ArticlePDF Available

Rapid Stabilization of a Linearized Bilinear 1D1-D Schr\"odinger Equation

Authors:

Abstract

We consider the one dimensional Schr\"odinger equation with a bilinear control and prove the rapid stabilization of the linearized equation around the ground state. The feedback law ensuring the rapid stabilization is obtained using a transformation mapping the solution to the linearized equation on the solution to an exponentially stable target linear equation. A suitable condition is imposed on the transformation in order to cancel the non-local terms arising in the kernel system. This conditions also insures the uniqueness of the transformation. The continuity and invertibility of the transformation follows from exact controllability of the linearized system.
arXiv:1611.03738v1 [math.OC] 11 Nov 2016
Rapid Stabilization of a Linearized Bilinear 1-D Schr¨
odinger
Equation
Jean-Michel CORON, Ludovick GAGNON, Morgan MORANCEY
Abstract
We consider the one dimensional Schr¨odinger equation with a bilinear control and prove the rapid
stabilization of the linearized equation around the ground state. The feedback law ensuring the rapid
stabilization is obtained using a transformation mapping the solution to the linearized equation on the
solution to an exponentially stable target linear equation. A suitable condition is imposed on the trans-
formation in order to cancel the non-local terms arising in the kernel system. This conditions also insures
the uniqueness of the transformation. The continuity and invertibility of the transformation follows from
exact controllability of the linearized system.
Keywords: Schr¨odinger equation, Rapid stabilization, Integral transform.
2010 Mathematics Subject Classification: 93D15, 93B52.
1 Introduction
1.1 Main result
Let T > 0. Consider the Schr¨odinger equation
(i∂tψ=ψu(t)µ(x)ψ, (t, x)(0, T )×(0,1),
ψ(t, 0) = ψ(t, 1) = 0, t (0, T ).(1.1)
In (1.1), ψis the complex-valued wave function, of L2-norm 1, of a particle confined in a 1Din-
finite square potential well. The particle is subjected to an electric field inside of the domain, where
uL2((0, T ); R)is the amplitude of the electric field and µH3((0,1); R)is the dipolar moment of the
particle.
Before stating our main result, we set some notations. Let A:D(A)L2((0,1); C)L2((0,1); C)be
defined by
Aφ:= φ, D(A) := H2((0,1); C)H1
0((0,1); C).(1.2)
The eigenvalues and eigenfunctions of Aare given by
λk:= ()2, ϕk(x) := 2 sin(kπx), k N.
The eigenstates of (1.1) (u= 0) are given by Φk(t, x) := ektϕk(x). The eigenstate Φ1associated to
the smallest eigenvalue is called the ground state.
Define the space Hs
(0)((0,1); C) := D(As/2)equipped with the inner product
hφ, ψiHs
(0) :=
+
X
k=1
λs
khφ, ϕkihψ, ϕki,
where ,·iis the L2((0,1); C)-inner product. The space Hs
(0) is endowed with the k.kHs
(0) -norm associated
to the Hs
(0)-inner product. We underline that the spaces used in this article can also be explicitly described
Universit´e Pierre et Marie Curie, LJLL UMR 7598, 4 place Jussieu, 75005, Paris, France.
Universit´e de Nice Sophia-Antipolis, CNRS UMR 7351, Laboratoire J.-A. Dieudonn´e, Parc Valrose, 06108, Nice, France.
Aix Marseille Universit´e, CNRS, Centrale Marseille, I2M UMR 7373 13453, Marseille, France.
1
by H2
(0)((0,1); C) = H2H1
0((0,1); C)and
H3
(0)((0,1); C) = φH3H1
0((0,1); C) ; φ′′ (0) = φ′′ (1) = 0,
H5
(0)((0,1); C) = nφH5H3
(0)((0,1); C) ; φ(4)(0) = φ(4)(1) = 0o.
We denote by Sthe radius 1sphere of L2((0,1); C).
Throughout this article, we assume that µsatisfies the following assumption.
Hypothesis 1.1 The function µbelongs to H3((0,1); R)and there exists c > 0satisfying
|hµϕ1, ϕki| c
k3,kN.(1.3)
Remark 1.2 A direct computation shows that, for every function µH3((0,1); R), we have
hµϕ1, ϕki=4
()3(1)k+1 µ(1) µ(0)2
()3Z1
0
(µϕ1)′′′(x) cos(kπx)dx (1.4)
and therefore there exists C > 0such that
|hµϕ1, ϕki| C
k3,kN.(1.5)
Moreover, since
lim
k+Z1
0
(µϕ1)′′′(x) cos(kπx)dx = 0,(1.6)
it follows from (1.4) that (1.3) implies that
µ(1) 6=µ(0) and µ(1) 6=µ(0).(1.7)
As proved in [8], Hypothesis 1.1 is not necessary to get local exact controllability of (1.1). However
(see [5]) it is a necessary and sufficient condition to get exact controllability of the following linearized
equation around the ground state
i∂tΨ = ∆Ψ v(t)µ(x1(t, x),(t, x)(0, T )×(0,1),
Ψ(t, 0) = Ψ(t, 1) = 0, t (0, T ),
Ψ(0, x) = Ψ0(x), x (0,1).
(1.8)
Theorem 1.3 ( [5] ) Let T > 0and assume that µsatisfies Hypothesis 1.1. Then, for every
Ψ0nφH3
(0) ;ℜhφ, ϕ1i= 0o=: H0,ΨTnφH3
(0) ;ℜhφ, Φ1(T)i= 0o=: HT,(1.9)
there exists vL2((0, T ); R)such that the solution Ψof (1.8) with the initial condition Ψ(0, .) = Ψ0
satisfies, Ψ(T , .) = ΨT.
Condition (1.9) means that Ψ0and ΨTlie in the tangent vector space of Sin ϕ1and Φ1(T), respectively.
Due to the linearization of the preservation of the norm for the bilinear problem, the solution of (1.8)
satisfies ℜhΨ(t),Φ1(t)i= 0 for every t0.
The main result of this paper is the construction of feedback laws leading to rapid stabilization of the linear
control system (1.8).
Theorem 1.4 Let T > 0. Assume that µsatisfies Hypothesis 1.1. Then, for every λ > 0, there exists
C > 0and a feedback law v(t) = K(Ψ(t, ·)) such that for every Ψ0 H0the associated solution of (1.8)
satisfies
kΨ(t, ·)kH3
(0) Ceλt kΨ0kH3
(0) .
2
For the sake of simplicity we will focus, for the rest of this article, on the rapid stabilization of the
following linear Schr¨odinger equation
i∂tΨ = ∆Ψ v(t)µ(x)ϕ1(x),(t, x)(0, T )×(0,1),
Ψ(t, 0) = Ψ(t, 1) = 0, t (0, T ),
Ψ(0, x) = Ψ0(x), x (0,1).
(1.10)
The only difference between (1.8) and (1.10) is that the control term v(t)µ(x1(t, x) has been replaced
by v(t)µ(x)ϕ1”. Using again [5, Proposition 4], we get the analogous of Theorem 1.3 that is exact
controllability with L2((0, T ); R)controls of system (1.10) but now in the state space H3
(0)((0,1); C).
We prove the following rapid stabilization result.
Theorem 1.5 Let T > 0. Assume that µsatisfies Hypothesis 1.1. Then, for every λ > 0, there exists
C > 0and a real-valued feedback law v(t) = K(ψ(t, ·)) such that, for every Ψ0H3
(0)((0,1); C), the
associated solution of (1.10) satisfies
kΨ(t, ·)kH3
(0) Ceλt kΨ0kH3
(0) .
Remark 1.6 The final goal would be to achieve local rapid stabilization of the bilinear problem (1.1)
toward the ground state Φ1. To avoid dealing with a moving target, notice that
kψ(t, ·)Φ1(t, ·)kH3
(0) =ke1tψ(t, ·)ϕ1kH3
(0) .
Thus it is simpler to look at the system satisfied by e1tψ(t, ·). In the same spirit, we will develop the proof
of Theorem 1.5 in this article and detail in Appendix C how we can modify the proof to obtain Theorem 1.4.
The obtained feedback law does not allow us, for now, to obtain rapid stabilization of (1.1).
1.2 A finite dimensional example
Let us explain the general idea of the proof of Theorem 1.5 in a finite dimensional setting. Let ARn×n
and BRn. Consider the control system
x(t) = Ax(t) + Bu(t)(1.11)
where, at time t, the state is x(t)Rnand the control is u(t)R. We assume that the system is
controllable, which is equivalent to the Kalman rank condition
rank(B , AB, . . . , An1B) = n(1.12)
(see e.g. [18, Theorem 1.16]). It is well-known that the controllability allows one to use the pole-shifting
theorem [18, Theorem 10.1]) to design a feedback law u(t) = Kx(t)to obtain the exponential stability
with an arbitrary exponential decay rate of (1.11). Let us present a different approach, more suitable for
PDEs, to this result. Let λRand denote the identity matrix of size nby I. Consider the target system
y(t) = (AλI)y(t) + Bv(t)(1.13)
where, at time t,y(t)Rnis the state and v(t)Ris the control. A straightforward computation shows
that, for v0, the solutions to (1.13) satisfy
ky(t)k e(λ−kAk)tky(0)k.
Let us assume, for the moment, that we can design a transformation (T, K )Rn×n×R1×nsuch that
if x(t)is the solution of (1.11) with
u(t) := Kx(t) + v(t),(1.14)
then y(t) := T x(t)is the solution of (1.13). Notice that if moreover Tis invertible, then
kx(t)k=kT1y(t)k kT1ke(λ−kAk)tky(0)k
kT1ke(λ−kAk)tkT x(0)k kT1kkTke(λ−kAk)tkx(0)k.
3
Therefore, the exponential stability of (1.11) with an arbitrary exponential decay rate is reduced to find
such a transformation (T, K)with Tinvertible. The transformation (T , K)maps (1.11) into
y(t) = T x(t) = T(Ax(t) + BK x(t) + Bv(t)) = (T A +T B K)x(t) + T Bv (t),
Hence this transformation maps (1.11) into (1.13) if and only if
T A +BK =AT λT, (1.15)
T B =B. (1.16)
One has the following theorem.
Theorem 1.7 ([19]) There exists one and only one (T , K)GLn(R)×R1×nsatisfying (1.15)-(1.16).
The proof of Theorem 1.7 provided in [19] relies on the phase variable canonical form (also called
controller form) of (1.11). We present here a different proof (in the case where the eigenvalues of Aare
simple) more suitable to deal with the infinite dimensional setting, with the additional assumption
λ > 0is such that ((λi+λ)IA)is invertible 1in. (1.17)
Proof: We first prove that the result holds for (T , K)GLn(C)×C1×n. The fact that (T , K)are
real-valued follows from the uniqueness of the transformations and the fact that Aand Bare real-valued.
Denote by {λi, ei}1inthe eigenvalues and eigenvectors of A. Then, (1.15)-(1.16) become
((λi+λ)IA)T ei=BKei,(1.18)
T Bei=Bei.(1.19)
The proof is then divided in four steps.
Step 1: Existence of a basis of the state space.
Assumption (1.17) implies that that there exists nvectors fi,1insatisfying
((λi+λ)IA)fi=B, 1in. (1.20)
We begin by proving that the set {fi}forms a basis of Cn. Notice that if Kis known, then one recovers
T eifrom the relation T ei=fiK ei.
Suppose there exists {ai}1inC,{ai}1in6= 0 such that
n
X
i=1
aifi= 0.(1.21)
Applying Ato this equation and using (1.20), we obtain
n
X
i=1
ai(λi+λ)fi= n
X
i=1
ai!B.
Applying successively A, we end up with
n
X
i=1
ai(λi+λ)pfi=
p
X
k=1 n
X
i=1
ai(λi+λ)pk!Ak1B, pN.(1.22)
Note that, for all jN, each coefficient
n
X
i=1
ai(λi+λ)j,(1.23)
appears in (1.22) for all pj+ 1 in front of Apj1B. We distinguish two cases. If there exists jN
such that there is a coefficient (1.23) that is not equal to zero, then it implies that {Apj1B}pj+1
4
span{fi}. From the controllability assumption it comes that span{Apj1B}pj+1 =Cn. Therefore, in
this case, the set of nvectors {fi}generates the whole space and consequently a basis of Cn.
The remaining case is the situation where every coefficient (1.23) vanishes i.e.
n
X
i=1
ai(λi+λ)j= 0,jN.(1.24)
In this case, consider the entire function
G:zC7→
n
X
i=1
aie(λi+λ)z.
From (1.24), we obtain for all jN,
G(j)(0) =
n
X
i=1
ai(λi+λ)j= 0.
Therefore G0. Let C:= Conv{λi+λ; 1 insuch that ai6= 0}where, for a nonempty subset R
of C, ConvRis the closed convex-hull of R. The set Chas at least one nonzero extremal, that is a point of
Csuch that there exists at least one hyperplane that meets Conly on this point. One such point must be of
the form λk0+λfor 1k0n. Therefore, there exists θ[0,2π]such that
(e(λk+λ)) <(e (λk0+λ)),1kn, k 6=k0.(1.25)
Let z=se where sR. We have
es(λk0+λ)e G(se) = ak0+
n
X
i=1,i6=k0
aies(λiλk0)e 0.(1.26)
From (1.25) and by letting s in (1.26), we obtain that ak0= 0. It is in contradiction with the fact
that the set Ccontains only nonzero ai. Therefore ai= 0,1inso the set {fi}is independant and
consequently a basis of Cn. The two cases were covered which implies that {fi}is a basis of Cn.
Remark 1.8 The latter part of the proof of the existence of a basis could have been done using the Vander-
monde matrix. The proof presented here has the advantage that it may be applied in the infinite dimensional
setting.
Step 2: Existence of the transformation (T, K ).
The transformation is obtained using (1.19). Indeed, let
B=
n
X
i=1
biei.
Notice that, by the controllability assumption, bi6= 0,1in. Then,
T B =B B=
n
X
i=1
biT ei=
n
X
i=1
biKeifi.(1.27)
Since {fi}1inis a basis of Cn, there exists {K ei}1inCsuch that the last equation is verified,
allowing to define TCn,n and KC1,n such that (1.18) and (1.19) hold.
Step 3: Uniqueness of (T , K).
To prove the uniqueness of the transformations (T , K), consider (T1, K1)and (T2, K2)solutions of
(1.18)-(1.19). Therefore (T1T2, K1K2)satisfies (1.18) and
(T1T2)B= 0.(1.28)
5
Since (T1T2, K1K2)satisfies (1.18), we use the basis constructed previously and (1.28) to prove that
K1=K2and T1=T2. With the uniqueness of the transformation and the fact that Aand Bare real-
valued, one ensures that the transformations are real-valued since (T , K )is also a solution of (1.18)-(1.19).
Step 4: Invertibility of T.
Let TCn,n and KC1,n be such that (1.18) and (1.19) hold. We prove that Tis invertible by
showing that Ker T={0}. Let xKer T. From (1.15)-(1.16), we obtain
TAx= (AT+KBT+λT )x= 0.
Therefore Ker Tis stable by A. From (1.16) it comes that
Bx=BTx= 0.
Thus there exists ˜xeigenvector of Ain Ker TKer B. From the controllability assumption and the
Hautus test (see for instance [48, Prop. 1.5.5]) it comes that Ker T={0}.
If the functional setting in the infinite dimensional case makes the proof more tricky, the strategy we
use remains the same. Riesz basis results will be used to prove the existence of a basis of the state space and
the invertibility of the transformation will be proved using the approximate controllability of the studied
system.
The main technical difficulty of this paper lies in the decomposition of B(1.27) in the basis of the state
space. Indeed, the control operator Bis admissible but not bounded in the state space. A careful analysis of
the Fourier components of the control operator Ballow us to define a transformation Twhich is bounded
from the state space into itself but the feedback transformation won’t be bounded from the state space into
R. Even so, the transformation Twill be proved to be invertible and the closed-loop linear equation will
be proved to be well-posed in the state space. It is important to note that this technical difficulty is in fact
essential for the invertibility of T(see Remark 3.6). Indeed, in our case, if Bwere to be bounded, then the
transformation Twould be compact and thus not continuously invertible. However, the unboundedness of
Kfrom the state space into Rprevents us to prove directly the well-posedness of the closed-loop nonlinear
equation.
Let us underline that the uniqueness condition T B =B, which was used implicitly in similar previous
works, will be crucial not only to obtain the existence and uniqueness of the transformation, but also to
define rigorously (1.15) in the infinite dimensional setting.
