Controllability Results For Cascade Systems of M C
Controllability Results For Cascade Systems of M C
∗1 †2
Takéo Takahashi , Luz de Teresa , and Ying Wu-Zhang‡2
1
Université de Lorraine, CNRS, Inria, IECL, F-54000 Nancy, France
2
Instituto de Matematicas, Universidad Nacional Autónoma de México, Circuito Exterior, C.U.,
C. P. 04510 Ciudad de México, México
Contents
1 Introduction 2
2 Preliminaries 4
Abstract
In this paper we deal with the controllability properties of a system of m coupled Stokes systems or m coupled
Navier-Stokes systems. We show the null-controllability of such systems in the case where the coupling is in
a cascade form and when the control acts only on one of the systems. Moreover, we impose that this control
has a vanishing component so that we control a m × N state (corresponding to the velocities of the fluids) by
N − 1 distributed scalar controls. The proof of the controllability of the coupled Stokes system is based on a
Carleman estimate for the adjoint system. The local null-controllability of the coupled Navier-Stokes systems is
then obtained by means of the source term method and a Banach fixed point.
1
1 Introduction
Controllability issues related to a single parabolic equation or to a single Stokes or Navier-Stokes system have been
intensively studied in the last fifty years giving rise to interesting techniques, new challenges and open problems.
See some seminal results [12,16,25] for the heat equation and [8,14,21] for the Navier-Stokes system. The literature
is vast and it is difficult to mention all the intensive studies about this subject. However, it is only in the last fifteen
years that the challenging issue of controlling coupled parabolic systems has attracted the interest of the control
community. This kind of systems appears mathematically in optimal control theory as a characterization of the
optimal control (with one equation coupled to its adjoint) but also appears, for example, in the study of chemical
reactions (see e.g. [11], [7]), and in a wide variety of mathematical biology and physical situations (see e.g. [20] ). In
the case of scalar (heat) coupled equations an important number of challenging problems has been solved (see [1] for
a survey of results until 2011) and sometimes the results have been surprising [2–4]. In the case of coupled Stokes
or Navier-Stokes systems, to our knowledge, only some cases of two coupled systems have been treated [5, 6, 18, 29].
Here our aim is to generalize results for a m scalar cascade system [17] to a m N -dimensional Stokes or Navier-Stokes
cascade system but including an extra deal: to eliminate one component on the N -dimensional control.
Let us be more specific: we consider a bounded domain Ω of RN (N = 2, 3) whose boundary ∂Ω is regular enough.
Let T > 0 and let ω ⊂ Ω be a (arbitrary small) nonempty open subset which will usually be referred as the control
domain. We will use the notation Q = Ω × (0, T ) and Σ = ∂Ω × (0, T ).
In this article, we are interested in the null controllability of a coupled system of m Stokes or Navier-Stokes
systems, with m > 2:
Pm
∂t y (i) − ∆y (i) + ∇p(i) + ε y (i) · ∇ y (i) = Ai,j y (j) + Di v1ω in Q, (1 6 i 6 m)
j=1
∇ · y (i) = 0 in Q, (1 6 i 6 m)
(i)
y = 0 on Σ, (1 6 i 6 m)
(i) (i)
y (·, 0) = y0 in Ω, (1 6 i 6 m)
with ε = 0 for Stokes systems and ε = 1 for Navier-Stokes systems, where Ai,j ∈ MN (R) and where Di ∈ MN,r (R)
for some r ∈ N∗ . We have denoted by 1ω the characteristic function of ω. Let us notice that we take in this work
the viscosities of the fluids constant and equal to 1 to simplify.
We can write the above systems in a more compact way as
∂t y − ∆y + ∇p = Ay + Dv1ω in Q,
∇ · y = 0 in Q,
(1.1)
y=0 on Σ,
y|t=0 = y0 in Ω,
or
∂t y − ∆y + (y · ∇)y + ∇p = Ay + Dv1ω
in Q,
∇ · y = 0 in Q,
(1.2)
y = 0
on Σ,
y|t=0 = y0 in Ω,
where we set
y = y (1) , . . . , y (m) , p = p(1) , . . . , p(m) , y0 = y0(1) , . . . , y0(m) ,
∆y = ∆y (1) , . . . , ∆y (m) , ∇p = ∇p(1) , . . . , ∇p(m) ,
(y · ∇) y = y (1) · ∇ y (1) , . . . , y (m) · ∇ y (m) , ∇ · y = ∇ · y (1) , . . . , ∇ · y (m)
and
P
m P
m
Ay = A1,j y (j) , . . . , Am,j y (j) Dv = D1 v, . . . , Dm v .
j=1 j=1
2
In this work, we will focus in the particular case where the partitioned matrix A has the form
A1,1 A1,2 A1,3 ... A1,m
A2,1 A2,2 A2,3 ... A2,m
0 A3,2 A3,3 ... A3,m
A= (1.3)
.. .. .. .. ..
. . . . .
0 0 . . . Am,m−1 Am,m
with all the blocks under the diagonal non zero and such that all its blocks are scalar matrices. More precisely, our
hypotheses on A are
Ai,j = ai,j IN , ai,i−1 6= 0 (2 6 i 6 m), ai,j = 0 if i > j + 2. (1.4)
We also control (1.1) or (1.2) by acting only on one of the Stokes or of the Navier-Stokes systems, for instance the
first one, and with N − 1 scalar controls on this system. Thus, without loss of generality, we assume
1 0
1
r = N − 1, Dj = 0 (j > 2), D1 = (if N = 2) or D1 = 0 1 (if N = 3). (1.5)
0
0 0
The above choice of the matrix A corresponds to a particular coupling considered in the context of the null-
controllability of systems of m linear heat equations, see [17] and our aim is to extend this result in the case of
coupled of Stokes or Navier-Stokes systems.
In order to state our main results, we recall some standard functional spaces associated with the Stokes system:
H = y ∈ L2 (Ω)N : ∇ · y = 0 in Ω, y · n = 0 on ∂Ω (1.6)
and
V = y ∈ H01 (Ω)N : ∇ · y = 0 in Ω . (1.7)
Our main result is the following theorem.
m
Theorem 1.1. Assume (1.4) and (1.5). Then, for any T > 0 and for any
2 2 N −1 (1) (m)
y0 ∈ H , there exists a control
v ∈ L (0, T ; L (ω) ) such that the corresponding solution y = y , ..., y to (1.1) satisfies
y(·, T ) = 0 in Ω.
