we can use our control to reachu1(T,·) = 0. Remember that, thanks to the choice of u0, the initial data will be fully flushed outside of the domain. Equation (4.20) also takes into account a residual term ∆u0. The easiest path to prove that it is possible to controlu1is to introduce w1:=∇ ×u1 and to write (4.20) in vorticity form :

(tw1+ u0· ∇

w1+ w1· ∇

u0= 0, in Ω fort≥0,

w1(0,·) =∇ ×u in Ω att= 0. (4.21) Note that the term w1· ∇

u0is specific to the 3D setting and does not appear in 2D (where the vorticity is merely transported). Nevertheless, even in 3D, the support of the vorticity is transported. Thus, thanks to hypothesis (4.18),w1will vanish inside Ω at timeT provided that we choose null boundary conditions forw1on the controlled boundary Γ. Hence, we can build a trajectory of (4.20) such that∇×u1(T,·) = 0.

Since, we can choose the boundary controls on Γ to be null at timeT, U :=u1(T,·) satisfies :





∇ ·U = 0 in Ω,

∇ ×U = 0 in Ω, U·n= 0 on.

(4.22)

For simply connected domains, this impliesU ≡0. For multiply connected domains, the situation is more complex. We refer to the demonstrations given in [53] and [94] which prove that, thanks to careful choices of the control and the reference trajectory, we can still reachu1(T,·) = 0. The interested reader can also start with the nice introduction given by Glass in [93].

Moreover, if we assume thatuH3(Ω) (which can be done without loss of generality as exposed in Section4.6),u1 can be constructed inL([0, T];H3(Ω)).

4.3.4 Size of the remainder

The equation for the remainder reads :













tRεεRε+ (uε· ∇)Rε+AεRε+∇πε=FεGε, in (0, T)×Ω, divRε= 0 in (0, T)×Ω, [∇ ×Rε]tan=−

∇ ×u1

tan in (0, TΩ\Γ, Rε·n= 0 in (0, TΩ\Γ, Rε(0,·) = 0 in Ω att= 0,

(4.23)

where :

AεRε= (Rε· ∇) u0+εu1

, Fε=εu1, Gε=ε(u1· ∇)u1. (4.24) We want to establish a standard L(L2)∩L2(H1) energy estimate for the remainder. Let g1 :=

∇ ×u1

tan. As usual, we multiply equation (4.23) by Rε and integrate by parts. Let us recall that Rε·n= 0 and divuε= divRε= 0, which simplifies most terms. We obtain :

1 2

d dt

Z

|Rε|2+ 2ε Z

|D(Rε)|2= Z

(Fε+Gε)Rε+ Z

AεRεRε−2ε Z

(g1+M Rε)Rε. (4.25) To integrate (∆Rε)Rεby parts, we used the formula :

− Z

f·g= 2 Z

D(f)D(g)−2 Z

D(f)n·g, (4.26)

which is valid as long asf is divergence-free (see [91, Lemma C.1]). Let us estimate the boundary integral by transforming it into an interior term. Choosing a smooth extension ¯M ofM to Ω, we write :

2ε Z

(g1+M Rε)Rε

= 2ε Z

div

(g1+ ¯M RεRε n

εC |Rε|L2+|g1|H1

|Rε|H1

εC |Rε|L2+|g1|H1

(|Rε|L2+|D(Rε)|L2)

εC |Rε|2L2+|g1|2H1

+ε|D(Rε)|2L2,

(4.27)

where we used the second Korn inequality to estimate theH1norm ofRεusingD(Rε) (see [127, Theorem 10.2, page 299]) and the constantsC depends on the domain andC1 norms ofnand ¯M. Plugging (4.27) into (4.25) yields :

1 2

d dt

Z

|Rε|2+ε Z

|D(Rε)|2

|Aε|+1 2+εC

Z

|Rε|2+ε|g1|2H1+ Z

|Fε|2+|Gε|2. (4.28) Applying the Gronwall inequality by integrating over (0, T) and using the null initial condition gives :

|Rε|2L(L2)+ε|D(Rε)|2L2(L2)εC, (4.29) whereCdepends on

u0 ,

u1 ,

u1

L1(L2) and g1

L1(H1). Hence,|Rε(T,)|L2 =O(√

ε) and this concludes the proof of the approximate null controllability sinceu0(T) =u1(T) = 0.

No documento DE L’UNIVERSITÉ PIERRE ET MARIE CURIE (páginas 99-102)