1.3 The linear Schr¨
odinger equation
As presented in the previous paragraph, the strategy to prove the rapid stabilization of the linear equation
(1.10) is inspired by the backstepping method. Recast the equations for Ψ1+iΨ2= Ψ as
t Ψ1
Ψ2!= 0
0 ! Ψ1
Ψ2!+v(t) 0
(µϕ1)(x)!,(t, x)(0, T )×(0,1),
Ψ1(t, 0) = Ψ1(t, 1) = 0,Ψ2(t, 0) = Ψ2(t, 1) = 0, t (0, T ),
Ψ1(0, x) = Ψ1
0(x),Ψ2(0, x) = Ψ2
0(x), x (0,1).
(1.29)
where Ψ1
0and Ψ2
0are the real and imaginary part of Ψ0respectively. From now on, all the functional
spaces are real-valued, except when specified. Moreover to deal with those real and imaginary parts, we
denote, for simplicity,
Xs
(0)((0,1); R) := Hs
(0)((0,1); C)L2((0,1); R)2,(1.30)
with the product topology. We will use the following operators
A:D(A) X3
(0) B:R C×(H3H1
0)((0,1); R)T
Ψ1
Ψ27− ∆Ψ2
∆Ψ1, a 7− a0
(µϕ1)(x),
6
with D(A) := X5
(0). Based on the previous work of the first author and Q. L ¨u ([22, 23]), instead of Volterra
transformations of the second kind usually used for the backstepping approach, we seek for transformations
(T, K)of the form
T:X3
(0) X3
(0)
Ψ1
Ψ27− Z1
0k11(x, y)k12(x, y)
k21(x, y)k22(x, y)Ψ1(y)
Ψ2(y)dy, (1.31)
K:D(K)X3
(0) R
Ψ1
Ψ27− Z1
0
α1(y1(y) + α2(y2(y)dy, (1.32)
such that if 1,Ψ2)Tis solution of (1.29) with
v(t) := KΨ1(t, .)
Ψ2(t, .),(1.33)
then (ξ1(t, .), ξ2(t, .))T:= T1(t, .),Ψ2(t, .))Tis solution to
t ξ1
ξ2!= 0
0 ! ξ1
ξ2!λ ξ1
ξ2!,(t, x)(0, T )×(0,1),
ξ1(t, 0) = ξ1(t, 1) = 0, ξ2(t, 0) = ξ2(t, 1) = 0, t ×(0, T ),
ξ1(0, x) = ξ1
0(x), ξ2(0, x) = ξ2
0(x), x ×(0,1),
(1.34)
with (ξ1
0, ξ2
0)T=T1
0,Ψ2
0)Tand Tis invertible in the state space. The decomposition in real and imagi-
nary part of the solution of (1.10) is made in order to ensure that the feedback v(t) = K1(t, .),Ψ2(t, .))T
is real-valued. Note that
ξ1(t)
ξ2(t)
X3
(0) eλt
ξ1
0
ξ2
0
X3
(0)
, t [0,+).(1.35)
The kernels are defined through the equations they must satisfy for (T , K)to map solutions of (1.29) to
solutions of (1.34). This is done formally in Section 2 together with a more detailed presentation of this
strategy.
1.4 A brief review of previous results
The controllability properties for the Schr ¨odinger equation were mostly studied in the usual (in opposition
to the bilinear model presented here) linear setting. For the control of the linear Schr¨odinger equation with
internal control (localized on a subdomain), we refer to the survey [33] and the references therein. In this
more classical setting we also mention [35, 32, 26] concerning stabilization.
Exact controllability of the bilinear Schr¨
odinger equation.
The first local controllability results on the bilinear Schr¨odinger equation appear in [2, 3, 5]. These local
controllability results have been extended with weaker assumptions in [8], in a more general setting in
infinite time [42] and also in the case of simultaneous controllability of a finite number of particles in
[38, 39]. Note that, despite the infinite speed of propagation, it was proved in [17, 4, 8, 38] that a minimal
amount of time is required for the controllability of some bilinear Schr¨odinger equations. More recently,
local exact controllability has been established in [6] for a Schr ¨odinger-Poisson model in 2D (see also [36]
for approximate controllability) and for the analogue of (1.1) with a control depending on time and space
in dimension less or equal than 3[45].
Approximate controllability and stabilization of the bilinear Schr¨
odinger equation.
The above mentioned results of exact controllability are mostly limited to the one dimensional case. In a
more general setting, the available results deal with approximate controllability. Using geometric control
7
techniques on appropriate Galerkin approximations, approximate controllability in different settings has
been proved [13, 10, 9, 16]. For a detailed presentation, see the survey [11].
However, most of these results are not suitable to prove approximate controllability in higher norms
(typically H3
(0)) and thus approximate controllability for bilinear Schr¨odinger equations has also been
studied from the Lyapunovfunctional point of view [37, 7, 40, 41]. Though it enabled global controllability
results, this strategy usually gives no indication on the convergencerate.
Rapid stabilization.
The strategy used in this article is inspired from backstepping techniques. Initially developed to design,
in a recursive way, more effective feedback laws for globally asymptotic stable finite dimensional system
for which a feedback law and a Lyapunov function are already known (we refer to [30, 46, 18] for a
comprehensive introduction in finite dimension and [20, 34] for an application of this method to partial
differential equations), the backstepping approach was later used in the infinite dimensional setting to
design feedback laws by mapping the system to stabilize to a target stable system. To our knowledge,
this strategy was first introduced in the context of partial differential equations to design a feedback law
for heat equation [1] and, later on, for a class of parabolic PDE [12]. The backstepping-like change of
coordinates of the semi-discretized equation maps the discrete solution to stabilize to a stable solution.
The corresponding continuous mapping obtained is a Volterra transformation of the second kind, seen
as an infinite dimensional backstepping transformation from the triangular domain of definition of the
transformation. This backstepping strategy in infinite dimensions led to a series of work (see [31] for
a global presentation and [29] for the use of the backstepping approach for the rapid stabilization of a
Schr¨odinger equation with a boundary control). The backstepping approach can be used to get stabilization
in finite time as shown in [24, 25]. Moreover it is not limited to linear equations, as shown in [14, 25].
Though using Volterra transformations of the second kind provides easily invertible transformations it is
also limited. Fredholm transformations, from which this paper is inspired, has already been used [22, 23,
21] for rapid stabilization using boundary feedback laws.
Abstract methods have been developed to obtain the rapid stabilization of linear partial differential
equations. Among them, we cite the works [28], [50] and [49], based on the Gramian approach and the
Riccati equations, which could be applied to obtain the rapid stabilization of the linearized equation (1.29)
as (0, µϕ1)D(A). However, it seems difficult to obtain, for various physical systems, the local
rapid stabilization of the nonlinear equation using those methods. For example, at least for the moment,
one does not know how to deduce from [15], where the rapid stabilization of a linearized Korteweg-de
Vries equation is obtained by using the method developed in [49], the rapid stabilization of the associated
nonlinear Korteweg-de Vries equation. This is in contrast with the method used here (linear transformations
to suitable target systems) applied to the same Korteweg-de Vries equation. Indeed, as shown in [22], the
feedback laws obtained by means of this method allows to get the rapid stabilization for the nonlinear
Korteweg-de Vries equation. One may therefore hope that, as in [22], our feedback law Kbeing quite
explicit it might allow to obtain the rapid stabilization of the nonlinear equation. Note however that in
[22] the feedback law Kis continuous, which is not the case in our situation. It makes the application to
the nonlinear system more complicated to study and requires suitable nonlinear modifications of the linear
feedback law K.
1.5 Structure of the article
In Sec. 2, we give a detailed presentation of the strategy used to construct the transformation (T, K )and
give a formal expression of this transformation. In Sec. 3 we prove that this formal transformation Tis well
defined and is continuous in the state space X3
(0). Then, we prove in Sec. 4 that the previous transformation
is indeed invertible in the state space. These properties of Twill follow using Hypothesis 1.1 i.e. exact
controllability of the linearized system. We end the proof of Theorem 1.5 in Sec. 5 by proving that the
constructed feedback Kleads to a well-posed closed-loop system (i.e. the equation (1.29) with vdefined
by (1.33)) and that Tactually maps the closed-loop system to the exponentially stable solutions of (1.34).
In Appendix A we study in a similar way a simplified Saint-Venant equation which exhibits the same
phenomenon but on which we explicitly compute the transformation (T, K ).
8
2 Heuristic construction of the transformations
We recall that we look for a transformation (T, K )of the form (1.31)- (1.32). Let us derive the set of
equations for (T , K)to map solutions of (1.29) to solutions of (1.34). First, to ensure that the boundary
conditions of (1.34) are satisfied, we assume that
kij (0, y) = kij (1, y) = 0,y(0,1),i, j {1,2}.(2.1)
Using the fact that 1,Ψ2)TX3
(0), and imposing the conditions
kij (x, 0) = kij (x, 1) = 0,x(0,1),i, j {1,2},(2.2)
formal computations, denoting xand ythe Dirichlet Laplacian with respect to xand yvariables re-
spectively, lead to
tξ1(t, x) + ξ2(t, x) + λξ1(t, x)
=Z1
0
k11(x, y)yΨ2(t, y)+k12 (x, y)yΨ1(t, y) + v(t)(µϕ1)(y)dy
+Z1
0
xk21(x, y1(t, y) + xk22 (x, y2(t, y )dy
+Z1
0
λk11(x, y1(t, y) + λk12 (x, y2(t, y )dy
=Z1
0
(∆xk21 + yk12 +λk11) (x, y1(t, y)dy
+Z1
0
(∆xk22 yk11 +λk12) (x, y2(t, y)dy
+v(t)Z1
0
(µϕ1) (z)k12(x, z)dz.
The boundary conditions (2.2) were imposed to avoid boundary terms in the integrations by parts. Using
the expression (1.32) of the feedback leads to
tξ1(t, x) + ξ2(t, x) + λξ1(t, x)
=Z1
0(∆yk12 + xk21 +λk11) (x, y) + α1(y)Z1
0
(µϕ1) (z)k12(x, z)dzΨ1(t, y)dy
Z1
0(∆yk11 xk22 λk12) (x, y)α2(y)Z1
0
(µϕ1) (z)k12(x, z )dzΨ2(t, y)dy. (2.3)
In the same way one gets
tξ2(t, x)ξ1(t, x) + λξ2(t, x)
=Z1
0(∆yk22 xk11 +λk21) (x, y) + α1(y)Z1
0
(µϕ1) (z)k22(x, z)dzΨ1(t, y)dy
Z1
0(∆yk21 + xk12 λk22) (x, y)α2(y)Z1
0
(µϕ1) (z)k22(x, z )dzΨ2(t, y)dy. (2.4)
9
If we want (ξ1, ξ2)to satisfy (1.34) then we need to find the functions kij and αjsatisfying
(∆yk11 xk22 λk12) (x, y)α2(y)Z1
0
(µϕ1) (z)k12(x, z)dz= 0,(x, y )(0,1)2,
(∆yk12 + xk21 +λk11) (x, y) + α1(y)Z1
0
(µϕ1) (z)k12(x, z)dz= 0,(x, y )(0,1)2,
(∆yk21 + xk12 λk22) (x, y)α2(y)Z1
0
(µϕ1) (z)k22(x, z)dz= 0,(x, y )(0,1)2,
(∆yk22 xk11 +λk21) (x, y) + α1(y)Z1
0
(µϕ1) (z)k22(x, z)dz= 0,(x, y )(0,1)2,
kij (x, 0) = kij (x, 1) = 0, x (0,1),
kij (0, y) = kij (1, y) = 0, y (0,1).
(2.5)
A fundamental extra condition. One could try to solve rightaway (2.5) and prove the invertibility of the
transformation Tbut the non-local terms yield a tedious task. To overcome this difficulty, one introduces,
as in the finite dimensional setting, what will be referred throughout this article as the T B =Bcondition,
T0
(µϕ1)(x)=Z1
0k12(x, y)(µϕ1)(y)
k22(x, y)(µϕ1)(y)dy =0
(µϕ1)(x).
Plugging this into (2.5) we obtain that we now seek for a solution to
(∆yk11 xk22 λk12) (x, y) = 0,(x, y)(0,1)2,
(∆yk12 + xk21 +λk11) (x, y) = 0,(x, y)(0,1)2,
(∆yk21 + xk12 λk22) (x, y)α2(y)(µϕ1)(x) = 0,(x, y)(0,1)2,
(∆yk22 xk11 +λk21) (x, y) + α1(y)(µϕ1)(x) = 0,(x, y)(0,1)2,
kij (x, 0) = kij (x, 1) = 0, x (0,1),
kij (0, y) = kij (1, y) = 0, y (0,1),
(2.6)
together to the T B =Bcondition
Z1
0
k12(x, y)(µϕ1)(y)dy = 0, x (0,1),
Z1
0
k22(x, y)(µϕ1)(y)dy = (µϕ1)(x), x (0,1).
(2.7)
Remark 2.1 In [22, 23], the authors were dealing with a boundary control for the Korteweg-de Vries
equation and for the Kuramoto-Sivashinsky equation. In their case, the T B =Bcondition writes
ky(x, L) = 0, x (0, L),
for the former and
kyy (x, L) = 0, x (0, L),
for the latter, where kis the kernel of the Fredholm transformation in each case. Contrary to our frame-
work, this boundary condition appeared naturally from the integration by parts performed in order to
obtain the equation on the kernel. It was not seen as a particular boundary condition although, a careful
analysis of their work shows that the T B =Bcondition was used to prove the uniqueness of the transfor-
mation (T , K). The common ground between their work and this article is the additional regularity that
the kernel needs in order to satisfy the T B =Bcondition. Notice that in what we present in this article,
the relation between the kernels kij and αjis more intricate and considerably modify the analysis.
10
Formal decomposition The global strategy to construct a solution of (2.6)-(2.7) is the following. First
assume that α1and α2are known. This enables us to compute the kernels kij satisfying (2.6) as functions
of α1and α2. Then we prove that we find α1and α2such that (2.7) is satisfied.
We decompose the functions in the following form
kij (x, y) =
+
X
n=1
fij
n(x)ϕn(y), αj(y) =
+
X
n=1
αj
nϕn(y).(2.8)
This leads to
f11
n′′(x) + λnf22
n(x)λf21
n(x) = α1
n(µϕ1)(x),
f12
n′′(x)λnf21
n(x)λf22
n(x) = α2
n(µϕ1)(x),
f21
n′′(x)λnf12
n(x) + λf11
n(x) = 0,
f22
n′′(x) + λnf11
n(x) + λf12
n(x) = 0,
fij
n(0) = fij
n(1) = 0.
(2.9)
The T B =Bcondition (2.7) becomes
+
X
n=1hµϕ1, ϕnif12
n(x) = 0,
+
X
n=1hµϕ1, ϕnif22
n(x) = (µϕ1)(x).
(2.10)
As mentioned, we begin by assuming that the feedback law is known. We considertwo sequences (β1
n)nN
and (β2
n)nNto be precised later on.
Let (g11
n, g12
n, g21
n, g22
n)Tbe the solution of system (2.9) with right-hand side β1
n(µϕ1),0,0,0Tand let
(h11
n, h12
n, h21
n, h22
n)Tbe the solution of system (2.9) with right-hand side 0, β2
n(µϕ1),0,0T. System (2.9)
being linear, it comes that
fij
n=α1
n
β1
n
gij
n+α2
n
β2
n
hij
n.(2.11)
Decomposing gij
nin the L2-orthonormal basis (ϕk)kN, if we denote by Ank the following matrix
Ank =
λk0λ λn
0λkλnλ
λλnλk0
λnλ0λk
,(2.12)
system (2.9) leads to
Ank
hg11
n, ϕki
hg12
n, ϕki
hg21
n, ϕki
hg22
n, ϕki
=
β1
nhµϕ1, ϕki
0
0
0
.