Remark 1.2. As a consequence, we deduce that we can control the system (1.1) of N × m scalar equations with
N − 1 scalar controls.
From Theorem 1.1 and a general method to deal with the controllability of nonlinear parabolic systems, we
deduce the local null controllability of the system (1.2):
Theorem 1.3. Assume (1.4) and (1.5). Then, for any T > 0, there exists δ > 0 such that, for any y0 ∈ V m
satisfying
ky0 kV m 6 δ,
there exists v ∈ L2 (0, T ; L2(ω)N −1 ) such that the corresponding solution y to (1.2) satisfies
y(·, T ) = 0 in Ω.
In order to prove Theorem 1.1, we introduce the adjoint system:
−∂t ϕ(i) − ∆ϕ(i) + ∇π (i) = P Aj,i ϕ(j)
i+1
in Q, (1 6 i 6 m − 1)
j=1
−∂ ϕ(m) − ∆ϕ(m) + ∇π (m) = P A ϕ(j)
m
t j,m in Q,
j=1 (1.8)
∇ · ϕ(i) = 0 in Q, (1 6 i 6 m)
ϕ(i) = 0
on Σ, (1 6 i 6 m)
(i) (i)
ϕ (·, T ) = ϕT in Ω, (1 6 i 6 m)
3
(i) (i)
with ϕT ∈ H (1 6 i 6 m) and we denote by ϕj , j = 1, . . . , N the coordinates of ϕ(i) . Note that by setting
Following a standard duality argument (see, for instance, [34, Theorem 11.2.1, p.357]), Theorem 1.1 will be obtained
as a consequence of the following observability inequality:
Z X
m 2 ZZ N
X −1
(i) (1) 2
ϕ (x, 0) dx 6 C(T ) ϕj dxdt. (1.10)
Ω i=1 ω×(0,T )
j=1
2 Preliminaries
In the whole article, we use the notation C for a generic positive constant that depends on Ω, ω. We can assume to
simplify that T ∈ (0, 1) and that will allow us to avoid some dependence on T for some constants. Finally we only
write the proof in the case N = 2, the case N = 3 can be done similarly.
To write our Carleman inequalities, we introduce standard weights and functions. First, we consider a nonempty
domain ω0 such that ω0 ⊂ ω. Then, using [16] (see also [34, Theorem 9.4.3, p.299]), there exists η 0 ∈ C 2 (Ω)
satisfying
η 0 > 0 in Ω, η 0 = 0 on ∂Ω, max η 0 = 1, ∇η 0 6= 0 in Ω \ ω0 . (2.1)
Ω
4
where ℓ > 14, λ > 1. In literature ℓ > 3 is enough. However, to prove our result we need to use ℓ > 14 (see Proof
of Proposition 3.2).
Note that we have the following useful relations: there exists C > 0 depending on Ω such that
|∂t α| + |∂t ξ| 6 CT ξ 1+1/ℓ , (2.4)
∗ ′ ∗ ′
(α ) + (ξ ) 6 CT (ξ ∗ )1+1/ℓ , (α∗ )′′ + (ξ ∗ )′′ 6 CT 2 (ξ ∗ )1+2/ℓ , (2.5)
C
ξ∗ >
, (2.6)
T 2ℓ
|∇α| = |∇ξ| 6 Cλξ, |∆α| = |∆ξ| 6 Cλ2 ξ. (2.7)
Weights of the kind (2.2) were first considered in [16]. In its present form, these weights have already been used
in [19] in order to obtain Carleman estimates for the controllability of strongly coupled parabolic equations and
later in [18] for the existence of insensitizing controls for Stokes systems.
Now, we recall some standard results. The first one is a Carleman estimate for the gradient operator, it is stated
and proved in [9]:
Lemma 2.1. Let r ∈ R. There exists C > 0 depending only on r, Ω and ω0 such that, for every T > 0, s > C and
every u ∈ L2 (0, T ; H 1(Ω)),
ZZ ZZ ZZ
e−2sα sr+2 ξ r+2 |u|2 dxdt 6 C e−2sα sr ξ r |∇u|2 dxdt + e−2sα sr+2 ξ r+2 |u|2 dxdt. .
Q Q ω0 ×(0,T )
The second result is a Carleman estimate for the heat equation with non-homogeneous Dirichlet boundary con-
ditions. It is proved in [24].
Lemma 2.2. There exists a constant C > 0 such that for any λ > C, s > C, f0 , f1 , ..., fN ∈ L2 (Q) and
u ∈ L2 0, T ; H 1(Ω) ∩ H 1 0, T ; H −1(Ω)
satisfying
N
X
∂t u − ∆u = f0 + ∂j fj in Q,
j=1
we have
ZZ ZZ
2 2 2
e−2sα s−1 ξ −1 |∇u| + sξ |u| dxdt 6 C e−2sα sξ |u| dxdt
ω0 ×(0,T )
Q
2
2
∗
∗
+
e−sα s−1/4 (ξ ∗ )−1/4+1/ℓ u
2 +
e−sα s−1/4 (ξ ∗ )−1/4 u
1 1
L (Σ) H 4 , 2 (Σ)
ZZ N ZZ
X
2 2
+ e−2sα s−2 ξ −2 |f0 | dxdt + e−2sα |fj | dxdt . (2.8)
Q j=1 Q
5
This result can be found in [33].
Lemma 2.3. Assume T > 0. For any f ∈ L2 (Q)N , there exists a unique solution
to the Stokes system (2.9) and there exists a constant C > 0 depending only on Ω such that
kuk2L2 (0,T ;H 2 (Ω)N ) + kuk2H 1 (0,T ;L2 (Ω)N ) 6 C kf k2L2 (0,T ;L2 (Ω)N ) . (2.10)
In order to prove the above proposition, we first start by estimating each I1 (s, ϕ(i) ) (i = 1, . . . , m) independently.
6
Proposition 3.2. Let ωb ⊂ Ω be a nonempty open set such that ω0 ⋐ ω b . Then, there exists a constant C such that
for any s > C and for any ϕT ∈ H m , the solution ϕ of (1.8) satisfies
ZZ
(i) (i) 2
I1 (s, ϕ ) 6 C e−2sα s5 ξ 5 ∆ϕ1 dxdt
ω
b ×(0,T )
m ZZ
X 2 2
(j)
dxdt
∗
+ e−2sα s9/2 (ξ ∗ )5 ϕ(j) + e−2sα s−2 ξ −2 ∇2 ∆ϕ1 (1 6 i 6 m). (3.3)
j=1 Q
Proof of Proposition 3.2. First taking the divergence of (1.9), we remark that
∆π (i) = 0 in Q (1 6 i 6 m).
Thus, using a standard method for the Stokes system (see, for instance [9]), we apply the operator ∇2 ∆ to the first
components of (1.9), and we deduce
i+1
X
(i) (i) (j)
− ∂t ∇2 ∆ϕ1 − ∆∇2 ∆ϕ1 = aj,i ∇2 ∆ϕ1 in Q, (1 6 i 6 m). (3.4)
j=1
(i)
Then, we apply Lemma 2.1 with u = ∆ϕ1 and r = 3: for any s > C and λ > C,
ZZ 2 ZZ 2 ZZ 2
(i) (i) (i)
e−2sα s5 ξ 5 ∆ϕ1 dxdt 6 C e−2sα s3 ξ 3 ∇∆ϕ1 dxdt + e−2sα s5 ξ 5 ∆ϕ1 dxdt . (3.7)
Q Q ω0 ×(0,T )
7
Now, using the divergence condition of ϕ(i) , we have
(i) (i) (i)
∂2 ϕ2 = ∂1 ϕ1 6 ∇ϕ1 .
Then using the Poincaré inequality and the ellipticity of the Laplace operator with Dirichlet boundary conditions,
we deduce the existence of a constant C depending on Ω such that
Z Z
(i) 2 (i) 2
ϕ dx 6 C ∆ϕ1 dx.
Ω Ω
Combining the above relation with (3.5), (3.6) and (3.7), we deduce that I1 (s, ϕ(i) ) defined by (3.1) satisfies
ZZ 2 2
(i) (i) (i) 2
I1 (s, ϕ(i) ) 6 C e−2sα sξ ∇2 ∆ϕ1 + s3 ξ 3 ∇∆ϕ1 + s5 ξ 5 ∆ϕ1 dxdt
ω0 ×(0,T )
2
2
∗ (i)
∗ (i)
+
e−sα s−1/4 (ξ ∗ )−1/4+1/ℓ ∇2 ∆ϕ1
2 +
e−sα (sξ ∗ )−1/4 ∇2 ∆ϕ1
1 1
L (Σ)4 H 4 , 2 (Σ)4
i+1 Z Z
X 2
(j)
+ e−2sα s−2 ξ −2 ∇2 ∆ϕ1 dxdt . (3.8)
j=1 Q
Step 2. In this step, we get rid of the boundary terms in the right-hand side of (3.8). To estimate the first term,
we notice that since ℓ > 4, (ξ ∗ )−1/4+1/ℓ is bounded in (0, T ) (see (2.6)). Thus
2
2
−sα∗ −1/4 ∗ −1/4+1/ℓ 2 (i)
∗ (i)
e s (ξ ) ∇ ∆ϕ1
2 4 6 Cs−1/2
e−sα ∇2 ∆ϕ1
2 4
L (Σ) L (Σ)
2
∗ (i)
∗ (i)
∗ (i)
6 Cs−1/2
e−sα ∇2 ∆ϕ1
2 4 +
s1/2 e−sα (ξ ∗ )1/2 ∇2 ∆ϕ1
2 4
s−1/2 e−sα (ξ ∗ )−1/2 ∇3 ∆ϕ1
2 8
L (Q) L (Q) L (Q)
ZZ
2 2
∗ (i) (i)
6 Cs−1/2 e−2sα sξ ∇2 ∆ϕ1 + s−1 ξ −1 ∇3 ∆ϕ1 dxdt 6 Cs−1/2 I1 (s, ϕ(i) ). (3.9)
Q
For the second boundary term, we will use a trace inequality (see, for instance, [27, Theorem 2.1, p.9]) and by
an interpolation argument (see, for instance, [26, Corollary 9.2, p.46]), we have
2
2
∗ −1/4 −sα∗ 2 (i)
∗ (i)
(ξ ) e ∇ ∆ϕ1
1,1 6 C
(ξ ∗ )−1/4 e−sα ∇2 ∆ϕ1
H 4 2 (Σ)4 L2 (0,T ;H 1 (Ω)4 )∩H 1/2 (0,T ;L2 (Ω)4 )
2
∗ (i)
6 C
(ξ ∗ )−1/4 e−sα ∇2 ∆ϕ1
2 . (3.10)
L (0,T ;H (Ω) )∩H (0,T ;H −1 (Ω)4 )
1 4 1
Our goal is to estimate the last term. First we consider the function
∗
θ1 (t) := s3/2 (ξ ∗ )3/2−1/ℓ e−sα .
From (2.5)-(2.6), for s > C, ∗
|θ1′ | 6 Cs5/2 (ξ ∗ )5/2 e−sα . (3.11)
(i) (i)
Then, from (1.8), (θ1 ϕ , θ1 π ) is the solution of
(i)
(i)
(i)
i+1
P
−∂ t θ 1 ϕ − ∆ θ 1 ϕ + ∇ θ 1 π = aj,i θ1 ϕ(j) − θ1′ ϕ(i) in Q,
j=1
∇ · θ1 ϕ(i) = 0 in Q, (3.12)
θ1 ϕ(i) = 0 on Σ,
θ1 ϕ(i) |t=T = 0 in Ω.
8
Applying Lemma 2.3 to the above system and using (3.11), we obtain
2
θ1 ϕ(i)
L2 (0,T ;H 2 (Ω)2 )∩H 1 (0,T ;L2 (Ω)2 )
i+1
X
2
3/2 ∗ 3/2 −sα∗ (j)
2
∗
.
6C
s (ξ ) e ϕ
+
s5/2 (ξ ∗ )5/2 e−sα ϕ(i)
L2 (Q)2 L2 (Q)2
j=1
Second, we introduce ∗
θ2 = s(ξ ∗ )1−3/(2ℓ) e−sα .