Then
g11
n(x) =
+
X
k=1
λkλ2
nλ2λ2
k
δnk(λ)β1
nhµϕ1, ϕkiϕk(x),
g12
n(x) =
+
X
k=1
2λλkλn
δnk(λ)β1
nhµϕ1, ϕkiϕk(x),
g21
n(x) =
+
X
k=1
λλ2+λ2
k+λ2
n
δnk(λ)β1
nhµϕ1, ϕkiϕk(x),
g22
n(x) =
+
X
k=1
λnλ2λ2
k+λ2
n
δnk(λ)β1
nhµϕ1, ϕkiϕk(x),
(2.13)
11
where
δnk(λ) = det(Ank) = λ2+ (λkλn)2λ2+ (λk+λn)2.(2.14)
The same computations can be carried out for hij
n, leading to the following relations
h11
n=β2
n
β1
n
g12
n, h12
n=β2
n
β1
n
g11
n,
h21
n=β2
n
β1
n
g22
n, h22
n=β2
n
β1
n
g21
n.
(2.15)
For the sake of simplicity, we denote by cij
nk and dij
nk the coefficients such that
gij
n(x) =
+
X
k=1
cij
nkβ1
nhµϕ1, ϕkiϕk(x), hij
n(x) =
+
X
k=1
dij
nkβ2
nhµϕ1, ϕkiϕk(x).(2.16)
Summary of the construction. Finally, using the definition of the transformation (1.31), the decomposi-
tions (2.8), (2.11) and the relations (2.15) it comes that the transformation Twe are looking for is defined
by
TΨ1
Ψ2=
+
X
n=1 α1
n
β1
n
g11
n+α2
n
β2
n
h11
nhΨ1, ϕni+
+
X
n=1 α1
n
β1
n
g12
n+α2
n
β2
n
h12
nhΨ2, ϕni
+
X
n=1 α1
n
β1
n
g21
n+α2
n
β2
n
h21
nhΨ1, ϕni+
+
X
n=1 α1
n
β1
n
g22
n+α2
n
β2
n
h22
nhΨ2, ϕni
=
+
X
n=1 α1
nhΨ2, ϕni
β1
nα2
nhΨ1, ϕni
β1
ng12
n
g22
n+α1
nhΨ1, ϕni
β2
n
+α2
nhΨ2, ϕni
β2
nh12
n
h22
n.
(2.17)
In the same spirit, the T B =Bcondition (2.10) becomes
0
µϕ1=
+
X
n=1
α1
nhµϕ1, ϕni
β1
ng12
n
g22
n+α2
nhµϕ1, ϕni
β2
nh12
n
h22
n.(2.18)
This ends the heuristic of the construction of the transformation and the feedback law. Indeed, in the
following section we prove that for a suitable choice of the sequences (β1
n)nNand (β2
n)nNthen
B:= g12
n
g22
n,h12
n
h22
n;nN,(2.19)
is a Riesz basis of X2
(0) (see Proposition 3.4). Then from (2.18) we get the feedback laws from the expan-
sion of (0, µϕ1)Tin the basis B. Finally, we study the behaviour of the coefficients α1
nand α2
nas ngoes
to infinity to prove that the transformation Tgiven by (2.17) is indeed continuous from X3
(0) to X3
(0).
Remark 2.2 From (2.18) it already appears that the behaviour of the coefficients hµϕ1, ϕni, and thus
Hypothesis 1.1, will play a crucial role.
3 Definition and properties of the transformation
In this section, we make rigorous the heuristic developed in the previous section. In subsection 3.1, we
prove that for a suitable choice of β1
nand β2
nthen Bdefined in (2.19) is a Riesz basis of X2
(0) where the
functions gij
nand hij
nare defined by (2.13) and (2.15). This enables us to define the feedback law and the
transformation Tusing the relations (2.18) and (2.17). However, this does not give enough regularity to
prove that T:X3
(0) X3
(0). We prove the extra regularity we need on the feedback laws in subsection 3.2.
This leads, in subsection 3.3, to the expected regularity for the transformation T.
12
3.1 Riesz basis property
Let us recall some results on Riesz basis.
Definition 3.1 Let Hbe an Hilbert space and {gn}nNH. We say that {gn}nNis ω-independent if
X
nN
angn= 0,with {an}nNR=an= 0,nN.
Theorem 3.2 [51, Theorem 15] Let Hbe a separable Hilbert space and let {en}nNbe an orthonormal
basis for H. If {gn}nNis an ω-independent sequence quadratically close to {en}nN, that is
X
nNkengnk2
H<+,
then {gn}nNis a Riesz basis for H.
Theorem 3.3 Let Hbe a separable Hilbert space and let {en}nNbe an orthonormal basis for H. If
{gn}nNis dense in Hand is quadratically close to {en}nN, i.e.
X
nNkengnk2
H<+,
then {gn}nNis a Riesz basis for H.
Let us provide a proof of Theorem 3.3, stated as a remark in [27, Remark 2.1, p. 318].
Proof:
Let us prove Theorem 3.3 by contradiction. Suppose that the {gn}nNis dense in Hand is quadrat-
ically close to {en}nNbut that {gn}nNis not a Riesz basis. Therefore, by Theorem 3.2, there must
exist a non-zero sequence {an}nNRsuch that
X
nN
angn= 0.
Since {gn}nNis quadratically close to {en}nN, there exists NNsuch that
X
nN+1 kengnk2
H<1.
Therefore, from [51, Theorem 13], the {en}1nN {gn}nN+1 is a Riesz basis of H. This implies that
there exists kNsuch that ak6= 0. Hence, with the density assumption, we have
H=span{gn|nN}= span{gn|nN\ {k}}.
Thus, codim(span{gn|nN+ 1})N1.
However
H/span{gn|nN+ 1},
is isomorphic to span{en|nN}, which is of dimension N, leading to a contradiction.
We will use the previous criteria to prove the following proposition.
Proposition 3.4 Let βi
nbeing defined as (3.1). Then,
B:= g12
n
g22
n,h12
n
h22
n;nN
is a Riesz basis of X2
(0) and
˜
B:= ( g12
n1/2
n
g22
n1/2
n!, h12
n1/2
n
h22
n1/2
n!;nN)
is a Riesz basis of X3
(0).
13
To apply the previous criterion for the Riesz basis, we prove that B(resp. ˜
B) is quadratically close to
the orthonormal basis of X2
(0) (resp. X3
(0)) given by
ϕns/2
n
0,0
ϕns/2
n;nN,with s= 2 (resp. s= 3).
Thus, we choose β1
nand β2
nsuch that g12
nand h22
nare close to ϕnni.e. hg12
n, ϕni=hh22
n, ϕni= 1n,
that is,
β1
n:= λ(λ2+ 4λ2
n)
2λ3
nhµϕ1, ϕni,
β2
n:= λ(λ2+ 4λ2
n)
(λ2+ 2λ2
n)λnhµϕ1, ϕni.
(3.1)
Notice that, from Hypothesis 1.1 and Remark 1.2, there exists C > 0such that
n
Cβ1
nC n, n
Cβ2
nC n, nN.(3.2)
During the proof of Proposition 3.4, we will use the following lemma. Its proof is purely technical and
postponed to Appendix B.
Lemma 3.5 With the above definition, for s= 2 and s= 3, one gets
X
nN
ϕns/2
n
0 g12
n(s2)/2
n
g22
n(s2)/2
n!
2
Xs
(0)
<+,
X
nN
0
ϕns/2
n h12
n(s2)/2
n
h22
n(s2)/2
n!
2
Xs
(0)
<+.
We are now ready to prove Proposition 3.4.
Proof: Let s= 2 or s= 3. In view of Theorem 3.2 and Lemma 3.5, assume that there exists a sequence
{an, bn}nNsuch that
X
nN
ang12
n
g22
n+bnh12
n
h22
n=0
0in Xs
(0).(3.3)
First step: we apply negative powers of the Laplacian to characterize elements of S:= span B.
Recall that Ais defined in (1.2). Let us write (2.9) for gij
nas
Agn=
λn
0 0 0 1
0 0 1 0
01 0 0
1 0 0 0
+λ
0 0 1 0
0 0 0 1
1 0 0 0
0 1 0 0
|{z }
=:Jn
gn
β1
nµϕ1
0
0
0
|{z }
=:Fn
,
where gn= (g11
n, g12
n, g21
n, g22
n)Tand where Agn= (Ag11
n,Ag12
n,Ag21
n,Ag22
n)T.
Since det(Jn) = (λ2+λ2
n)26= 0, we have
A1gn=J1
ngn+A1J1
nFn,
where
J1
n=1
λ2+λ2
n
0 0 λ λn
0 0 λnλ
λλn0 0
λnλ0 0
.
14
Therefore, using (2.15), we obtain
A1g12
n=1
λ2
n+λ2λg22
nλn
β1
n
β2
n
h22
n,
A1g22
n=1
λ2
n+λ2λn
β1
n
β2
n
h12
nλg12
n+λnβ1
nA1(µϕ1),
A1h12
n=1
λ2
n+λ2λn
β2
n
β1
n
g22
n+λh22
n,
A1h22
n=1
λ2
n+λ2λn
β2
n
β1
n
g12
n+λh12
n+λβ2
nA1(µϕ1).
(3.4)
Thus, applying A1to (3.3) leads to
c0A1(µϕ1)
0=X
nN
λan+λnβ2
n
β1
nbn
λ2
n+λ2g12
n
g22
n+λnβ1
n
β2
nan+λbn
λ2
n+λ2h12
n
h22
n,(3.5)
where
c0:= X
nN
λnβ1
nanλβ2
nbn
λ2
n+λ2.(3.6)
Applying A1to (3.5) we obtain
c1A1(µϕ1)
c0A2(µϕ1)=X
nN
(λ2λ2
n)an+ 2λλnβ2
n
β1
nbn
(λ2
n+λ2)2g12
n
g22
n+2λλnβ1
n
β2
nan+ (λ2λ2
n)bn
(λ2
n+λ2)2h12
n
h22
n,
(3.7)
where
c1:= X
nN
2λλnβ1
nan+ (λ2
nλ2)β2
nbn
(λ2
n+λ2)2.(3.8)
In order to iterate (3.5) and (3.7), notice that applying A1to the relations (3.4) leads to
A2g12
n=1
(λ2
n+λ2)2h(λ2
nλ2)g12
n+ 2λλn
β1
n
β2
n
h12
n+ 2λλnβ1
nA1(µϕ1)i,
A2g22
n=1
(λ2
n+λ2)2h(λ2
nλ2)g22
n+ 2λλn
β1
n
β2
n
h22
ni+λnβ1
n
λ2
n+λ2A2(µϕ1),
A2h12
n=1
(λ2
n+λ2)2h(λ2
nλ2)h12
n2λλn
β2
n
β1
n
g12
n+ (λ2
nλ2)β2
nA1(µϕ1)i,
A2h22
n=1
(λ2
n+λ2)2h(λ2
nλ2)h22
n2λλn
β2
n
β1
n
g22
niλβ2
n
λ2
n+λ2A2(µϕ1).
(3.9)
Applying successively A2to (3.5) we obtain, for every pN(with the convention that the sum from 1
to 0is equal to 0),
c0A(2p+1)(µϕ1) + Pp
j=1 PnN
2λλnβ1
nkj
n+(λ2
nλ2)β2
nlj
n
(λ2
n+λ2)2j+1 A2(pj)1(µϕ1)
Pp
j=1 PnN
λnβ1
nkj
nλβ2
nlj
n
(λ2
n+λ2)2jA2(pj+1)(µϕ1)
=X
nN
1
(λ2
n+λ2)2p+1 kp+1
ng12
n
g22
n+lp+1
nh12
n
h22
n,
(3.10)
where the coefficients (kj
n, lj
n)are defined by
kj+1
n
lj+1
n=
(λ2λ2
n) 2λλnβ2
n
β1
n
2λλnβ1
n
β2
n(λ2λ2
n)
kj
n
lj
n,with k1
n
l1
n:=
λan+λnβ2
n
β1
nbn
λnβ1
n
β2
nan+λbn
.(3.11)
15
Applying successively A2to (3.7) we obtain, for every pN,
c1A(2p+1)(µϕ1) + Pp
j=1 PnN
2λλnβ1
n˜
kj
n+(λ2
nλ2)β2
n˜
lj
n
(λ2
n+λ2)2(j+1) A2(pj)1(µϕ1)
c0A2(p+1) Pp
j=1 PnN
λnβ1
n˜
kj
nλβ2
n˜
lj
n
(λ2
n+λ2)2j+1 A2(pj+1)(µϕ1)
=X
nN
1
(λ2
n+λ2)2(p+1) ˜
kp+1
ng12
n
g22
n+˜
lp+1
nh12
n
h22
n,
(3.12)
where the coefficients (˜
kj
n,˜
lj
n)are defined by
˜
kj+1
n
˜
lj+1
n=
(λ2λ2
n) 2λλnβ2
n
β1
n
2λλnβ1
n
β2
n(λ2λ2
n)
˜
kj
n
˜
lj
n,with ˜
k0
n
˜
l0
n=an
bn.(3.13)
Assuming that c06= 0, equality (3.5) implies that
A1(µϕ1)
0span B.
Then using (3.7) we obtain
0
c0A2(µϕ1)+c1A1(µϕ1)
0span B=0
A2(µϕ1)span B.
Iterating with the relations (3.10) and (3.12) it comes that, for every pN,
A(2p1)(µϕ1)
0span Band 0
A2p(µϕ1)span B.(3.14)
Notice that if c0= 0 and c16= 0, the same argument can be repeated to obtain (3.14). Actually, one
gets (3.14) as soon as there exists a non-zero coefficient in the left-hand side of (3.10) or in the left-hand
side of (3.12). Thus it comes that either (3.14) holds or, for every jN,
0 = X
nN
2λλnβ1
nkj
n+ (λ2
nλ2)β2
nlj
n
(λ2
n+λ2)2j+1 ,
0 = X
nN
λnβ1
nkj
nλβ2
nlj
n
(λ2
n+λ2)2j,
0 = X
nN
2λλnβ1
n˜
kj
n+ (λ2
nλ2)β2
n˜
lj
n
(λ2
n+λ2)2(j+1) ,
0 = X
nN
λnβ1
n˜
kj
nλβ2
n˜
lj
n
(λ2
n+λ2)2j+1 .
(3.15)
Second step: we prove that if (3.14) holds, then Bis a Riesz basis.
Let (d1, d2)TXs
(0) such that
d1
d2,φ1
φ2= 0,φ1
φ2span B.
Using (3.14), we obtain that, for every pN,
0 = d1
d2,A(2p1)(µϕ1)
0Xs
(0)
=hd1,A(2p1)(µϕ1)iHs
(0)
=X
kN
λs
khd1, ϕkihϕk,A(2p1)(µϕ1)i
=X
kN
λs+1
khµϕ1, ϕkihd1, ϕki
λ2p
k
.
16
Let d1
k:= hd1, ϕkiand define
G:zC7→ X
kN
λs1
khµϕ1, ϕkid1
kez/λ2
kC.
From uniform convergence on compact sets, it comes that Gis an entire function and the previous relation
imply that, for every pN,
G(p1)(0) = X
kN
λs+1
khµϕ1, ϕkid1
k
λ2p
k
= 0.
Thus, G0. If d16= 0, let
n0:= min{nN;d1
n6= 0}.
It comes that
0 = λs1
n0hµϕ1, ϕn0id1
n0+X
k>n0
λs1
khµϕ1, ϕkid1
kexp z1
λ2
k1
λ2
n0.(3.16)
As hµϕ1, ϕn0i 6= 0, considering zreal and letting zgo to +in (3.16) imply d1
n0= 0. Therefore, from
the definition of n0,d1= 0.
Using the exact same strategy for
0
A2p(µϕ1)span B,pN,
implies d2= 0. Thus, it comes that d1=d2= 0 i.e. span B=Xs
(0). From Theorem 3.3, we obtain that
Bis a Riesz basis of X2
(0) (resp. ˜
Bis a Riesz basis of X3
(0)).
Third step: we prove that in the remaining case (3.15), one has an=bn= 0 for any nN.
Let us define
˜
G:zC7→ X
nN
1
(λ2
n+λ2)2 2λλnβ1
n
(λ2
nλ2)β2
n,exp (zM )an
bn,(3.17)
with
M:= 1
(λ2
n+λ2)2
(λ2λ2
n) 2λλnβ2
n
β1
n
2λλnβ1
n
β2
n(λ2λ2
n)
.(3.18)
Notice that the matrix appearing in this definition is the one used in the recurrence relations (3.11) and (3.13).
This matrix is diagonalized as follows
M=1
(λ2
n+λ2)2
iβ2
n
β1
niβ2
n
β1
n
1 1
(λ2λ2
n) + 2iλλn0
0 (λ2λ2
n)2iλλn
i
2
β1
n
β2
n
1
2
i
2
β1
n
β2
n
1
2
.