From (2.5)-(2.6), for s > C, ∗
|θ2′ | 6 Cs2 (ξ ∗ )2−1/(2ℓ) e−sα . (3.14)
Then, from (1.8), (θ2 ϕ(i) , θ2 π (i) ) is the solution of
i+1
P
−∂t θ2 ϕ(i) − ∆ θ2 ϕ(i) + ∇ θ2 π (i) = aj,i θ2 ϕ(j) − θ2′ ϕ(i) in Q,
j=1
∇ · θ2 ϕ(i) = 0 in Q, (3.15)
θ2 ϕ(i) = 0 on Σ,
θ2 ϕ(i) |t=T = 0 in Ω.
Using that the right-hand side of the first equation is divergence free and with null trace on ∂Ω, we can apply
Lemma 2.3 to the above system and we deduce
!
2 i+1
X
2
(i)
(j)
′ (i)
2
θ2 ϕ
6C
θ2 ϕ
+
θ2 ϕ
.
L2 (0,T ;H 3 (Ω)2 )∩H 1 (0,T ;V ) L2 (0,T ;V ) L2 (0,T ;V )
j=1
∗
Using θ2 6 Cs2 (ξ ∗ )2−1/(2ℓ) e−sα and combining the above result with (3.14) and (3.13)
2 m
X
5/2 ∗ 5/2 −sα∗ (j)
2
θ2 ϕ(i)
2 6C
s (ξ ) e ϕ
. (3.16)
L (0,T ;H 3 (Ω)2 )∩H 1 (0,T ;V ) 2 L (Q)2
j=1
Finally, we introduce ∗
θ3 = (ξ ∗ )−5/(2ℓ) e−sα .
From (2.5)-(2.6), for s > C, ∗
|θ3′ | 6 Cθ2 , |θ3′′ | 6 Cs2 (ξ ∗ )2−1/(2ℓ) e−sα . (3.17)
(i) (i)
Then, from (1.8), (θ3 ϕ , θ3 π ) is the solution of
(i)
(i)
(i)
i+1
P
−∂ t θ 3 ϕ − ∆ θ 3 ϕ + ∇ θ 3 π = Aj,i θ3 ϕ(j) − θ3′ ϕ(i) in Q,
j=1
∇ · θ3 ϕ(i) = 0 in Q, (3.18)
θ3 ϕ(i) = 0 on Σ,
θ3 ϕ(i) |t=T = 0 in Ω.
9
We can again apply Lemma 2.3 to the above estimate to deduce
2
θ3 ϕ(i)
L2 (0,T ;H 5 (Ω)2 )∩H 1 (0,T ;H 3 (Ω)2 )∩H 2 (0,T ;V )
Xi+1
2
2
6C
θ3 ϕ(j)
2 +
∂t θ3 ϕ(j)
2
L (0,T ;H 3 (Ω)2 ) L (0,T ;H 1 (Ω)2 )
j=1
2
2
+
θ3′ ϕ(i)
2 +
∂t θ3′ ϕ(i)
2 . (3.19)
L (0,T ;H 3 (Ω)2 ) L (0,T ;H 1 (Ω)2 )
Using θ3 6 Cθ2 and applying (3.17), (3.16), and (3.13), we deduce from (3.19)
2 Xm
5/2 ∗ 5/2 −sα∗ (j)
2 .
θ3 ϕ(i)
2 5 2 1 3 2 2
6 C
s (ξ ) e ϕ
2 (3.20)
L (0,T ;H (Ω) )∩H (0,T ;H (Ω) )∩H (0,T ;V ) L (Q)2
j=1
Using the trace inequality (3.10) and that ℓ > 14, we have
2
2
∗ −1/4 −sα∗ 2 (i)
∗ (i)
(ξ ) e ∇ ∆ϕ1
1 , 1 6 C
(ξ ∗ )−1/4 e−sα ∇2 ∆ϕ1
H 4 2 (Σ)4 L2 (0,T ;H 1 (Ω)4 )∩H 1 (0,T ;H −1 (Ω)4 )
2
2
∗ (i)
6 C
(ξ ∗ )−1/4 e−sα ϕ1
6 C
θ3 ϕ(i)
. (3.21)
L2 (0,T ;H 5 (Ω))∩H 1 (0,T ;H 3 (Ω)) L2 (0,T ;H 5 (Ω)2 )∩H 1 (0,T ;H 3 (Ω)2 )
Putting together (3.8), the above relation and (3.9), we deduce at this step the following inequality:
ZZ 2 2
(i) (i) (i) 2
I1 (s, ϕ(i) ) 6 C e−2sα sξ ∇2 ∆ϕ1 + s3 ξ 3 ∇∆ϕ1 + s5 ξ 5 ∆ϕ1 dxdt
ω0 ×(0,T )
m ZZ
X 2 2
(j)
dxdt . (3.22)
∗
+ e−2sα s9/2 (ξ ∗ )5 ϕ(j) + e−2sα s−2 ξ −2 ∇2 ∆ϕ1
j=1 Q
Step 3. To estimate the local terms, we proceed in a standard way: we consider ω1 an open subset satisfying
ω0 ⋐ ω1 ⋐ ω
b and
η1 ∈ Cc2 (ω1 ), η1 ≡ 1 in ω0 , η1 > 0.
Then, an integration by parts gives
ZZ 2 ZZ 2
−2sα ∂2 (i) −2sα ∂2 (i)
e sξ ∆ϕ1 dxdt 6 η1 e sξ ∆ϕ1 dxdt
ω0 ×(0,T ) ∂xk ∂xq ω1 ×(0,T ) ∂xk ∂xq
ZZ
∂ ∂2 (i) ∂ (i)
=− η1 e−2sα sξ ∆ϕ1 ∆ϕ1 dxdt
ω1 ×(0,T ) ∂xk ∂xk ∂xq ∂xq
ZZ
∂3 (i) ∂ (i)
− η1 e−2sα sξ 2 ∆ϕ1 ∆ϕ1 dxdt.
ω1 ×(0,T ) ∂xk ∂xq ∂x q
10
Using (2.7) and Young’s inequality, we deduce from the above relation that there exists C > 0 such that for all
ε > 0,
ZZ 2 ZZ 2 2
−2sα 2 (i) −2sα −1 −1 3 (i) 2 (i)
e sξ ∇ ∆ϕ1 dxdt 6 ε e s ξ ∇ ∆ϕ1 + sξ ∇ ∆ϕ1 dxdt
ω0 ×(0,T )
Q
ZZ 2
C (i)
+ e−2sα s3 ξ 3 ∇∆ϕ1 dxdt. (3.23)
ε ω1 ×(0,T )
(i)
Now we estimate, in an analogous way, the local term associated with ∇∆ϕ1 : we consider
η2 ∈ Cc2 (b
ω ), η2 ≡ 1 in ω1 , η2 > 0.