From this diagonalization we deduce that
˜
G(z) = X
nN
exp (λ+n)2
(λ2+λ2
n)2z"1
nan
2
(λn)2
(λ2
n+λ2)2β2
nbn
2
(λ+n)2
(λ2
n+λ2)2#
+ exp (λn)2
(λ2+λ2
n)2z"1
nan
2
(λ+n)2
(λ2
n+λ2)2+β2
nbn
2
(λn)2
(λ2
n+λ2)2#.(3.19)
Again, ˜
Gis an entire function from the uniform convergence on compact sets. The recurrence rela-
tion (3.13) implies that, for every pN,
˜
G(p)(0) = X
nN
1
(λ2
n+λ2)2 2λλnβ1
n
(λ2
nλ2)β2
n, M pan
bn
=X
nN
2λλnβ1
n˜
kp
n+ (λ2
nλ2)β2
n˜
lp
n
(λ2
n+λ2)2(p+1)
= 0.
17
Thus we obtain ˜
G0.(3.20)
We claim that (λn)2
(λ2
n+λ2)26=(λ+m)2
(λ2
m+λ2)2,(n, m)(N)2.(3.21)
Indeed, let (n, m)(N)2be such that
(λn)2
(λ2
n+λ2)26=(λ+m)2
(λ2
m+λ2)2.(3.22)
Taking the modulus of both sides of (3.22), one gets
1
λ2
n+λ2=1
λ2
m+λ2,
which implies that λn=λmand therefore n=m. Hence
(λn)2
(λ2
n+λ2)2=(λ+n)2
(λ2
n+λ2)2,(3.23)
Taking the imaginary part of (3.23) yields a contradiction and proves that (3.21) holds.
For simplicity, we rewrite Gas
G(z) = X
nN
Cneµnz,
with µnequal to (λk)2/(λ2
k+λ2)2or (λ+k)2/(λ2
k+λ2)2for some kNand Cnis the
corresponding coefficient in (3.19). Notice that µnare all different, µn0as n and Cn
2(N;C).
We repeat the same argument as in the finite dimensional case. Let
C:= Conv{µn|nNsuch that Cn6= 0}.
Consider a nonzero extremal point of C, which is therefore of the form µn0for some n0N. Hence,
there exists θ[0,2π]such that
eµn<e µn0,nN\ {n0}(3.24)
By letting z=ρe with ρRand using (3.20), we obtain
0 = Gρe eρeµn0=Cn0+X
nN\{n0}
Cneρ(eµne µn0).(3.25)
We then let ρ+in (3.25) to obtain, using (3.24), that Cn0= 0 which is a contradiction with the
construction C. It implies that Cn= 0,nN. The expression of the Cnimplies for all nN
1
nan
2
(λn)2
(λ2
n+λ2)2β2
nbn
2
(λ+n)2
(λ2
n+λ2)2= 0,
1
nan
2
(λ+n)2
(λ2
n+λ2)2+β2
nbn
2
(λn)2
(λ2
n+λ2)2= 0.
One then easily concludes that an=bn= 0 by using (3.22) and the fact that β1
n6= 0 and β2
n6= 0 for all
nN. Theorem 3.2 thus implies the Riesz basis property.
18
3.2 Definition and regularity of the feedback law
So far, we have obtained from (2.9) the expression of the kernels kij with respect to the Fourier coefficients
αi
nof the feedback
KΨ1
Ψ2:=
+
X
n=1
α1
nhΨ1, ϕni+α2
nhΨ2, ϕni.(3.26)
The regularity of the kernels and, consequently, the regularity of T, will be directly related to the decay
rate of those coefficients as n+. It remains to use the T B =Bcondition (2.18) to construct Kand
T.As µϕ1H2
(0) and Bis a Riesz basis of X2
(0) (see Proposition 3.4), it comes that there exist sequences
(a1
n)nN,(a2
n)nN2(N,R)such that
0
µϕ1=
+
X
n=1
a1
ng12
n
g22
n+a2
nh12
n
h22
n,in X2
(0).(3.27)
Getting back to the T B =Bcondition (2.18), we define
α1
n:= β1
n
hµϕ1, ϕnia1
n, α2
n:= β2
n
hµϕ1, ϕnia2
n.(3.28)
Then, following the heuristic of Sec. 2, the transformation Tis defined by (2.17) and the feedback law K
is defined by (3.26).
Unfortunately, the regularity of the coefficients is only (aj
n)nN2(N,R)for the X2
(0) Riesz basis
and will not be sufficient to prove that Tis continuous in X3
(0).
Remark 3.6 Recall that we assumed controllability of the linearized system i.e. Hypothesis 1.1 . From [5,
Remark 2], it follows that µ(0) ±µ(1) 6= 0 and then that (µϕ1)6∈ H3
(0). Thus, (0, µϕ1)Tcannot be
decomposed in the Riesz basis of X3
(0). This would have led to more regularity for the sequences (αj
n)n.
It is fortunate since, as it will be noticed in Remark 4.2, if (µϕ1)H3
(0), the obtained transformation
Twould have been compact in X3
(0) and thus not invertible in X3
(0).
Performing a suitable decomposition of the function µϕ1, we provethat the coefficients of the feedback
law satisfy the following regularity.
Proposition 3.7 We define the sequence (hn)nNby
hn:= 4
n3π2(1)n+1µ(1) µ(0).
Then the sequences (α1
n)nNand (α2
n)nNdefined in (3.28) satisfy
α1
n
n32(N,R),1
n3α2
nλnβ2
n
hn
hµϕ1, ϕni2(N,R).
Remark 3.8 As a corollary, it comes that, for every j {1,2},
αj
n
n3nN(N,R).
Proof:
First step: splitting of the problem. We start using the ideas developed in [44] to extend the regulariza-
tion result [5, Lemma 1] to higher dimensions. Let
h:x(0,1) 7→ π2
3(µ(0) + µ(1))x33µ(0)x2+ (2µ(0) µ(1))x.
It comes that
g:= µϕ1hH3
(0).(3.29)
19
The Fourier coefficients of hare given by
hh, ϕki=hk=4
k3π2(1)k+1µ(1) µ(0),kN,(3.30)
which is the leading term in the asymptotic expansion of hµϕ1, ϕkigiven in [5, Remark 2]. Let us split the
left-hand side of (3.27) in two parts
0
µϕ1=0
g+0
h.
As gH3
(0), using the Riesz basis of X3
(0) we get the existence of sequences (δ1
n)nNand (δ2
n)nN
such that 0
g=
+
X
n=1hµϕ1, ϕni"λ1/2
nδ1
n
β1
n g12
n1/2
n
g22
n1/2
n!+λ1/2
nδ2
n
β2
n h12
n1/2
n
h22
n1/2
n!#.(3.31)
The coefficients of the decomposition in a Riesz basis being 2sequences, Hypothesis 1.1 and the behaviour
of coefficients βj
nin (3.2) imply that, for every j {1,2},
δj
n
n3nN2(N,R).(3.32)
Second step: decomposition of h.Using the Riesz basis Bof X2
(0), we get coefficients (ρ1
n)nNand
(ρ2
n)nNsuch that 0
h=
+
X
n=1hµϕ1, ϕniρ1
n
β1
ng12
n
g22
n+ρ2
n
β2
nh12
n
h22
n.(3.33)
Recall that the basis Bis obtained as a perturbation of the L2-orthonormal basis. To highlight this, we
define the sequences (γj
n)nN,j {1,2}, such that
ρ1
n=λnβ1
nγ1
n,
ρ2
n=λnβ2
nhh, ϕni
hµϕ1, ϕni+λnβ2
nγ2
n.(3.34)
From (2.16), (3.33), and (3.34), we get, for every kN,
0
hh, ϕki=
+
X
n=1hµϕ1, ϕnihµϕ1, ϕkiρ1
nc12
nk
c22
nk+ρ2
nd12
nk
d22
nk
=
+
X
n=1hµϕ1, ϕnihµϕ1, ϕkihh, ϕni
hµϕ1, ϕniλnβ2
nd12
nk
d22
nk
+
+
X
n=1hµϕ1, ϕnihµϕ1, ϕkiλnβ1
nγ1
nc12
nk
c22
nk+λnβ2
nγ2
nd12
nk
d22
nk.
which, using (2.13), (2.14), (2.15), and (3.1), is equivalent to
P+
n=1hµϕ1, ϕkihh, ϕniλnβ2
nd12
nk
P+
n=1,n6=khµϕ1, ϕkihh, ϕniλnβ2
nd22
nk!
=
+
X
n=1hµϕ1, ϕnihµϕ1, ϕkiλnβ1
nγ1
nc12
nk
c22
nk+λnβ2
nγ2
nd12
nk
d22
nk,
for every kN. Hence, if we define
˜
h1
˜
h2:=
+
X
k=1
P+
n=1hh, ϕnihµϕ1, ϕkiλnβ2
nd12
nkϕk
P+
n=1,n6=khh, ϕnihµϕ1, ϕkiλnβ2
nd22
nkϕk
,(3.35)
20
we get, using also (2.16),
˜
h1
˜
h2=
+
X
n=1
+
X
k=1hµϕ1, ϕniλnγ1
nhµϕ1, ϕkiβ1
nc12
nk
hµϕ1, ϕkiβ1
nc22
nk+λnγ2
nhµϕ1, ϕkiβ2
nd12
nk
hµϕ1, ϕkiβ2
nd22
nkϕk
=
+
X
n=1
λ3/2
nhµϕ1, ϕni"γ1
n g12
n1/2
n
g22
n1/2
n!+γ2
n h12
n1/2
n
h22
n1/2
n!#.(3.36)
Finally, if (˜
h1,˜
h2)TX3
(0) (which will be proved in the next two steps), it comes that, for every j {1,2},
λ3/2
nhµϕ1, ϕniγj
nn2(N,R),
which, thanks to Hypothesis 1.1, implies that
(γj
n)n2(N,R).(3.37)
Using (3.27), (3.28), (3.29), (3.31), (3.33), and (3.34), we obtain
α1
n
n3=ρ1
n
n3+δ1
n
n3=λnβ1
n
n3γ1
n+δ1
n
n3,
α2
n
n3=ρ2
n
n3+δ2
n
n3=λnβ2
n
n3γ2
n+λnβ2
n
n3hh, ϕni
hµϕ1, ϕni+δ2
n
n3,
which, with (1.3), (3.2), (3.30), (3.32), and (3.37) will end the proof of Proposition 3.7.
Third step: H3
(0) regularity of ˜
h1.We start by proving that ˜
h1H3
(0) i.e.
+
X
k=1 k3h˜
h1, ϕki
2=
+
X
k=1 k3
+
X
n=1hµϕ1, ϕkihh, ϕniλnβ2
nd12
nk
2
<+.(3.38)
From (3.2) and (3.30) it comes that there exists C > 0such that
|hh, ϕniλnβ2
n| C, nN.
Using (1.5) together with (2.13), (2.14), (2.15) and (2.16), it suffices to prove that
M12 :=
+
X
k=1
+
X
n=1
d12
nk
2
=
+
X
k=1
+
X
n=1
λk(λ2+λ2
kλ2
n)
(λ2+ (λkλn)2) (λ2+ (λk+λn)2)
2
<+.(3.39)
Notice that
M12 2
+
X
k=1
λ2λk
λ2(λ2+ 4λ2
k)2
+
+
X
n=1
n6=k
d12
nk
2
.
As +
X
k=1 λ2λk
λ2(λ2+ 4λ2
k)2
<+,
we only deal with the second term. Straightforward computations lead to
+
X
n=1
n6=k
d12
nk =1
2
+
X
n=1
n6=kλkλn
λ2+ (λkλn)2+λk+λn
λ2+ (λk+λn)2.(3.40)
21
The second term of this sum is easily dealt with. As λ2+ (λk+λn)2(λk+λn)2it comes that
+
X
n=1
n6=k
λk+λn
λ2+ (λk+λn)2
+
X
n=1
n6=k
1
λk+λn
+
X
n=1 Zn
n1
1
λk+π2x2dx
=1
2 .
Thus,
+
X
k=1
+
X
n=1
n6=k
λk+λn
λ2+ (λk+λn)2
2
<+.(3.41)
We now turn back to the first term of the right-hand side of (3.40). We have
+
X
n=1
n6=k
λkλn
λ2+ (λkλn)2
+
X
n=1
n6=k
1
|λkλn|=
k1
X
n=1
1
λkλn
+
+
X
n=k+1
1
λnλk
.
Straightforward computations lead to
k1
X
n=1
1
λkλn
=1
22
k1
X
n=1 1
kn+1
k+n
=1
22
k1
X
j=1
1
j+
2k1
X
j=k+1
1
j
=1
22
2k1
X
j=1
1
j1
k
.
In the same spirit, for Nk+ 1,
0
N
X
n=k+1
1
λkλn
=1
22
N+k
X
j=2k+1
1
j1
22
Nk
X
j=1
1
j
and thus,
0
+
X
n=k+1
1
λnλk1
22
2k
X
j=1
1
j1
22
N+k
X
j=nk+1
1
j1
22
2k
X
j=1
1
j.
Finally,
+
X
n=1
n6=k
λkλn
λ2+ (λkλn)2
+
X
n=1
n6=k
1
|λkλn|1
22
2
2k1
X
j=1
1
j1
2k
,(3.42)
which belongs to 2(N)as P2k1
j=1 1
j
k+ln(2k1). From (3.40), this proves that ˜
h1H3
(0).
Fourth step: H3
(0) regularity of ˜
h2.We end the proof of Proposition 3.7 by proving that ˜
h2H3
(0) i.e.,
by (3.35),
+
X
k=1 k3h˜
h2, ϕki
2=
+
X
k=1
k3
+
X
n=1
n6=k
hµϕ1, ϕkihh, ϕniλnβ2
nd22
nk
2
<+.(3.43)
22
From (3.2) and (3.30) it comes that there exists C > 0such that
|hh, ϕniλnβ2
n| C, nN.
Using (1.5) it suffices to prove that
+
X
k=1
+
X
n=1
n6=k
d22
nk
2
=
+
X
k=1
+
X
n=1
n6=k
λ(λ2+λ2
k+λ2
n)
(λ2+ (λkλn)2) (λ2+ (λk+λn)2)
2
<+.(3.44)
Notice that
λ(λ2+λ2
k+λ2
n)
(λ2+ (λkλn)2) (λ2+ (λk+λn)2)=λ
21
λ2+ (λk+λn)2+1
λ2+ (λkλn)2
λ
2λk+λn
λ2+ (λk+λn)2+|λkλn|
λ2+ (λkλn)2
λ
2λk+λn
λ2+ (λk+λn)2+1
|λkλn|.
From (3.41) and (3.42), it comes that
+
X
k=1
+
X
n=1
n6=k
λ(λ2+λ2
k+λ2
n)
(λ2+ (λkλn)2) (λ2+ (λk+λn)2)
2
<+.
This ends the proof of Proposition 3.7.
3.3 Domain of definition and continuity of the transformation
The regularity of the coefficients obtained in the previous section is sufficient to define a continuous oper-
ator Tin the state space.
Proposition 3.9 The transformation Tdefined on X3
(0) by (2.17) and (3.28) is linear continuous in X3
(0).
Proof: Let 1,Ψ2)TX3
(0). From (2.17), using the Riesz basis property of Proposition 3.4, it comes
that TΨ1
Ψ2belongs to X3
(0) if the following sequences,
λ1/2
nα1
nhΨ2, ϕni
β1
nλ1/2
nα2
nhΨ1, ϕni
β1
n!nN
,
λ1/2
nα1
nhΨ1, ϕni
β2
n
+λ1/2
nα2
nhΨ2, ϕni
β2
n!nN
,
belong to 2(N,R). From (3.2) and Remark 3.8, it comes that, for every i, j, k {1,2},
+
X
n=1
λ1/2
nαj
nhΨi, ϕni
βk
n
2
C
+
X
n=1
αj
n
n3n3hΨi, ϕni
2
C
+
X
n=1 n3hΨi, ϕni2
=CkΨik2
H3
(0) .
Finally, we obtain
TΨ1
Ψ2
X3
(0) C
Ψ1
Ψ2
X3
(0)
.
23
4 Invertibility of the transformation
This section aims at proving the invertibility of T. As a first step, we prove in subsection 4.1 that Tis a
Fredholm operator. In subsection 4.1, we prove that the analogous of (1.15) in finite dimension holds on a
certain functional space. This will be used in subsection (4.3) to obtain the invertibility of T.