Using (2.7) and the Young’s inequality, we deduce from the above relation that there exists C > 0 such that for
all ε > 0,
ZZ 2 ZZ 2 2
(i) (i) (i)
e−2sα s3 ξ 3 ∇∆ϕ1 dxdt 6 ε e−2sα s3 ξ 3 ∇∆ϕ1 + sξ ∇2 ∆ϕ1 dxdt
ω1 ×(0,T )
Q
ZZ
C (i) 2
+ e−2sα s5 ξ 5 ∆ϕ1 dxdt. (3.24)
ε ω
b ×(0,T )
The above estimate, together with (3.22) and (3.23) implies (3.3). This concludes the proof of Proposition 3.2.
We are now in a position to prove Theorem 3.1.
Proof of Theorem 3.1. Summing (3.3) for i = 1, . . . , m, and taking s > C for a constant C large enough, we deduce
that
m
X Xm ZZ
(i) (i) 2
I1 (s, ϕ ) 6 C e−2sα s5 ξ 5 ∆ϕ1 dxdt. (3.25)
i=1 i=1 ω
b ×(0,T )
In order to get rid of the local terms in the right-hand side (except the term corresponding to i = 1), we introduce
a sequence of open sets Oi , (0 6 i 6 m) such that
b =: O0 ⋐ O1 ⋐ ... ⋐ Oi ⋐ ... ⋐ Om ⋐ ω
ω
and functions
ζi ∈ Cc2 (Oi ) such that ζi ≡ 1 in Oi−1 , ζi > 0 (1 6 i 6 m).
Then, we consider the equation m−1 of (1.9), we apply the Laplace operator on the first component of this equation
and we multiply it by ζ1 :
m−1
X
(m) (m−1) (m−1) (j)
ζ1 am,m−1 ∆ϕ1 = −ζ1 ∂t ∆ϕ1 − ζ1 ∆2 ϕ1 − ζ1 aj,m−1 ∆ϕ1 . (3.26)
j=1
11
Then, using the above equation combined with (1.4) and integrating by parts, we deduce
ZZ ZZ
(m) 2 (m) 2
e−2sα s5 ξ 5 ∆ϕ1 dxdt 6 ζ1 e−2sα s5 ξ 5 ∆ϕ1 dxdt
ω
b ×(0,T ) O1 ×(0,T )
ZZ m−1
X
−1 (m) (m−1) (m−1) (j)
= ζ1 e−2sα s5 ξ 5 ∆ϕ1 ∂t ∆ϕ1 + ∆2 ϕ1 + aj,m−1 ∆ϕ1 dxdt. (3.27)
am,m−1 O1 ×(0,T ) j=1
Integrating by parts and applying ∆ to the first component of the equation (1.9) with i = m, we obtain
ZZ ZZ
−2sα 5 5 (m) (m−1) (m) (m−1)
ζ1 e s ξ ∆ϕ1 ∂t ∆ϕ1 dxdt = − ζ1 ∂t e−2sα s5 ξ 5 ∆ϕ1 ∆ϕ1 dxdt
O1 ×(0,T ) O1 ×(0,T )
ZZ
(m) (m−1)
+ ζ1 e−2sα s5 ξ 5 ∆2 ϕ1 ∆ϕ1 dxdt
O1 ×(0,T )
m ZZ
X (j) (m−1)
+ ζ1 e−2sα s5 ξ 5 aj,m ∆ϕ1 ∆ϕ1 dxdt. (3.28)
j=1 O1 ×(0,T )
Let us now estimate the right-hand side of (3.28) and (3.29) and the last term in (3.27).
Using (2.4), we have
∂t e−2sα s5 ξ 5 6 Ce−2sα s6 ξ 6+1/ℓ .
Combining this inequality and Young’s inequality, we deduce the existence of C such that for any s > C and for
any ε > 0,
Z Z
−2sα 5 5 (m) (m−1)
ζ1 ∂t e s ξ ∆ϕ1 ∆ϕ1 dxdt
O1 ×(0,T )
ZZ ZZ
(m) 2 C (m−1) 2
6ε e−2sα s5 ξ 5 ∆ϕ1 dxdt + e−2sα s7 ξ 7+2/ℓ ∆ϕ1 dxdt. (3.30)
ε
Q O1 ×(0,T )
12
Using (2.7), we have
∇(ζ1 e−2sα s5 ξ 5 ) 6 Ce−2sα s6 ξ 6 .
Combining this inequality and Young’s inequality, we deduce the existence of C such that for any s > C and for
any ε > 0,
Z Z
−2sα 5 5 (m) (m−1)
2∇ ζ1 e s ξ · ∇∆ϕ1 ∆ϕ1 dxdt
O1 ×(0,T )
ZZ 2 ZZ
(m) C (m−1) 2
6ε e−2sα s3 ξ 3 ∇∆ϕ1 dxdt + e−2sα s9 ξ 9 ∆ϕ1 dxdt. (3.32)
ε
Q O1 ×(0,T )
m ZZ
X (m−1)
ζ1 e −2sα 5 5 (j)
s ξ aj,m ∆ϕ ∆ϕ1 dxdt
j=1 O1 ×(0,T )
Xm ZZ ZZ
(j) 2 C (m−1) 2
6ε e−2sα s5 ξ 5 ∆ϕ1 dxdt + e−2sα s5 ξ 5 ∆ϕ1 dxdt, (3.34)
j=1
ε
Q O1 ×(0,T )
ZZ m−1
X
−1 (m) (j)
ζ1 e −2sα 5 5
s ξ ∆ϕ1 aj,m−1 ∆ϕ1 dxdt
am,m−1
O1 ×(0,T ) j=1
ZZ m−1 Z Z
(m) 2 C X (j) 2
6ε e−2sα s5 ξ 5 ∆ϕ1 dxdt + e−2sα s5 ξ 5 ∆ϕ1 dxdt. (3.35)
ε j=1
Q O1 ×(0,T )
The combination of (3.27) with (3.30), (3.31), (3.32), (3.33), (3.34), and (3.35) yields the existence of a constant
C such that for any s > C and for any ε > 0,
ZZ Xm ZZ
(m) 2 C (m−1) 2
e−2sα s5 ξ 5 ∆ϕ1 dxdt 6 ε I1 (s, ϕ(j) ) + e−2sα s9 ξ 9 ∆ϕ1 dxdt
ω
b ×(0,T ) j=1
ε
O1 ×(0,T )
m−2 Z Z
C X (j) 2
+ e−2sα s5 ξ 5 ∆ϕ1 dxdt. (3.36)
ε j=1
O1 ×(0,T )
(m−1)
We can repeat the same analysis for ∆ϕ1 and we obtain the existence of a constant C such that for any s > C
13
and for any ε > 0,
ZZ Xm ZZ
−2sα 9 9 (m−1) 2 C (m−2) 2
e s ξ ∆ϕ1 dxdt 6 ε I1 (s, ϕ(j) ) + e−2sα s17 ξ 17 ∆ϕ1 dxdt.