4.1 Fredholm form
The goal of this subsection is the proof of the following result.
Proposition 4.1 There exists e
T:X3
(0) X3
(0) invertible such that Te
Tis a compact operator.
The proof of this proposition rely on the study of the feedback law done in Proposition 3.7.
Proof: Let e
Tbe the transformation defined by (2.17) where the coefficients α1
nand α2
nare respectively
replaced by
˜α1
n= 0,˜α2
n=hh, ϕni
hµϕ1, ϕniλnβ2
n.
From Proposition 3.7 (recall that hn=hh, ϕniis given in (3.30)), it follows that defining ˆαj
n=αj
n˜αj
n,
we get that (ˆαj
n/n3)n2(N,R).
The computations done in the proof of Proposition 3.9 show that e
Tis a linear continuous operator of
X3
(0). We prove that e
Tis invertible. For any 1,Ψ2)TX3
(0),
e
TΨ1
Ψ2=
+
X
n=1 ˜α2
nλ1/2
n
β1
nhΨ1, ϕni g12
n1/2
n
g22
n1/2
n!+˜α2
nλ1/2
n
β2
nhΨ2, ϕni h12
n1/2
n
h22
n1/2
n!.
From (1.5), (1.7), (3.2), and (3.30), we have that, for every j {1,2},(βj
nλ3/2
n/(˜α2
nλ1/2
n))n(N,R).
Then, the Riesz basis property of Proposition 3.4 ends the proof of the invertibility of e
T.
Finally, we prove that Te
Tis compact using the Hilbert-Schmidt criterion, i.e., we prove that
+
X
n=1
(Te
T)ϕn3/2
n
0
2
X3
(0)
+
(Te
T)0
ϕn3/2
n
2
X3
(0)
<+.(4.1)
From (2.17) and the definition of e
Tit comes that
Te
TΨ1
Ψ2=
+
X
n=1 ˆα1
nhΨ2, ϕni
β1
nˆα2
nhΨ1, ϕni
β1
ng12
n
g22
n
+
+
X
n=1 ˆα1
nhΨ1, ϕni
β2
n
+ˆα2
nhΨ2, ϕni
β2
nh12
n
h22
n.
Thus,
(Te
T)ϕn3/2
n
0
2
X3
(0)
=
ˆα2
n
λ3/2
nβ1
ng12
n
g22
n+ˆα1
n
λ3/2
nβ2
nh12
n
h22
n
2
X3
(0)
=
ˆα2
n
λnβ1
n g12
n1/2
n
g22
n1/2
n!+ˆα1
n
λnβ2
n h12
n1/2
n
h22
n1/2
n!
2
X3
(0)
.
24
The first term is estimated in the following way
ˆα2
n
λnβ1
n g12
n1/2
n
g22
n1/2
n!
2
X3
(0)
2
ˆα2
n
λnβ1
n g12
n1/2
nϕn3/2
n
g22
n1/2
n!
2
X3
(0)
+ 2
ˆα2
n
λnβ1
nϕn3/2
n
0
2
X3
(0)
2ˆα2
n
λnβ1
n2
g12
n1/2
nϕn3/2
n
g22
n1/2
n!
2
X3
(0)
+ 2 ˆα2
n
λnβ1
n2
.
We proved in Lemma 3.5 that
+
X
n=1
g12
n1/2
nϕn3/2
n
g22
n1/2
n!
2
X3
(0)
<+.
As (ˆα2
n/n3)n2(N,R), this gives, using (3.2), that
+
X
n=1
ˆα2
n
λnβ1
n g12
n1/2
n
g22
n1/2
n!
2
X3
(0)
<+.
The term
+
X
n=1
ˆα1
n
λnβ2
n h12
n1/2
n
h22
n1/2
n!
2
X3
(0)
,
is dealt with in the exact same way, ending the proof of Proposition 4.1.
Remark 4.2 Notice that the key point in proving the invertibility of e
Tis that for any nN,hh, ϕni 6= 0.
We also underline that the compactness of Te
Tcomes from the fact that (ˆαj
n/n3)nN2(N,R).
If we had µϕ1H3
(0) (this is not possible due to Hypothesis 1.1, see [5, Remark 2]), then from (2.18)
we would have obtained that (αj
n/n3)nN2(N,R). This would have led to the compactness of the
transformation Tpreventing its invertibility.
4.2 Operator equality
We prove that the formal operator equality
T(A+BK) = AT λT (4.2)
holds true on an appropriate functional space. Recall that Kis defined by (3.26).
Remark 4.3 Notice that due to the regularity of the coefficients αj
nobtained in Proposition 3.7, the oper-
ator Kis not defined on X3
(0). Otherwise, taking any (0, ψ)TX3
(0) with ψ /Hs
(0) for any s > 3would
imply K0
ψ=X
nN
α2
nhψ, ϕni=X
nN
α2
n
n3(n3hψ, ϕni)<+,
and therefore (α2
n/n3)nNl2(N,R)which is in contradiction with Proposition 3.7.
25
Due to the previous remark, the functional setting for (4.2) to hold needs to be specified. Let us first deal
with K. Proposition 3.7 implies that, for every 1,Ψ2)TH3
(0) ×H4
(0),
X
nN|α1
nhΨ1, ϕni+α2
nhΨ2, ϕni| X
nN
α1
n
n3
2!1/2
kΨ1kH3
(0) +CX
nN
λ3/2
nhΨ2, ϕni
X
nN
α1
n
n3
2!1/2
kΨ1kH3
(0) +C X
nN
1
λn!1/2 X
nN
λ4
nhΨ2, ϕni2!1/2
CkΨ1kH3
(0) +CkΨ2kH4
(0) .
This shows that Kis well defined and continuous on H3
(0) ×H4
(0).
We now turn to A+BK. Recall that we expect
(A+BK)Ψ1
Ψ2=0
0 Ψ1
Ψ2+KΨ1
Ψ2 0
µϕ1.(4.3)
We define
D(A+BK) := Ψ1
Ψ2(H5H3
(0))×H5
(0) ; ∆Ψ1+KΨ1
Ψ2µϕ1H3
(0).
and then define A+BK on D(A+BK )by (4.3). Note that 1,Ψ2)TD(A+BK )if and only if
1,Ψ2)(H5H3
(0))×H5
(0) and
2Ψ1(x) + 2KΨ1
Ψ2µ(x)ϕ
1(x) = 0, x {0,1}.(4.4)
We now prove the density of D(A+BK).
Lemma 4.4 The domain D(A+BK )is dense in X3
(0).
Proof: Let us prove that D(A+BK)={0}in X3
(0). Let 1,Ψ2)TD(A+BK ).
First step: we prove that Ψ1= 0.
Let kN. Consider φ1:= ϕk. From the asymptotic behaviour of α2
n(see Proposition 3.7), there
exists NNsuch that nN,α2
n6= 0. Therefore, setting φ2=α1
kϕn2
nH5
(0) implies that
Kφ1
φ2=α1
k+α2
nα1
k
α2
n= 0
which means that (4.4) is satisfied and (φ1, φ2)TD(A+BK). Thus,
0 = *Ψ1
Ψ2, ϕk
α1
k
α2
nϕn!+X3
(0)
=Ψ1, ϕkH3
(0) α1
k
α2
nΨ2, ϕnH3
(0)
=Ψ1, ϕkH3
(0) λ3/2
nα1
k
α2
nλ3/2
nΨ2, ϕn
n+Ψ1, ϕkH3
(0)
.
Indeed, since Ψ2H3
(0), it implies that n3hΨ2, ϕninN2(N;R)and Remark 1.2, (3.2) and (3.34)
imply that λ3/2
nα1
k2
nn(N\ {1, ..., N 1};R). We conclude that
Ψ1, ϕkH3
(0)
= 0,kN,
and thus Ψ1= 0.
26
Second step : we prove that Ψ2= 0.
For all kN, let φ2=ϕk. If there exists nNsuch that α1
n6= 0, then setting φ1=α2
kϕn1
n
implies that Kφ1
φ2= 0 and therefore (φ1, φ2)TD(A+BK). Thus,
0 = *Ψ1
Ψ2, α2
k
α1
nϕn
ϕk!+X3
(0)
=Ψ2, ϕkH3
(0)
.
Otherwise, for every φ1H3
(0),
Kφ1
φ2=α2
k.
Then we consider φ1solution to (φ1=α2
k(µϕ1)
φ1(0) = φ1(1) = 0.
Since µϕ1H3H1
0, then φ1H5H3
(0) and, since (4.4) is satisfied, (φ1, φ2)TD(A+BK).
Moreover, as Ψ1= 0,
0 = Ψ1
Ψ2,φ1
φ2X3
(0)
=Ψ2, ϕkH3
(0)
.
This proves that 1,Ψ2)T= 0.
Let us now turn our attention to the kernel system. More precisely, we prove the following.
Proposition 4.5 For 1,Ψ2)TD(A+BK),
T(A+BK)Ψ1
Ψ2= (AλI)TΨ1
Ψ2,in X3
(0).
Proof: Let 1,Ψ2)TD(A+BK ). Then
(A+BK)Ψ1
Ψ2=
∆Ψ2
∆Ψ1+KΨ1
Ψ2µϕ1
.
Therefore
T(A+BK)Ψ1
Ψ2=X
nNα1
n
β1
n∆Ψ1+KΨ1
Ψ2µϕ1, ϕn+α2
n
β1
n∆Ψ2, ϕng12
n
g22
n
+α1
n
β2
n∆Ψ2, ϕn+α2
n
β2
n∆Ψ1+KΨ1
Ψ2µϕ1, ϕnh12
n
h22
n
=KΨ1
Ψ2X
nNα1
n
β1
ng12
n
g22
n+α2
n
β2
nh12
n
h22
nhµϕ1, ϕni
X
nN
λnα1
n
β1
nΨ1, ϕn+α2
n
β1
nΨ2, ϕng12
n
g22
n
+λnα1
n
β2
nΨ2, ϕnα2
n
β2
nΨ1, ϕnh12
n
h22
n
= X
nN
α1
nhΨ1, ϕni+α2
nhΨ2, ϕni!0
µϕ1
+X
nNλnα1
n
β1
nΨ1, ϕn+α2
n
β1
nΨ2, ϕng12
n
g22
n
+λnα1
n
β2
nΨ2, ϕnα2
n
β2
nΨ1, ϕnh12
n
h22
n,in X2
(0),
27
using the expression of Tin (2.17), the T B =Bcondition (2.18) and the expression of Kgiven in (3.26).
Moreover, by the definition of gij
n, hij
ngiven in (2.13) and by the relations (2.15), we have, on one hand,
using (2.17) once more,
(AλI)TΨ1
Ψ2,ϕk
0(L2)2
=TΨ1
Ψ2, Aϕk
0(L2)2λTΨ1
Ψ2,ϕk
0(L2)2
=TΨ1
Ψ2,λϕk
λkϕk(L2)2
=X
nNα1
n
β1
nhΨ2, ϕni α2
n
β1
nhΨ1, ϕniλkhg22
n, ϕki λhg12
n, ϕki
+α1
n
β2
nhΨ1, ϕni+α2
n
β2
nhΨ2, ϕniλkhh22
n, ϕki λhh12
n, ϕki
=X
nNα1
n
β1
nhΨ2, ϕni α2
n
β1
nhΨ1, ϕniλnhg11
n, ϕki
+α1
n
β2
nhΨ1, ϕni+α2
n
β2
nhΨ2, ϕniλnhh11
n, ϕki
=X
nNα1
n
β2
nhΨ2, ϕni α2
n
β2
nhΨ1, ϕniλnhh12
n, ϕki
α1
n
β2
nhΨ1, ϕni+α2
n
β2
nhΨ2, ϕniλnhg12
n, ϕki
=T(A+BK)Ψ1
Ψ2,ϕk
0(L2)2
.
On the other hand, using again (2.13), (2.15) and (2.17)
(AλI)TΨ1
Ψ2,0
ϕk(L2)2
=TΨ1
Ψ2,λkϕk
λϕk(L2)2
=X
nNα1
n
β1
nhΨ2, ϕni α2
n
β1
nhΨ1, ϕniλhg22
n, ϕki λkhg12
n, ϕki
+α1
n
β2
nhΨ1, ϕni+α2
n
β2
nhΨ2, ϕniλhh22
n, ϕki λkhh12
n, ϕki
=X
nNα1
n
β2
nhΨ2, ϕni α2
n
β2
nhΨ1, ϕniλnhh22
n, ϕki
α1
n
β1
nhΨ1, ϕni+α2
n
β1
nhΨ2, ϕniλnhg22
n, ϕki
+X
nNα1
nhΨ1, ϕni+α2
nhΨ2, ϕnihµϕ1, ϕki
=T(A+BK)Ψ1
Ψ2,0
ϕk(L2)2
.
Indeed, (2.13) and (2.15) imply
λkhh12
n, ϕki λhh22
n, ϕki=λnhh21
n, ϕki+β2
nhµϕ1, ϕki
=λn
β2
n
β1
nhg22
n, ϕki+β2
nhµϕ1, ϕki.
Consequently, by the definition of D(A+BK)and by the continuity of Tfrom X3
(0) into itself, T(A+
BK ) = (AλI )Tholds in X3
(0). Notice that all the previous infinite sums are converging due to the
regularity assumptions on the functions of D(A+BK).
28
We conclude this section by noting that, following Proposition 4.5, if 1
Ψ2)TD(A+BK), then TΨ1
Ψ2X5
(0). Indeed,
AT Ψ1
Ψ2=T(A+BK)Ψ1
Ψ2λT Ψ1
Ψ2,
where the right-hand side of the last equation is in X3
(0).
4.3 Invertibility
Let us now turn to the invertibility of Tand prove the following result.
Proposition 4.6 The operator Tis invertible from X3
(0) into itself.
Proof: In the previous section, we have proven that e
Tis invertible and Te
Tis a compact operator.
Consequently, the index of Tis equal to zero. Thus, it is sufficient to prove that Ker T={0}to show the
invertibility of T.
Let us rewrite the operator equality (4.2) under the form
T(A+BK +λI +ρI) = AT +ρT = (A+ρI )T , (4.5)
where ρCwill be chosen later. Assume for the moment that ρCis such that
(A+ρI +BK +λI)is an invertible operator from D(A+BK )to X3
(0),(4.6)
(A+ρI)is an invertible operator from D(A)to X3
(0).(4.7)
Note that here, the vector spaces D(A+BK), D(A)and X3
(0) are complexified. Then, from (4.5), one has
(A+ρI)1T=T(A+BK +λI +ρI)1.(4.8)
Consider (χ1, χ2)TKer T. Then, for all (φ1, φ2)TX3
(0), we have
0 = h(A+ρI)1Tφ1
φ2T(A+BK +λI +ρI)1φ1
φ2,χ1
χ2iX3
(0)
=hφ1
φ2, T ((A+ρI)1)χ1
χ2iX3
(0) h(A+BK +λI +ρI)1φ1
φ2, T χ1
χ2iX3
(0)
=hφ1
φ2, T ((A+ρI)1)χ1
χ2iX3
(0) .
Thus the space Ker T, which is of finite dimension, is stable by ((A+ρI))1. Hence, if
Ker T6={0},(4.9)
((A+ρI))1has an eigenfunction in Ker T. This eigenfunction is also an eigenfunction of (A)1.
Thus, if (4.9) holds, there exists νCand (χ1, χ2)TKer T\ {0}such that
(A)1χ1
χ2=1χ2
1χ1=νχ1
χ2.(4.10)
Hence, for every jN,
ν2hχ1, ϕji=νh−1χ2, ϕji=1
λj
νhχ2, ϕji=λjh1χ1, ϕji=1
λ2
jhχ1, ϕji.(4.11)
Therefore,
(ν2+1
λ2
j
)hχ1, ϕji= 0.(4.12)
29
Note that χ1= 0 together with (4.11) implies χ2= 0. Hence, since (χ1, χ2)T6= 0,χ16= 0, which with
(4.12) implies that there exists one and only one kNsuch that
ν=±i1
λk
, χ1=ckϕk, ckC\ {0}.
Furthermore, from (4.10), we obtain χ2=ickϕk. Finally, we have, by the T B =Bcondition (2.18),
±ickλ3
khµϕ1, ϕki=hµϕ1, χ2iH1
0,H5
(0)
=h0
µϕ1,χ1
χ2i(H1
0)2,(H5
(0))2
=hT0
µϕ1,χ1
χ2i(H1
0)2,(H5
(0))2
=h0
µϕ1, T χ1
χ2i(H1
0)2,(H5
(0))2
=0.