O1 ×(0,T ) j=1
ε
O2 ×(0,T )
m−3 Z Z
C X (j) 2
+ e−2sα s13 ξ 13 ∆ϕ1 dxdt. (3.37)
ε j=1
O2 ×(0,T )
Proceeding by induction, we can estimate all the local terms and we deduce from (3.25) that
m
X ZZ 2
(m+1)
+1 2(m+1) +1 (1)
I1 (s, ϕ(i) ) 6 C e−2sα s2 ξ ∆ϕ1 dxdt. (3.38)
i=1 Om ×(0,T )
(1)
Finally, we estimate the above local term in terms of ϕ1 . In order to do this, we consider ω
e an open subset
satisfying Om ⋐ ωe ⋐ ω and
ζe ∈ Cc2 (e
ω ) such that ζe ≡ 1 in Om , ζe > 0.
Then by integrating by parts, we obtain
ZZ 2 ZZ 2
(m+1)
+1 2(m+1) +1 (1) e −2sα s2(m+1) +1 ξ 2(m+1) +1 ∆ϕ(1) dxdt
e−2sα s2 ξ ∆ϕ1 dxdt 6 ζe 1
Om ×(0,T ) ω
e ×(0,T )
ZZ
=− e −2sα s2(m+1) +1 ξ 2(m+1) +1 ∆ϕ(1) · ∇ϕ(1) dxdt
∇ ζe 1 1
ω
e ×(0,T )
ZZ
− e −2sα s2(m+1) +1 ξ 2(m+1) +1 ∆∇ϕ(1) · ∇ϕ(1) dxdt.
ζe (3.39)
1 1
ω
e ×(0,T )
Then, we consider
ζ ∈ Cc2 (ω) such that ζ ≡ 1 in ω
e, ζ >0
and we integrate by parts:
ZZ 2 ZZ
(m+2)
−1 2(m+2) −1 (1) (m+2)
−1 2(m+2) −1 (1) 2
e−2sα s2 ξ ∇ϕ1 dxdt 6 ζe−2sα s2 ξ ∇ϕ1 dxdt
ω
e ×(0,T ) ω×(0,T )
ZZ
(m+2)
−1 2(m+2) −1 (1) (1)
=− ∇ ζe−2sα s2 ξ · ∇ϕ1 ϕ1 dxdt
ω×(0,T )
ZZ
(m+2)
−1 2(m+2) −1 (1) (1)
− ζe−2sα s2 ξ ∆ϕ1 ϕ1 dxdt.
ω×(0,T )
14
Using (2.7), we have
(m+2)
−1 2(m+2) −1 (m+2) (m+2)
∇(ζe−2sα s2 ξ ) 6 Ce−2sα s2 ξ2 .
Combining this inequality and Young’s inequality, we deduce the existence of C such that for any s > C and for
any ε > 0,
ZZ
(m+2)
−1 2(m+2) −1 (1) 2
e−2sα s2 ξ ∇ϕ1 dxdt
ω
e ×(0,T )
ZZ ZZ
(1) (1)
6 ε e−2sα s6 ξ 6 |∇ϕ1 |2 dxdt + e−2sα s5 ξ 5 |∆ϕ1 |2 dxdt
Q Q
ZZ
C (m+3)
−6 2(m+3) −6 (1) 2
+ e−2sα s2 ξ ϕ1 dxdt. (3.41)
ε
ω×(0,T )
Using Lemma 4 in [9] (see [23] for the proof) and the definition (3.1) of I1 , we deduce that
ZZ ZZ ZZ
(i) 2 (i) 2 (i) 2
I1 (s, ϕ(i) ) + e−2sα s8 ξ 8 ϕ1 dxdt > e−2sα s5 ξ 5 ∆ϕ1 dxdt + e−2sα s8 ξ 8 ϕ1 dxdt
ω×(0,T ) Q ω×(0,T )
ZZ 2 ZZ 2
(i) (i)
>C e−2sα s8 ξ 8 ϕ1 dxdt + e−2sα s6 ξ 6 ∇ϕ1 dxdt .
Q Q
This estimate, combined with (3.38), (3.40) and (3.41) yields the conclusion of Theorem 3.1.
Proof. First, we consider an energy estimate of the adjoint system (1.8). Multiplying each equation (1.9) by ϕ(i)
and integrating by parts, we deduce
Z X
m Z X
m Z X
m
1 d (i) 2 (i) 2
− ϕ dx + ∇ϕ dx = Aj,i ϕ(i) · ϕ(j) dx. (4.2)
2 dt
Ω i=1 Ω i=1 Ω i,j=1
Thus, using the Grönwall lemma, there exists C > 0 such that
Z X
m 2
(i)
t 7→ eCt ϕ (x, t) dx
Ω i=1
15
is nondecreasing. In particular, for some constant C > 0,
Z X 3T
Z /4Z X
m 2 m 2
(i) 2 CT (i)
ϕ (x, 0) dx 6 e ϕ (x, t) dx dt. (4.3)
i=1
T i=1
Ω T /4 Ω
Similarly, from (2.2), there exist two constants c1 , c2 > 0 such that for (x, t) ∈ Ω × [0, T ],
c1 c2
α(x, t) > , ξ(x, t) 6
tℓ (T − t)ℓ tℓ (T − t)ℓ
and consequently, for some constant c3 > 0,
2m+3 −6
−2sα 2m+3 −6 −2s
c1
tℓ (T −t)ℓ
c2
e (sξ) 6e s 6 c3 . (4.6)
t (T − t)ℓ
ℓ
Combining (4.3), (4.4), (4.5) and (4.6), we deduce that for some constant C,
Z Xm 2 ZZ
(i) C C T+ 1 (1) 2
ϕ (x, 0) dx 6 e ( T 2ℓ ) ϕ1 dxdt.
i=1
T
Ω ω×(0,T )
H := H m , V := V m . (4.7)
We define P0 : L2 (Ω)N → H the Leray projector and the projection P defined as:
m
P : L2 (Ω)N → H, y = y (1) , . . . , y (m) 7→ P0 y (1) , . . . , P0 y (m) .