Since hµϕ1, ϕki 6= 0, we conclude that ckmust be zero, which implies that (4.9) does not hold and
therefore Ker T={0}.
It remains to prove the existence of ρCsuch that (4.6) and (4.7) hold. Let κ:= ρ+λ. Applying
A1to A+BK +κI yields the operator
I+A1BK +κA1:D(A+BK)D(A),(4.13)
where A1B= (∆1(µϕ1),0)T. Let us prove that the set of κCsuch that I+A1BK +κA1is
invertible from D(A+BK )to D(A)is non-empty.
First, if K(A1B)6=1, then the operator I+A1BK :D(A+BK)D(A)is invertible and the
proof is over. Indeed, to solve
(I+A1BK)ψ=f, (4.14)
for any fD(A), one applies Kto (4.14) (K(ψ),K(A1B)and K(f)are well-defined in this case)
leading to
K(ψ)(1 + K(A1B)) = K(f).
Since K(A1B)6=1, we use the expression of K(ψ)in (4.14) to obtain
ψ=fA1BK(f)
1 + K(A1B).
Suppose then that K(A1B) = 1. It corresponds to the case where A1BD(A+BK). Notice that
0is an eigenvalue of I+A1BK of algebraic multiplicity 1. Then, from [47], there exists an open set
Cof 0Csuch that there exist an holomorphic function κ7→ λ(κ)Cand an holomorphic
function κ7→ x(κ)D(A+BK )such that
x(0) = A1B=1(µϕ1)
0,
(I+A1BK +κA1)x(κ) = λ(κ)x(κ).(4.15)
If λ(κ)6= 0 in a small neighborhood of 0, then I+A1BK +κA1is invertible for κclose to 0and the
proof is over. Suppose then that λ(κ) = 0 in a small neighborhood of 0. In this case, consider the power
serie expansion of xaround 0
x=1(µϕ1)
0+
X
k=1
κkx1
k
x2
k.
Notice that since xD(A+BK )and A1BD(A+BK), we obtain that (x1
k, x2
k)TD(A). At the
zeroth order, (4.15) writes
1(µϕ1)
0+1(µϕ1)
0K1(µϕ1)
0=0
0,
30
from the hypothesis K(A1B) = 1. At the higher order, we have
x1
k
x2
k+1(µϕ1)
0Kx1
k
x2
k+κA1x1
k1
x2
k1=0
0,(4.16)
where (x1
0, x2
0)T:= (∆1(µϕ1),0)T. Taking Kof (4.16) yields
KA1x1
k
x2
k= 0,k0.
By successively taking A1and Kof (4.16), we obtain
KAnx1
k
x2
k= 0,k0,n1.(4.17)
Therefore from (4.16) and (4.17)
KA12n0
(µϕ1)=X
jN
α1
jh(1)n12n(µϕ1), ϕni
=X
jN
(1)nα1
jhµϕ1, ϕni
λ1+2n
j
=0.(4.18)
Consider the entire function
H(z) := X
jN
α1
jhµϕ1, ϕniez/λ2
j
λ3
j
.
From (4.18), we obtain that H(p)(0) = 0 and therefore H0. By letting z −∞ and by Hypothesis
1.1, we deduce that
α1
j= 0,j1.
In the same fashion,
KA2n0
(µϕ1)=X
jN
α2
jh(1)n2n(µϕ1), ϕni
=X
jN
(1)nα1
jhµϕ1, ϕni
λ2n
j
=0.(4.19)
Consider the entire function
ˆ
H(z) := X
jN
α2
jhµϕ1, ϕniez/λ2
j
λ2
j
.
From (4.19), we obtain that ˆ
H(p)(0) = 0 and therefore H0. By letting z −∞ and by Hypothesis
1.1, we deduce that
α2
j= 0,j1.
From Proposition 3.7, we know that α2
j6= 0,j1. Hence a contradiction either with K(A1B) = 1,
which implies the invertibility of I+A1BK +κA1for all κC, or with the fact that λ(κ) = 0 in a
small neighborhood of 0, which implies that I+A1BK +κA1is invertible in a small neighborhood of
0. Since (A+ρI)has discrete eigenvalues, it is possible in those two cases to choose ρsuch that (4.6)-(4.7)
are satisfied.
31
5 Well-posedness of the closed-loop linear system and rapid stabi-
lization
This section ends the proof of Theorem 1.5. Due to Remark 4.3, the feedback Kis not well defined for
functions in X3
(0). In subsection 5.1, we give a meaning to the solution 1,Ψ2)Tof the closed-loop
system
t Ψ1
Ψ2!= 0
0 ! Ψ1
Ψ2!+K Ψ1
Ψ2! 0
(µϕ1)(x)!,(t, x)(0, T )×(0,1),
Ψ1(t, 0) = Ψ1(t, 1) = 0,Ψ2(t, 0) = Ψ2(t, 1) = 0, t (0, T ),
Ψ1(0, x) = Ψ1
0(x),Ψ2(0, x) = Ψ2
0(x), x (0,1)
(5.1)
by proving that A+BK generates a C0semigroup. Finally we conclude to the exponential stability using
the operator equality of Proposition 4.5 and the invertibility of the transformation T.
5.1 Well-posedness of the closed-loop linear system
Let us first show,
Proposition 5.1 The operator (A+BK)defined on D(A+BK)generates a C0semigroup on X3
(0).
Thus, there exists a unique solution C([0, T ]; X3
(0))of (1.29) with v(t) = Kψ1(t, .)
ψ2(t, .).
Proof: We prove that (A+BK)is the infinitesimal generator of a C0semigroup on X3
(0).
First step. The density of D(A+BK )in X3
(0) was proven in Lemma 4.4.
Second step. Let us prove that (A+BK)is closed. Let (ψ1
n, ψ2
n)TD(A+BK)such that
ψ1
n
ψ2
n
n+ψ1
ψ2,in X3
(0),
(A+BK)ψ1
n
ψ2
n
n+φ1
φ2,in X3
(0).
We have
(A+BK)ψ1
n
ψ2
n=
ψ2
n
ψ1
n+Kψ1
n
ψ2
nµϕ1
.
Hence
ψ2
n
n+ψ2,in H1
0
ψ2
n
n+φ1,in H3
(0),
and, consequently, ψ2=φ1,ψ2H5
(0) and ψ2
n
n+ψ2in H5
(0). Therefore, Kψ1
ψ2is well-
defined and
Kψ1
n
ψ2
nKψ1
ψ2=X
kN
α1
khψ1
nψ1, ϕki+α2
khψ2
nψ2, ϕki
X
kNα1
k
k32!1/2
kψ1
nψ1kH3
(0) + X
kNα2
k
k42!1/2
kψ2
nψ2kH4
(0)
n+0.
32
We obtain
ψ1
n+Kψ1
n
ψ2
nµϕ1
n+φ2in H3
(0),
ψ1
n+Kψ1
n
ψ2
nµϕ1
n+ψ1+Kψ1
ψ2µϕ1in H1
0.
We conclude that ψ1+Kψ1
ψ2µϕ1=φ2and therefore (ψ1, ψ2)TD(A+BK ).
Third step. Let us now prove the dissipativity of (A+BK). Since Tis invertible from X3
(0) into itself,
we define the norm k · kT:= kT· kX3
(0) , which is equivalent to the X3
(0) norm. We denote ,·iTthe
associated inner product.
Consider (ψ1, ψ2)TD(A+BK). From Proposition 4.5, we have, since Tψ1
ψ2X5
(0),
h(A+BK)ψ1
ψ2,ψ1
ψ2iT=hT(A+BK)ψ1
ψ2, T ψ1
ψ2iX3
(0)
=hAT ψ1
ψ2, T ψ1
ψ2iX3
(0) λ
ψ1
ψ2
2
T
=λ
ψ1
ψ2
2
T0.
Consider now the dissipativity of (A+BK ). First, let (ψ1, ψ2)TD(A+BK )and (φ1, φ2)T
X3
(0). We have
h(A+BK)ψ1
ψ2,φ1
φ2iT=hT(A+BK)ψ1
ψ2, T φ1
φ2iX3
(0)
=hψ1
ψ2,(T1ATλI)φ1
φ2iT.
which implies that D((A+BK)) = n(φ1, φ2)TX3
(0) Tφ1
φ2X5
(0)o. Then, for (φ1, φ2)T
D((A+BK)), we have
h(A+BK)φ1
φ2,φ1
φ2iT=h(T1ATλI)φ1
φ2,φ1
φ2iT
=hATφ1
φ2, T φ1
φ2iX3
(0) λ
Tφ1
φ2
2
X3
(0)
0.
Thus from the Lumer-Philipps theorem (see e.g. [43, Corollary 4.4]) we obtain that A+BK generates
aC0semigroup on X3
(0).
5.2 Proof of the rapid stabilization
This section is dedicated to the proof of the main result, the rapid stabilization stated in Theorem 1.5.
Proof: To begin, let us assume that 1
0,Ψ2
0)TD(A+BK). Then, from Proposition 5.1, we get that
the associated solution of (1.29) with the feedback law v(t) = K1(t, .),Ψ2(t, .))Tis given by
Ψ1(t, .)
Ψ2(t, .)=et(A+BK)Ψ1
0
Ψ2
0.(5.2)
33
Let ξ1(t, .)
ξ2(t, .):= TΨ1(t, .)
Ψ2(t, .). Using the equality of operators given in Proposition 4.5, it comes that
d
dtξ1(t, .)
ξ2(t, .)=Td
dtΨ1(t, .)
Ψ2(t, .)
=T(A+BK)Ψ1(t, .)
Ψ2(t, .)
= (AT λT )Ψ1(t, .)
Ψ2(t, .)
=Aξ1(t, .)
ξ2(t, .)λξ1(t, .)
ξ2(t, .).
Thus, we obtain the following exponential stability of (ξ1(t, .), ξ2(t, .))T
ξ1(t, .)
ξ2(t, .)
X3
(0) eλt
ξ1
0
ξ2
0
X3
(0)
=eλt
TΨ1
0
Ψ2
0
X3
(0)
.
From the continuity and invertibility of Tin X3
(0) (see Proposition 4.6) it comes that
Ψ1(t, .)
Ψ2(t, .)
X3
(0)
T1
ξ1(t, .)
ξ2(t, .)
X3
(0)
T1
kTkeλt
Ψ1
0
Ψ2
0
X3
(0)
.(5.3)
Finally, let Ψ0H3
(0). Then, (0)
0))TX3
(0). The stability estimate (5.3) and the density proved in Lemma 4.4 ends the proof of
Theorem 1.5.
Acknowledgements. This work has been partially carried out thanks to the support of ARCHIMEDE
LabEx (ANR-11-LABX- 0033), the A*MIDEX project (ANR-11-IDEX-0001-02) funded by the “In-
vestissements d’Avenir” French government program managed by the ANR, the ANR grant EMAQS
No.ANR-2011-BS01-017-01, the ERC avanced grant 266907 (CPDENL) of the 7th Research Framework
Programme (FP7) and the FQRNT.
A Simplified Saint-Venant Equation Example
Let us provide an explicit transformation (T , K)which allows to stabilize exponentially rapidly the sim-
plified Saint-Venant equation
ht+vx= 0,(t, x)(0, T )×(0,1),
vt+hx=u(t),(x, t)(0, T )×(0,1),
h(t, 0) = v(t, 1) = 0, t (0, T ),
h(0, x) = h0(x), v(0, x) = v0(x), x (0,1),
(A.1)
which is controllable in time T > 2.
Let H:= hxand V:= vx. Then, the equation on (H, V )writes
Ht+Vx= 0,(t, x)(0, T )×(0,1),
Vt+Hx= 0,(t, x)(0, T )×(0,1),
H(t, 1) = u(t), t (0, T ),
V(t, 0) = 0, t (0, T ),
H(0, x) = H0(x), V (0, x) = V0(x), x (0,1),
34
with H0= (h0)xand V0= (v0)x. Consider now R1:= H+Vand R2:= HV. Then
R1
t+R1
x= 0,(x, t)(0, T )×(0,1),
R2
tR2
x= 0,(x, t)(0, T )×(0,1),
(R1+R2)(t, 1) = 2u(t), t (0, T ),
(R1R2)(t, 0) = 0, t (0, T ),
R1(0, x) = R1
0(x), R2(0, x) = R2
0(x), x (0,1),
with R1
0:= H0+V0and R2
0:= H0V0. Let us consider a transformation which maps (R1, R2)to a
solution of a target stable system, that is, e
R1:= eλxR1/cosh(λ)and e
R2:= eλxR2/cosh(λ), for λ > 0.
A straightforward computation leads to
e
R1
t+e
R1
x+λe
R1= 0,(x, t)(0, T )×(0,1),
e
R2
te
R2
x+λe
R2= 0,(x, t)(0, T )×(0,1),
(e
R1+e
R2)(t, 1) = 2eλ(u(t)/cosh(λ)) + 2 tanh(λ)R2(t, 1), t (0, T )
(e
R1e
R2)(t, 0) = 0, t (0, T ),
e
R1(0, x) = e
R1
0(x),e
R2(0, x) = e
R2
0(x), x (0,1).
(A.2)
Hence, the exponential stability of (A.2) is obtained if 2eλu(t)/cosh(λ) + 2 tanh(λ)R2(t, 1) = 0. In
terms of the original variables, it implies that
0 = 2eλu(t)/cosh(λ) + 2 tanh(λ)R2(t, 1)
=2eλu(t)/cosh(λ) + 2 tanh(λ)(hx(t, 1) vx(t, 1))
=2eλu(t)/cosh(λ)2 tanh(λ)(vx(t, 1) + u(t)),
that is, u(t) = tanh(λ)vx(t, 1).
The target system of (A.1) is given by
e
ht+evx+λe
h= 0,(x, t)(0, T )×(0,1),
evt+e
hx+λev= 0,(x, t)(0, T )×(0,1),
e
h(t, 0) = ev(t, 1) = 0, t (0, T ),
e
h(0, x) = e
h0(x),ev(0, x) = ev0(x), x (0,1),
(A.3)
One can recover an explicit expression of the transformation (T , K)leading to this target system using
e
hx+evx=eλx(hx+vx)/cosh(λ),
e
hxevx=eλx(hxvx)/cosh(λ),(A.4)
which boils down to
e
hx= (cosh(λx)hxsinh(λx)vx)/cosh(λ)
evx= (sinh(λx)hx+ cosh(λx)vx)/cosh(λ).
Using the boundary conditions, one obtains the explicit transformation T
e
h(x) = 1
cosh(λ)cosh(λx)h(x)λZx
0
sinh(λy)h(y)dy sinh(λx)v(x) + λZx
0
cosh(λy)v(y)dy
=1
cosh(λ)Z1
0δx=ycosh(λy)λ1(0,x)(y) sinh(λy)h(y)dy
+Z1
0λ1(0,x)(y) cosh(λy)δx=ysinh(λy)v(y)dy,
ev(x) = 1
cosh(λ)sinh(λ)h(1) sinh(λx)h(x)Z1
x
λcosh(λy)h(y)dy
35
+ cosh(λx)v(x) + λZ1
x
sinh(λy)v(y)dy
=1
cosh(λ)Z1
0δy=1 sinh(λ)δx=ysinh(λy)λ1(x,1)(y) cosh(λy)h(y)dy
+Z1
0λ1(x,1)(y) sinh(λy) + δx=ycosh(λy)v(y)dy.
Moreover, the explicit transformation Kwrites
u(t) = tanh(λ)Z1
0
δ
y=1v(t, y)dy.
If one writes, in the same spirit as (2.6), the kernels equation for (A.1), then one obtains that the kernels of
the transformations (T , K)exhibited here are the solution of this system. One also verifies that, thanks to
the factor 1/cosh(λ), the T B =Bcondition is verified by the transformation T. Getting back to (A.4)
one sees that the inverse of Tcan be computed explicitly performing similar computations.
Moreover, the Fourier coefficients of the kernels system associated to (A.1) have the same expression
as (2.13), where the eigenvalues/eigenfunctions are replaced by those associated with (A.1) and the Fourier
coefficients of the control operator µϕ1are replaced by the one of (A.1), that is 1.