16
Finally, we define the control operator B ∈ L(U, H) by
U := L2 (ω)N −1 , Bv = B v1 , . . . , vN −1 := P0 v1 , . . . , vN −1 , 0 1ω , 0, . . . , 0 . (4.9)
As it is well-known (see, for instance, [34, p.357]), system (4.10) is null-controllable in time T > 0 if and only if
there exists K(T ) > 0 such that
Z T
T A∗
2
∗ tA∗
2
e ϕ0
6 K(T )2
B e ϕ0
dt (ϕ0 ∈ H). (4.11)
H 0 H
with K : (0, ∞) → [0, ∞) continuous and non increasing. Let us consider T > 0 and suppose there exist ρ0 , ρ1 , ρ ∈
C 0 ([0, T ], R+ ), non increasing, positive in [0, T ) such that ρ0 (T ) = ρ1 (T ) = ρ(T ) = 0 and such that, for some
constant q > 1,
1
ρ0 (t) := ρ1 (q 2 (t − T ) + T )K((q − 1)(T − t)) t ∈ T 1 − 2 ,T , (4.13)
q
ρ0 6 Cρ, ρ1 6 Cρ, |ρ′ |ρ0 6 Cρ2 (t ∈ [0, T ]) . (4.14)
for some constant C > 0. We denote by L2ρ1 (0, T ; H) the space
f
L2ρ1 (0, T ; H) := f ∈ L2 (0, T ; H) ; ∈ L2 (0, T ; H)
ρ1
17
Theorem 4.2. With the above assumptions, there exists a bounded operator
1/2
ET ∈ L D (−A) × L2ρ1 (0, T ; H), L2ρ0 (0, T ; U)
such that for any y0 ∈ D (−A)1/2 and for any f ∈ L2ρ1 (0, T ; H), the solution y of (4.15) with u = ET (y0 , f )
satisfies
y 1/2
∈ L2 (0, T ; D(A)) ∩ C 0 [0, T ]; D (−A) ∩ H 1 (0, T ; H).
ρ
Moreover there exists a constant C such that
y
6 C ky k 1/2 + kf kL2 (0,T ;H) .
ρ
2 0 D ((−A) ) ρ1
L (0,T ;D(A))∩C 0 ([0,T ];D ((−A)1/2 ))∩H 1 (0,T ;H)
Remark 4.3. Note that in [28], A is assumed to be self-adjoint negative but the result can be extended to the
case where A is the generator of an analytic semigroup. Indeed, the hypothesis used in the proof is the maximal
regularity of (4.15) for v = 0.
Remark 4.4. Since ρ(T ) = 0, the above result implies in particular that y(T ) = 0, that is the null-controllability
of (4.15).
In the previous section, we have defined for our problem the spaces H, U, A and B, see (4.7), (4.8) and (4.9).
To show that A is the generator of an analytic semigroup, we first note that P∆ is self-adjoint negative (see, for
instance, [31, Theorem 2.1.1, p.128]) and thus is the generator of an analytic semigroup. Then using a perturbation
argument (see, for instance, [30, Theorem 2.1, p.80]), we deduce that A is the generator of an analytic semigroup
in H.
Finally, applying Lemma 4.1, we deduce that (4.12) holds with
CK
K(T ) = CK e T 2ℓ ,
Then we can check that (4.13) and (4.14) hold and we have moreover that
ρ2 6 ρ1 . (4.16)
18
More precisely, there exists
m
ET ∈ L V × L2ρ1 (0, T ; L2 (Ω)N ), L2ρ0 (0, T ; L2 (ω))
m
such that for any y0 ∈ V and for any f = f (1) , . . . , f (m) ∈ L2ρ1 (0, T ; L2 (Ω)N ), the solution y of (4.17) with the
control v = ET (y0 , f ) satisfies
y m m
∈ L2 (0, T ; H 2 (Ω)N ) ∩ C 0 [0, T ]; H 1 (Ω)N ∩ H 1 (0, T ; H). (4.18)
ρ
Moreover we have the following estimate
y
C ky k + kf k
ρ
2 6 0 V Lρ1 (0,T ;[L (Ω) ] ) .
2 2 N m (4.19)
L (0,T ;[H 2 (Ω)N ]m )∩C 0 ([0,T ];[H 1 (Ω)N ]m )∩H 1 (0,T ;H)
19
Using this and (4.20), we obtain
N (fe) N (fb)
ye y
y yb
T T
−
6
·∇ + ·∇
ρ1 ρ1
ρ ρ
L2 (0,T ;[L2 (Ω)N ]m )
ρ ρ
L2 (0,T ;[L2 (Ω)N ]m )
L2 (0,T ;[L2 (Ω)N ]m )
6 CR kf kL2 2 (Ω)N ]m ) .
ρ1 (0,T ;[L
Thus for R small enough, (NT )|BR is a strict contraction and this ends the proof of Theorem 1.3
References
[1] F. Ammar-Khodja, A. Benabdallah, M. González-Burgos, and L. de Teresa. Recent results on the controllability
of linear coupled parabolic problems: a survey. Math. Control Relat. Fields, 1(3):267–306, 2011.
[2] F. Ammar Khodja, A. Benabdallah, M. González-Burgos, and L. de Teresa. Minimal time for the null con-
trollability of parabolic systems: the effect of the condensation index of complex sequences. J. Funct. Anal.,
267(7):2077–2151, 2014.
[3] F. Ammar Khodja, A. Benabdallah, M. González-Burgos, and L. de Teresa. New phenomena for the null con-
trollability of parabolic systems: minimal time and geometrical dependence. J. Math. Anal. Appl., 444(2):1071–
1113, 2016.
[4] A. Benabdallah, F. Boyer, and M. Morancey. A block moment method to handle spectral condensation
phenomenon in parabolic control problems. Ann. H. Lebesgue, 3:717–793, 2020.
[5] N. Carreño, S. Guerrero, and M. Gueye. Insensitizing controls with two vanishing components for the three-
dimensional Boussinesq system. ESAIM Control Optim. Calc. Var., 21(1):73–100, 2015.
[6] N. Carreño and M. Gueye. Insensitizing controls with one vanishing component for the Navier-Stokes system.
J. Math. Pures Appl. (9), 101(1):27–53, 2014.
[7] F. Conforto, L. Desvillettes, and R. Monaco. Some asymptotic limits of reaction-diffusion systems appearing
in reversible chemistry. Ric. Mat., 66(1):99–111, 2017.
[8] J.-M. Coron and A. V. Fursikov. Global exact controllability of the 2D Navier-Stokes equations on a manifold
without boundary. Russian J. Math. Phys., 4(4):429–448, 1996.
[9] J.-M. Coron and S. Guerrero. Null controllability of the N -dimensional Stokes system with N − 1 scalar
controls. J. Differential Equations, 246(7):2908–2921, 2009.
[10] J.-M. Coron and P. Lissy. Local null controllability of the three-dimensional Navier-Stokes system with a
distributed control having two vanishing components. Invent. Math., 198(3):833–880, 2014.
[11] P. Érdi and J. Tóth. Mathematical models of chemical reactions. Nonlinear Science: Theory and Applications.
Princeton University Press, Princeton, NJ, 1989. Theory and applications of deterministic and stochastic
models.
[12] H. O. Fattorini and D. L. Russell. Uniform bounds on biorthogonal functions for real exponentials with an
application to the control theory of parabolic equations. Quart. Appl. Math., 32:45–69, 1974/75.
[13] E. Fernández-Cara, M. González-Burgos, S. Guerrero, and J.-P. Puel. Null controllability of the heat equation
with boundary Fourier conditions: the linear case. ESAIM Control Optim. Calc. Var., 12(3):442–465, 2006.
[14] E. Fernández-Cara, S. Guerrero, O. Y. Imanuvilov, and J.-P. Puel. Local exact controllability of the Navier-
Stokes system. J. Math. Pures Appl. (9), 83(12):1501–1542, 2004.
20
[15] E. Fernández-Cara, S. Guerrero, O. Y. Imanuvilov, and J.-P. Puel. Some controllability results for the N -
dimensional Navier-Stokes and Boussinesq systems with N − 1 scalar controls. SIAM J. Control Optim.,
45(1):146–173, 2006.
[16] A. V. Fursikov and O. Y. Imanuvilov. Controllability of evolution equations, volume 34 of Lecture Notes Series.
Seoul National University, Research Institute of Mathematics, Global Analysis Research Center, Seoul, 1996.
[17] M. González-Burgos and L. de Teresa. Controllability results for cascade systems of m coupled parabolic PDEs
by one control force. Port. Math., 67(1):91–113, 2010.
[18] S. Guerrero. Controllability of systems of Stokes equations with one control force: existence of insensitizing
controls. Ann. Inst. H. Poincaré Anal. Non Linéaire, 24(6):1029–1054, 2007.
[19] S. Guerrero. Null controllability of some systems of two parabolic equations with one control force. SIAM J.
Control Optim., 46(2):379–394, 2007.
[20] M. Iida, H. Monobe, H. Murakawa, and H. Ninomiya. Vanishing, moving and immovable interfaces in fast
reaction limits. J. Differential Equations, 263(5):2715–2735, 2017.
[21] O. Y. Imanuvilov. On exact controllability for the Navier-Stokes equations. ESAIM Control Optim. Calc. Var.,
3:97–131, 1998.
[22] O. Y. Imanuvilov. Remarks on exact controllability for the Navier-Stokes equations. ESAIM Control Optim.
Calc. Var., 6:39–72, 2001.
[23] O. Y. Imanuvilov and J.-P. Puel. Global Carleman estimates for weak solutions of elliptic nonhomogeneous
Dirichlet problems. Int. Math. Res. Not., (16):883–913, 2003.
[24] O. Y. Imanuvilov, J.-P. Puel, and M. Yamamoto. Carleman estimates for parabolic equations with nonhomo-
geneous boundary conditions. Chin. Ann. Math. Ser. B, 30(4):333–378, 2009.
[25] G. Lebeau and L. Robbiano. Contrôle exacte de l’équation de la chaleur. In Séminaire sur les Équations aux
Dérivées Partielles, 1994–1995, pages Exp. No. VII, 13. École Polytech., Palaiseau, 1995.
[26] J.-L. Lions and E. Magenes. Non-homogeneous boundary value problems and applications. Vol. I. Die
Grundlehren der mathematischen Wissenschaften, Band 181. Springer-Verlag, New York-Heidelberg, 1972.
Translated from the French by P. Kenneth.
[27] J.-L. Lions and E. Magenes. Non-homogeneous boundary value problems and applications. Vol. II. Die
Grundlehren der mathematischen Wissenschaften, Band 182. Springer-Verlag, New York-Heidelberg, 1972.
Translated from the French by P. Kenneth.
[28] Y. Liu, T. Takahashi, and M. Tucsnak. Single input controllability of a simplified fluid-structure interaction
model. ESAIM Control Optim. Calc. Var., 19(1):20–42, 2013.
[29] C. Montoya and L. de Teresa. Robust Stackelberg controllability for the Navier-Stokes equations. NoDEA
Nonlinear Differential Equations Appl., 25(5):Paper No. 46, 33, 2018.
[30] A. Pazy. Semigroups of linear operators and applications to partial differential equations, volume 44 of Applied
Mathematical Sciences. Springer-Verlag, New York, 1983.
[31] H. Sohr. The Navier-Stokes equations. Modern Birkhäuser Classics. Birkhäuser/Springer Basel AG, Basel,
2001. An elementary functional analytic approach, [2013 reprint of the 2001 original] [MR1928881].
[32] T. Takahashi. Boundary local null-controllability of the Kuramoto-Sivashinsky equation. Math. Control Signals
Systems, 29(1):Art. 2, 21, 2017.
21
[33] R. Temam. Behaviour at time t = 0 of the solutions of semilinear evolution equations. J. Differential Equations,
43(1):73–92, 1982.
[34] M. Tucsnak and G. Weiss. Observation and control for operator semigroups. Birkhäuser Advanced Texts:
Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks]. Birkhäuser Verlag, Basel, 2009.
22