It is also noticeable that the Fourier coefficients of the kernel α2are
α2
n= (1)n(πn) tanh(λ),
which is adequate with the perturbation argument used in Proposition 3.7 to obtain the Fourier coefficients
from the T B =Bcondition. One notice that, for (A.1), α10. Whether α10or not in the case of the
linearized Schr¨odinger equation cannot be verified with our analysis.
B Quadratically close families
This section is devoted to the proof of Lemma 3.5.
Proof: Let s= 2 or 3. To simplify the notations, let,
˜cij
nk := cij
nkβ1
nhµϕ1, ϕki,˜
dij
nk := dij
nkβ2
nhµϕ1, ϕki,
where cij
nk and dij
nk are defined by (2.16) and βj
nare defined by (3.1).
First step: let us prove that
X
nN
ϕns/2
n
0 g12
n(s2)/2
n
g22
n(s2)/2
n!
2
Xs
(0)
<+.
First, by denoting k=m+n, we have using in particular (1.3), (1.5), (2.13) and (2.15)
X
nN
ϕn
λs/2
ng12
n
λ(s2)/2
n
2
Hs
(0)
=X
nNX
kN\{n}
λs/2
k˜c12
nk
λ(s2)/2
n
2
=X
nNX
0<|m|<n
mZ
+X
n<m
mNλs
n+m
λs
nλn˜c12
nn+m2
=λ4X
nNX
0<|m|<n
mZ
+X
n<m
mNλs
n+m
λs
n
(λ2+ 4λ2
n)λn+m
δnn+m(λ)λn
2hµϕ1, ϕn+mi
hµϕ1, ϕni
2
4X
nNX
0<|m|<n
mZ
+X
n<m
mNλn+m
λns1
λ2+ 4λ2
n
δnn+m(λ)
2
.(B.1)
36
The two sums of (B.1) are dealt with separately.
Consider first the case where 0<|m|< n and mZ. We have, using in particular (2.14)
(λ2+ 4λ2
n)
δnn+m(λ)=1
m2n2
λ2
n4+ 4π4
λ2
n4+π2+π21 + m
n2 λ2
(mn)2+m
nπ2+ 2π22
1
m2n2
λ4+ 4π4
π8,(B.2)
and
λn+m=1 + m
n2λn4λn.(B.3)
Thus, for the first term of the right-hand side of (B.1)
X
nNX
0<|m|<n
mZλn+m
λns1
(λ2+ 4λ2
n)
δnn+m(λ)
2
CX
nNX
0<|m|<n
mZ
1
m4n4<+.(B.4)
Consider now the case where m > n. We have
(λ2+ 4λ2
n)
δnn+m(λ)=n4
m8
λ2
n4+ 4π4
λ2
m4+1 + n
m2π2+n
m2π22λ2
m4+π2+ 2 n
mπ22,
n4
m8
λ2+ 4π4
π8(B.5)
and λn+m
λn
=(n+m)2
n2=m2
n21 + n
m24m2
n2.(B.6)
Thus, for the second term of the right-hand side of (B.1) yields,
X
nNX
n<m
mNλn+m
λns1
(λ2+ 4λ2
n)
δnn+m(λ)
2
CX
nNX
n<m
mNn
m82(s1) 1
m8
CX
nNX
n<m
mN
1
m4
1
n4<+.(B.7)
Then inequalities (B.1), (B.4) and (B.7) imply that
X
nN
ϕn
λs/2
ng12
n
λ(s2)/2
n
2
Hs
(0)
<+.(B.8)
The other sum is treated in a similar way. Indeed,
X
nN
g22
n
λ(s2)/2
n
2
Hs
(0)
=X
nNX
kN
λs/2
k˜c22
nk
λ(s2)/2
n
2
=X
nNX
0<|m|<n
mZ
+X
n<m
mNλn+m
λnsλn˜c22
nn+m2+X
nNλn˜c22
nn2
=λ2
4X
nNX
0<|m|<n
mZ
+X
n<m
mNλn+m
λns
λ2+ 4λ2
n
δnn+m(λ)
2
λ2λ2
n+m+λ2
n
λn
2hµϕ1, ϕn+mi
hµϕ1, ϕni
2
+X
nNλ
2λn2
C+CX
nNX
0<|m|<n
mZ
+X
n<m
mNλn+m
λns3
λ2+ 4λ2
n
δnn+m(λ)
2
λ2λ2
n+m+λ2
n
λn
2
.(B.9)
37
Consider first the case where 0<|m|< n. We have
λ2λ2
n+m+λ2
n
λnCn3|m|
n2
λ2
n3mπ44 + 6 m
n+ 4 m
n2+m
n3Cn|m|λ2+ 15π4.
Using (B.2), it comes that
X
nNX
0<|m|<n
mZλn+m
λns3
λ2+ 4λ2
n
δnn+m(λ)
2
λ2λ2
n+m+λ2
n
λn
2
CX
nNX
0<|m|<n
mZλn+m
λns31
n2m2.
This sum is clearly finite for s= 3. For s= 2, we notice that the previous series is the general term of a
Cauchy product. Indeed,
X
0<|m|<n
mZλn+m
λns31
n2m2=X
0<|m|<n
mZ
1
m2
1
(n+m)2
2
n1
X
m=1
1
m2
1
(nm)2+2
n2X
mN
1
m2.
Thus, for s {2,3},
X
nNX
0<|m|<n
mZλn+m
λns3
λ2+ 4λ2
n
δnn+m(λ)
2
λ2λ2
n+m+λ2
n
λn
2
<+.(B.10)
Consider now the case where m > n. We have
λ2λ2
n+m+λ2
n
λn=
m4λ2
m41 + n
m4π4+n
m4
n2π2Cm4
n2,
and λn+m
λns3
1.
Using (B.5), it then comes that,
X
nNX
n<m
mNλn+m
λns3
λ2+ 4λ2
n
δnn+m(λ)
2
λ2λ2
n+m+λ2
n
λn
2
CX
nNX
n<m
mNm4
n2
n4
m82
CX
nNX
n<m
mN
1
m2
1
n2<+.(B.11)
Then, inequalities (B.9), (B.10) and (B.11) imply that
X
nN
g22
n
λ(s2)/2
n
2
Hs
(0)
<+.(B.12)
Together with (B.8) it ends the first step.
Second step: let us prove that
X
nN
0
ϕns/2
n h12
n(s2)/2
n
h22
n(s2)/2
n!
2
Xs
(0)
<+.
38
This proof is very similar to the first step. Thus we give the expressions of the different sums but we
do not detail every computation:
X
nN
h12
n
λ(s2)/2
n
2
Hs
(0)
=X
nNX
kN
λs/2
k˜
d12
nk
λ(s2)/2
n
2
=X
nNX
0<|m|<n
mZ
+X
n<m
mNλn+m
λnsλn˜
d12
nn+m
2+X
nNλn˜
d12
nn
2
=λ2X
nNX
0<|m|<n
mZ
+X
n<m
mNλs
n+m
λs
n
λ2+ 4λ2
n
δnn+m(λ)
2
λn+mλ2+λ2
n+mλ2
n
λ2+ 2λ2
n
2hµϕ1, ϕn+mi
hµϕ1, ϕni
2
+X
nN
λ2λn
λ2+λ2
n
2
<+,
and
X
nN
ϕn
λs/2
nh22
n
λ(s2)/2
n
2
Hs
(0)
=X
nNX
kN\{n}
λs/2
k˜
d22
nk
λ(s2)/2
n
2
=X
nNX
0<|m|<n
mZ
+X
n<m
mNλs
n+m
λs
nλn˜
d22
nn+m
2
=λ4X
nNX
0<|m|<n
mZ
+X
n<m
mNλs
n+m
λs
n
λ2+ 4λ2
n
δnn+m(λ)
2
λ2+λ2
n+m+λ2
n
λ2+ 2λ2
n
2hµϕ1, ϕn+mi
hµϕ1, ϕni
2
<+.
This completes the proof of Lemma 3.5.
C Rapid stabilization of the linearized system
In this section we detail how Theorem 1.4 can be obtained from the results developed in this article.
Let Ψ0 H0and Ψthe associated solution for a control u. Then, if we define, e
Ψ(t, ·) := Ψ(t, ·)e1t,
it comes that e
Ψsatisfies
i∂te
Ψ = e
Ψλ1e
Ψu(t)µϕ1,(t, x)(0, T )×(0,1)
e
Ψ(t, 0) = e
Ψ(t, 1) = 0, t (0, T )
e
Ψ(0,·) = Ψ0, x (0,1)
(C.1)
and ke
Ψ(t, .)kH3
(0) =kΨ(t, .)kH3
(0) .
Rapid stabilization of (C.1). Notice that system (C.1) is almost identical to (1.10) except that the spec-
trum of the underlying operator is shifted by λ1. This modifies the state space. Indeed, for every t0,
0 = ℜhΨ(t, .),Φ1(t, .)i=ℜhe
Ψ(t, .), ϕ1i.
Thus, e
Ψ H0. Notice that due to Theorem 1.3, one gets that system (C.1) is exactly controllable in H0.
The rest of the analysis is barely modified. For instance, the Riesz bases have instead the form
B=g12
n
g22
n;n2h12
n
h22
n;n1,
39
where, for example,
g12
n(x) =
+
X
k=1
2λ(λkλ1)(λnλ1)
(λ2+ (λkλn)2)(λ2+ (λk+λn2λ1)2)β1
nhµϕ1, ϕkiϕk(x),
with β1
nchosen such that hg12
n, ϕni= 1n.
This leads to the existence of a feedback law ˜
K (e
Ψ)
(e
Ψ)!such that, for the closed-loop system
(e
Ψ(t, .))
(e
Ψ(t, .))!
X3
(0)
Ceλt
0)
0)
X3
(0)
.
Rapid stabilization of (1.8). Due to the previous relation, e
Ψ(t, .) = Ψ(t, .)e1tit comes that if Ψis the
solution of (1.8) with the feedback law
v(t) = ˜
Kcos(λ1t)(Ψ(t, .)) sin(λ1t)(Ψ(t, .))
sin(λ1t)(Ψ(t, .)) + cos(λ1t)(Ψ(t, .)),
then
kΨ(t, .)kH3
(0) Ceλt kΨ0kH3
(0) .
References
[1] A. Balogh and M. Krsti´c. Infinite dimensional backsteppingstyle feedback transformations for a heat
equation with an arbitrary level of instability. Eur. J. Control, 8(3):165–175, 2002.
[2] K. Beauchard. Local controllability of a 1-D Schr¨odinger equation. J. Math. Pures Appl. (9),
84(7):851–956, 2005.
[3] K. Beauchard and J.-M. Coron. Controllability of a quantum particle in a moving potential well. J.
Funct. Anal., 232(2):328–389, 2006.
[4] K. Beauchard, J.-M. Coron, and H. Teismann. Minimal time for the bilinear control of Schr¨odinger
equations. Systems Control Lett., 71:1–6, 2014.
[5] K. Beauchard and C. Laurent. Local controllability of 1D linear and nonlinear Schr ¨odinger equations
with bilinear control. J. Math. Pures Appl. (9), 94(5):520–554, 2010.
[6] K. Beauchard and C. Laurent. Exact controllability of the 2D Schr¨odinger-Poisson system. Preprint,
2016.
[7] K. Beauchard and M. Mirrahimi. Practical stabilization of a quantum particle in a one-dimensional
infinite square potential well. SIAM J. Control Optim., 48(2):1179–1205, 2009.
[8] K. Beauchard and M. Morancey. Local controllability of 1D Schr¨odinger equations with bilinear
control and minimal time. Math. Control Relat. Fields, 4(2):125–160, 2014.
[9] U. Boscain, M. Caponigro, T. Chambrion, and M. Sigalotti. A weak spectral condition for the con-
trollability of the bilinear Schr¨odinger equation with application to the control of a rotating planar
molecule. Comm. Math. Phys., 311(2):423–455, 2012.
[10] U. Boscain, M. Caponigro, and M. Sigalotti. Multi-input Schr ¨odinger equation: controllability, track-
ing, and application to the quantum angular momentum. preprint, arXiv:1302.4173, 2013.
[11] U. Boscain, T. Chambrion, and M. Sigalotti. On some open questions in bilinear quantum control.
preprint, arXiv:1304.7181, 2013.
40
[12] D. M. Boˇskovi´c, A. Balogh, and M. Krsti´c. Backstepping in infinite dimension for a class of parabolic
distributed parameter systems. Math. Control Signals Systems, 16(1):44–75, 2003.
[13] N. Boussaid, M. Caponigro, and T. Chambrion. Regular propagators of bilinear quantum systems.
preprint, June 2014.
[14] E. Cerpa and J.-M. Coron. Rapid stabilization for a Korteweg-de Vries equation from the left Dirichlet
boundary condition. IEEE Trans. Automat. Control, 58(7):1688–1695, 2013.
[15] E. Cerpa and E. Cr´epeau. Rapid exponential stabilization for a linear Korteweg-de Vries equation.
Discrete Contin. Dyn. Syst. Ser. B, 11(3):655–668, 2009.
[16] T. Chambrion, P. Mason, M. Sigalotti, and U. Boscain. Controllability of the discrete-spectrum
Schr¨odinger equation driven by an external field. Ann. Inst. H. Poincar´
e Anal. Non Lin´
eaire,
26(1):329–349, 2009.
[17] J.-M. Coron. On the small-time local controllability of a quantum particle in a moving one-
dimensional infinite square potential well. C. R. Math. Acad. Sci. Paris, 342(2):103–108,2006.
[18] J.-M. Coron. Control and nonlinearity, volume 136 of Mathematical Surveys and Monographs.
American Mathematical Society, Providence, RI, 2007.
[19] J.-M. Coron. Stabilization of control systems and nonlinearities. In Proceedings of the 8th Inter-
national Congress on Industrial and Applied Mathematics, pages 17–40. Higher Ed. Press, Beijing,
2015.
[20] J.-M. Coron and B. d’Andr´ea Novel. Stabilization of a rotating body beam without damping. IEEE
Trans. Automat. Control, 43(5):608–618, 1998.
[21] J.-M. Coron, L. Hu, and G. Olive. Stabilization and controllability of first-order integro-differential
hyperbolic equations. Preprint, arXiv:1511.01078, 2015.
[22] J.-M. Coron and Q. L¨u. Local rapid stabilization for a Korteweg-de Vries equation with a Neumann
boundary control on the right. J. Math. Pures Appl. (9), 102(6):1080–1120, 2014.
[23] J.-M. Coron and Q. L ¨u. Fredholm transform and local rapid stabilization for a Kuramoto–Sivashinsky
equation. J. Differential Equations, 259(8):3683–3729, 2015.
[24] J.-M. Coron and H.-M. Nguyen. Null controllability and finite time stabilization for the heat equations
with variable coefficients in space in one dimension via backstepping approach. 2015. Preprint, hal-
01228895, version 1.
[25] J.-M. Coron, R. Vazquez, M. Krsti´c, and G. Bastin. Local exponential H2stabilization of a 2×2
quasilinear hyperbolic system using backstepping. SIAM J. Control Optim., 51(3):2005–2035,2013.
[26] B. Dehman, P. G´erard, and G. Lebeau. Stabilization and control for the nonlinear Schr¨odinger equa-
tion on a compact surface. Math. Z., 254(4):729–749, 2006.
[27] I. C. Gohberg and M. G. Krın. Introduction to the theory of linear nonselfadjoint operators. Trans-
lated from the Russian by A. Feinstein. Translations of Mathematical Monographs, Vol. 18. American
Mathematical Society, Providence, R.I., 1969.
[28] V. Komornik. Rapid boundary stabilization of linear distributed systems. SIAM J. Control Optim.,
35(5):1591–1613, 1997.
[29] M. Krsti´c, B.-Z. Guo, and A. Smyshlyaev. Boundary controllers and observers for the linearized
Schr¨odinger equation. SIAM J. Control Optim., 49(4):1479–1497, 2011.
[30] M. Krsti´c, I. Kanellakopoulos, and P.V. Kokotovic. Nonlinear and Adaptive Control Design. Adap-
tative and Learning Systems for Signal Processing, Communications, and Control. John Wiley and
Sons, July 1995. ISBN: 978-0-471-12732-1.
41
[31] M. Krsti´c and A. Smyshlyaev. Boundary control of PDEs, volume 16 of Advances in Design and
Control. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2008. A course
on backstepping designs.
[32] I. Lasiecka, R. Triggiani, and X. Zhang. Global uniqueness, observability and stabilization of non-
conservative Schr¨odinger equations via pointwise Carleman estimates. II. L2(Ω)-estimates. J. Inverse
Ill-Posed Probl., 12(2):183–231, 2004.
[33] C. Laurent. Internal control of the Schr¨odinger equation. Math. Control Relat. Fields, 4(2):161–186,
2014.
[34] W.-J. Liu and M. Krsti ´c. Backstepping boundary control of Burgers’ equation with actuator dynamics.
Systems Control Lett., 41(4):291–303, 2000.
[35] E. Machtyngier and E. Zuazua. Stabilization of the Schr¨odinger equation. Portugal. Math.,
51(2):243–256, 1994.
[36] F. ehats, Y. Privat, and M. Sigalotti. On the controllability of quantum transport in an electronic
nanostructure. SIAM J. Appl. Math., 74(6):1870–1894, 2014.
[37] M. Mirrahimi. Lyapunov control of a quantum particle in a decaying potential. Ann. Inst. H. Poincar´
e
Anal. Non Lin´
eaire, 26(5):1743–1765, 2009.
[38] M. Morancey. Simultaneous local exact controllability of 1D bilinear Schr¨odinger equations. Ann.
Inst. H. Poincar´
e Anal. Non Lin´
eaire, 31(3):501–529, 2014.
[39] M. Morancey and V. Nersesyan. Simultaneous global exact controllability of an arbitrary number of
1d bilinear Schr ¨odinger equations. J. Math. Pures Appl. (9), 103(1):228–254, 2015.
[40] V. Nersesyan. Growth of Sobolev norms and controllability of the Schr¨odinger equation. Comm.
Math. Phys., 290(1):371–387, 2009.
[41] V. Nersesyan. Global approximate controllability for Schr¨odinger equation in higher Sobolev norms
and applications. Ann. Inst. H. Poincar´
e Anal. Non Lin´
eaire, 27(3):901–915, 2010.
[42] V. Nersesyan and H. Nersisyan. Global exact controllability in infinite time of Schr¨odinger equation.
J. Math. Pures Appl. (9), 97(4):295–317, 2012.
[43] A. Pazy. Semigroups of linear operators and applications to partial differential equations, volume 44
of Applied Mathematical Sciences. Springer-Verlag, New York, 1983.
[44] J.-P. Puel. A regularity property for Schr ¨odinger equations on bounded domains. Rev. Mat. Complut.,
26(1):183–192, 2013.
[45] J.-P. Puel. Local exact bilinear control of the Schr¨odinger equation. Preprint, 2016.
[46] E.D. Sontag. Mathematical control theory, volume 6 of Texts in Applied Mathematics. Springer-
Verlag, New York, second edition, 1998. Deterministic finite-dimensional systems.
[47] B. Sz.-Nagy. Perturbations des transformations lin´eaires ferm´ees. Acta Sci. Math. Szeged, 14:125–
137, 1951.
[48] M. Tucsnak and G. Weiss. Observation and control for operator semigroups. Birkh¨auser Advanced
Texts: Basler Lehrb¨ucher. Birkh¨auser Verlag, Basel, 2009.
[49] J. M. Urquiza. Rapid exponential feedback stabilization with unbounded control operators. SIAM J.
Control Optim., 43(6):2233–2244 (electronic), 2005.
[50] A. Vest. Rapid stabilization in a semigroup framework. SIAM J. Control Optim., 51(5):4169–4188,
2013.
[51] R.M. Young. An introduction to nonharmonic Fourier series, volume 93 of Pure and Applied Math-
ematics. Academic Press Inc. [Harcourt Brace Jovanovich Publishers], New York, 1980.
42
... As in previous works, see, e.g., [6,8,18], we are interested in the solutions in the space H (for each time t) defined by (1.9) H " ! Ψ P H 1 0 pI; Cq; ÿ kě1 |k 3 xΨ, ϕ k y L 2 pIq | 2 ă`8 ) , 1 Hereafter, given a Hilbert space H, we denote x¨,¨yH its scalar product. ...
... y " py 1 , y 2 q T P H such that xy 1 , ϕ 1 y L 2 pIq " 0 ) , and we equip with the following scalar products for the spaces H and H 1,7 : (1. 18) xy, r yy H :" xy, r yy H 3 pIq " ż I 2 ÿ ℓ"1´y ℓ r y ℓ`y 1 ℓ r y 1 ℓ`y 2 ℓ r y 2 ℓ`y 3 ℓ r y 3 ℓ¯d s for y " py 1 , y 2 q T , r y " pr y 1 , r y 2 q T P H, and (1. 19) xy, r yy H 1,7 " xy, r yy H for y, r y P H 1,7 . ...
... One can check, see e.g., [18], that (1.20) H " ! y " py 1 , y 2 q T P H 3 pI; R 2 q; y 1 pxq " y 2 pxq " y 2 1 pxq " y 2 2 pxq " 0 on BI ) and (1.21) H 1,7 " ! ...
Preprint
Full-text available
We propose a method to establish the rapid stabilization of the bilinear Schr\"odinger control system and its linearized system, and the finite time stabilization of the linearized system using the Grammian operators. The analysis of the rapid stabilization involves a new quantity (variable) which is inspired by the adjoint state in the optimal control theory and is proposed in our recent work on control systems associated with strongly continuous group. The analysis of the finite time stabilization follows the strategy introduced by Coron and Nguyen in the study of the finite time stabilization of the heat equation and incorporate a new ingredient involving the estimate of the cost of controls of the linearized system in small time derived in this paper.
... 93C20, 93B17, 93D15, 33C10. in the present paper, using a Volterra transformation would lead to solve a wave equation in two dimensions, that is degenerate, on a triangular domain, where both the functional setting and the qualitative properties of the equation are unclear 1 2 PIERRE LISSY AND CLAUDIA MORENO the transformation is more difficult to obtain. Here, we will follow the abstract approach developed in [11]. ...
... To study the stabilization of the system (1), as already mentioned, we will use the backstepping method. We focus on the construction of an appropriate Fredholm transformation, in the spirit of [11], which has emerged as an alternative to the Volterra transformation during the last years (see also [12,13,28,16,17,15]). Let us emphasize that our main stabilization result can be obtained from the abstract theorem given in [1, Theorem 1.6], but the backstepping method has some additional advantages (see [16, Remark 2]) that make relevant the present study. ...
... Then, one can prove the following result (see [27,Theorem 15] and [11,Theorem 3.3]. ...
... In the present work, the control action takes the form of a bilinear control, that is, a control given by the multiplicative coefficient α in (1). General references in the area of multiplicative controllability are the seminal work [4] by Ball, Marsden, and Slemrod, some important results about bilinear control of the Schrödinger equation obtained by Beauchard, Coron, Gagnon, Laurent and Morancey in [5], [20], and in the references therein, and some results obtained by Khapalov for parabolic and hyperbolic equations, see [33], and the references therein. See also some results for reaction-diffusion equations (both degenerate and uniformly parabolic) obtained by Cannarsa and Floridia in [9] and [10], by Floridia in [29], and by Cannarsa, Floridia and Khapalov in [12]. ...
Preprint
In this paper we study the global approximate multiplicative controllability for nonlinear degenerate parabolic Cauchy problems. In particular, we consider a one-dimensional semilinear degenerate reaction-diffusion equation in divergence form governed via the coefficient of the \-reaction term (bilinear or multiplicative control). The above one-dimensional equation is degenerate since the diffusion coefficient is positive on the interior of the spatial domain and vanishes at the boundary points. Furthermore, two different kinds of degenerate diffusion coefficient are distinguished and studied in this paper: the weakly degenerate case, that is, if the reciprocal of the diffusion coefficient is summable, and the strongly degenerate case, that is, if that reciprocal isn't summable. In our main result we show that the above systems can be steered from an initial continuous state that admits a finite number of points of sign change to a target state with the same number of changes of sign in the same order. Our method uses a recent technique introduced for uniformly parabolic equations employing the shifting of the points of sign change by making use of a finite sequence of initial-value pure diffusion pro\-blems. Our interest in degenerate reaction-diffusion equations is motivated by the study of some \-energy balance models in climatology (see, e.g., the Budyko-Sellers model) and some models in population genetics (see, e.g., the Fleming-Viot model).
... The stabilization problem has attracted many scholars attention, and it has been widely studied in many kinds of partial differential equations, for instance, Navier-Stokes equation [9], Schrödinger equation [15,27,35], Burgers equation [16], Korteweg-de Vries equation [22]. Zhang-Zhu-Li-Wang considered the stabilization of an unstable wave equation through a dynamic boundary compensator in [48]. ...
Article
Full-text available
This paper concerns with the stabilizability for a quasilinear Klein–Gordon–Schrödinger system with variable coefficients in dimensionless form. The stabilizability of quaslinear Klein–Gordon-Wave system with the Kelvin–Voigt damping has been considered by Liu–Yan–Zhang (SIAM J Control Optim 61:1651–1678, 2023). Our main contribution is to find a suitable linear feedback control law such that the quasilinear Klein–Gordon–Schrödinger system is exponentially stable under certain smallness conditions.
... where g is a given profile and U is the scalar control. Rapid exponential stabilization by feedback for the linearized Schrodinger equation has been studied in [63] by using interior control. In the paper [64], the author explored backstepping approach for the one dimensional transport equation. ...
Article
Full-text available
In this article, we study the boundary feedback stabilization of some one-dimensional nonlinear coupled parabolic-ODE systems, namely Rogers-McCulloch and FitzHugh-Nagumo systems, in the interval (0, 1). Our goal is to construct an explicit linear feedback control law acting only at the right end of the Dirichlet boundary to establish the local exponential stabilizability of these two different nonlinear systems with a decay e^ {-ωt} , where ω ∈ (0, δ] for the FitzHugh-Nagumo system and ω ∈ (0, δ) for the Rogers-McCulloch system and δ is the system parameter that presents in the ODE of both coupled systems. The feedback control law, derived by the backstepping method forces the exponential decay of solution of the closed-loop nonlinear system in both L^2 (0, 1) and H^1 (0, 1) norms, respectively, if the initial data is small enough. We also show that the linearized FitzHugh-Nagumo system is not stabilizable with exponential decay e^ {-ωt} , where ω > δ.
... This method, first introduced by Krstic and his collaborators [28], corresponds to moving the spectrum with the help of some feedback laws. It has been generalized in [29,30] by Fredholm transformation, and has turned out to be efficient for various one-dimensional models [31][32][33][34][35]. However, it is still a challenging open problem to introduce the backstepping method on general multidimensional models. ...
Article
Full-text available
The finite time stabilizability of the one dimensional heat equation is proved by Coron–Nguyên [16], while the same question for multidimensional spaces remained open. Inspired by Coron–Trélat [17] we introduce a new method to stabilize multidimensional heat equations quantitatively in finite time and call it Frequency Lyapunov method. This method naturally combines spectral inequality [35] and constructive feedback stabilization. As application this approach also yields a constructive proof for null controllability, which gives sharing optimal cost with explicit controls and works perfectly for related nonlinear models such as Navier–Stokes equations [52].
... The above considerations motivate our investigation of the multiplicative controllability. As regards the topic of multiplicative controllability of PDEs we recall the pioneering paper [5] by Ball, Marsden, and Slemrod, and in the framework of the Schrödinger equation we especially mention [6] by Beauchard and Laurent,and [22] by Coron, Gagnon and Moranceu. As regards parabolic and hyperbolic equations we focus on some results by Khapalov contained in the book [41], and in the references therein. ...
Article
Full-text available
We consider a nonlinear degenerate reaction-diffusion equation. First we prove that if the initial state is nonnegative, then the solution remains nonnegative for all time. Then we prove the approximate controllability between nonnegative states via multiplicative controls, this is done using the reaction coefficient as control. For more information see https://ejde.math.txstate.edu/Volumes/2020/59/abstr.html
... Verification of the hypotheses. It is well-known that a fundamental solution of the Schrödinger system can be obtained by the Fourier expansion, see for instance [19,25]. So, in this way, considering the eigenvalues and the eigenfunctions that form an orthonormal basis of L 2 (Ω), we can define in H an inner product similarly as in [16]. ...
Article
Full-text available
This article presents some controllability and stabilization results for a system of two coupled linear Schrödinger equations in the one-dimensional case where the state components are interacting through the Kirchhoff boundary conditions. Considering the system in a bounded domain, the null boundary controllability result is shown. The result is achieved thanks to a new Carleman estimate, which ensures a boundary observation. Additionally, this boundary observation together with some trace estimates, helps us to use the Gramian approach, with a suitable choice of feedback law, to prove that the system under consideration decays exponentially to zero at least as fast as the function e-2ωte2ωte^{-2\omega t} for some ω>0ω>0\omega >0.
Article
This paper deals with global stability dynamics for the Klein–Gordon–Zakharov system in R 2 . We first establish that this system admits a family of linear mode unstable explicit quasi-periodic wave solutions. Next, we prove that the Kelvin–Voigt damping can help to stabilize those linear mode unstable explicit quasi-periodic wave solutions for the Klein–Gordon–Zakharov system in the Sobolev space H s + 1 ( R 2 ) × H s + 1 ( R 2 ) × H s + 1 ( R 2 ) for any s ⩾ 1. Moreover, the Kelvin–Voigt damped Klein–Gordon–Zakharov system admits a unique Sobolev regular solution exponentially convergent to some special solutions (including quasi-periodic wave solutions) of it. Our result can be extended to the n-dimension dissipative Klein–Gordon–Zakharov system for any n ⩾ 1.
Article
Full-text available
Using the backstepping approach we recover the null controllability for the heat equations with variable coefficients in space in one dimension and prove that these equations can be stabilized in finite time by means of periodic time-varying feedback laws. To this end, on the one hand, we provide a new proof of the well-posedness and the “optimal” bound with respect to damping constants for the solutions of the kernel equations; this allows one to deal with variable coefficients, even with a weak regularity of these coefficients. On the other hand, we establish the well-posedness and estimates for the heat equations with a nonlocal boundary condition at one side.
Article
In this article, we investigate the exact controllability of the 2D-Schrödinger-Poisson system, which couples a Schrödinger equation on a bounded domain of R2\mathbb{R}^2 with a Poisson equation for the electrical potential. The control acts on the system through a Neumann boundary condition on the potential, locally distributed on the boundary of the space domain. We prove several results, with or without nonlinearity and with different boundary conditions on the wave function, of Dirichlet type or of Neumann type.
Article
We are going to prove the local exact bilinear controllability for a Schr " odinger equation, set in a bounded regular domain, in a neighborhood of an eigenfunction corresponding to a simple eigenvalue in dimension N <= 3. For a general domain we will require a non degeneracy condition of the normal derivative of the eigenfunction on a part Gamma(0) of the boundary satisfying the Geometric Control Condition (see [G. Lebeau. J. Math. Pures Appl. 71 (1992) 267-291]) and for a rectangle when N = 2 or an interval for N = 1 no further condition. In the general case we will use real potentials concentrated in the neighborhood of Gamma(0) and the linear controllability results with real and sufficiently regular controls.
Article
In this work, we consider the problem of boundary stabilization for a quasilinear 2×2 system of first-order hyperbolic PDEs. We design a new full-state feedback control law, with actuation on only one end of the domain, which achieves H2 exponential stability of the closed-loop system. Our proof uses a backstepping transformation to find new variables for which a strict Lyapunov function can be constructed. The kernels of the transformation are found to verify a Goursat-type 4×4 system of first-order hyperbolic PDEs, whose well-posedness is shown using the method of characteristics and successive approximations. Once the kernels are computed, the stabilizing feedback law can be explicitly constructed from them.
Article
In the present article we study the stabilization of first-order linear integro-differential hyperbolic equations. For such equations we prove that the stabilization in finite time is equivalent to the exact controllability property. The proof relies on a Fredholm transformation that maps the original system into a finite-time stable target system. The controllability assumption is used to prove the invertibility of such a transformation. Finally, using the method of moments, we show in a particular case that the controllability is reduced to the criterion of Fattorini.
Article
This paper is devoted to the study of the local rapid exponential stabilization problem for a controlled Kuramoto-Sivashinsky equation on a bounded interval. We build a feedback control law to force the solution of the closed-loop system to decay exponentially to zero with arbitrarily prescribed decay rates, provided that the initial datum is small enough. Our approach uses a method we introduced for the rapid stabilization of a Korteweg-de Vries equation. It relies on the construction of a suitable integral transform and can be applied to many other equations.