• Aucun résultat trouvé

Carleman estimates for time-discrete parabolic equations and applications to controllability

N/A
N/A
Protected

Academic year: 2021

Partager "Carleman estimates for time-discrete parabolic equations and applications to controllability"

Copied!
36
0
0

Texte intégral

(1)

HAL Id: hal-02121672

https://hal.archives-ouvertes.fr/hal-02121672

Submitted on 6 May 2019

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Carleman estimates for time-discrete parabolic equations and applications to controllability

Franck Boyer, Víctor Hernández-Santamaría

To cite this version:

Franck Boyer, Víctor Hernández-Santamaría. Carleman estimates for time-discrete parabolic equa- tions and applications to controllability. ESAIM: Control, Optimisation and Calculus of Variations, EDP Sciences, 2020, 26, �10.1051/cocv/2019072�. �hal-02121672�

(2)

Carleman estimates for time-discrete parabolic equations and applications to controllability

Franck Boyer V´ıctor Hern´andez-Santamar´ıa May 6, 2019

Abstract

In this paper, we prove a Carleman estimate for a time-discrete parabolic equation in which the large parameter is connected to the time step of the discretization scheme. This estimate is then used to obtain a relaxed observability estimate that yields some controllability results for linear and semi-linear time-discrete parabolic equations. We also discuss the application of this Carleman estimate to the controllability of coupled systems.

Keywords: Carleman estimates, time-discrete heat equation, observability, null controllability.

1 Introduction

1.1 Null-controllability of the heat equation

Let ΩRd,d1 be a bounded open set with boundary∂Ω regular enough. For given timeT >0, we denoteQ= Ω×(0, T) and Σ =∂Ω×(0, T). Let ωbe a nonempty subset of Ω. We consider the linear heat equation

ty∆y=1ωv inQ,

y= 0 on Σ,

y(x,0) =y0(x) in Ω.

(1.1) In (1.1),y=y(x, t) is the state,v=v(x, t) is the control function acting on the system on the control domainω, andy0 is a given initial data. As usual,1ω denotes the characteristic function ofω.

It is well-known that for anyy0 L2(Ω) andvL2×(0, T)), the corresponding solution to (1.1) is globally defined in [0, T]. More precisely, one has

yC([0, T];L2(Ω))L2(0, T;H01(Ω)). (1.2) One of the key questions in control theory is to determine whether a system enjoys the so-callednull controllability property. System (1.1) is said to be null-controllable at timeT if, for any y0 L2(Ω), there exists a controlvL2×(0, T)) such that the corresponding solution satisfies

y(·, T) = 0 in Ω.

Observe that the regularity (1.2) justifies the definition we have introduced.

It is by now well-known that (1.1) is null-controllable for anyT >0 and for any nonempty open set ωΩ. In fact, this problem was addressed independently in the 90’s by Lebeau & Robbiano [15] and by Fursikov & Imanuvilov [12]. By a duality argument, the null-controllability of (1.1) is equivalent to the observability of the associated adjoint states. More precisely, for eachqT L2(Ω), consider the adjoint system

−∂tq∆q= 0 inQ,

q= 0 on Σ,

q(x, T) =qT(x) in Ω.

(1.3)

(3)

Then, (1.1) is null-controllable if and only if there existsCobs>0 such that the following observability inequality holds

|q(0)|L2(Ω)Cobs

Z Z

ω×(0,T)

|q|2dxdt

!12

, ∀qT L2(Ω). (1.4)

1.2 Time-discrete setting

In this paper, we shall use the notationJa, bK= [a, b]N, for any real numbersa < b.

We are interested in studying controllability and observability properties for the time-discrete coun- terparts to (1.1) and (1.3), but also for more general parabolic systems (see Sections 4and 5). To be more precise, for any givenMN, we set4t=T /M and introduce the following discretization for the time variable

0 =t0< t1< . . . < tM =T, (1.5) withtn=n4tandnJ0, MK. We also introducetn+1

2 = (tn+1+tn)/2, fornJ0, M1K, see Figure1.

For any time discrete control sequencev ={vn+12}n∈J0,M−1K L2(Ω), consider the sequence y = {yn}n∈J0,MKL2(Ω) verifying the recursive formula

yn+1yn

4t ∆yn+1=1ωvn+12, nJ0, M1K, yn+1|∂Ω = 0, nJ0, M1K, y0=y0.

(1.6)

where yn (resp. vn+12) denotes an approximation ofy(resp. v) at timetn(resp. tn+1

2). Observe that (1.6) is precisely an implicit Euler discretization of the heat equation. Evidently, there exist many other methods to discretize (1.1), but we have chosen this method for the sake of simplicity.

As in the continuous case, we can introduce the notion of controllability for the discrete scheme, this is, system (1.6) is said to be null controllable if for anyy0L2(Ω) there exists a sequence{vn+12}n∈J0,M−1K

such that the corresponding solution satisfies

yM = 0. (1.7)

Notice that for any fixed4tand for eachnJ0, M1K, (1.6) can be regarded as a system of controlled elliptic equations.

There are only a handful of works in the literature addressing the controllability of time-discrete systems such as (1.6). This may come from the fact that, as pointed out in [22, Theorem 1.1], system (1.6) is not even approximately controllable for any given 4t > 0, except for the trivial case when ω = Ω. In view of this negative result, it is natural to ask whether the controllability requirement (1.7) for system (1.6) can be relaxed. In this direction, in [22], the controllability for (1.6) is achieved by considering the projections of solutions over a suitable class of low frequency Fourier components. Using a time-discretized Lebeau-Robbiano strategy (see e.g. [15,16]), the author proves a uniform controllability result (with respect to4t) for the low frequency part of the solution.

In [7], the authors prove in a quite general framework that any controllable parabolic equation (even if discretized in the space variable) is null controllable after time discretization by an appropriate filtering of the high-frequencies. In [5], the authors study in a general setting the null-controllability of fully- discrete approximations for parabolic equations and its convergence rate towards the semi-discrete (in space) case. There, the authors prove somerelaxed observability estimates (uniform with respect to the discretization parameters) allowing to recover classical results for the continuous case. However, both of these works rely on spectral analysis tools and therefore the results are limited to autonomous linear control systems. We finally mention the works [8, 21,23] which encompass the controllability of time discretization schemes for wave-like, Schr¨odinger and KdV equations.

Here, inspired in the strategy outlined in [6] (see also [4]) for the space discretization of parabolic equations, we derivate a Carleman estimate for time-discrete approximations of the parabolic operator

−∂t∆ and from there we deduce a relaxed observability inequality for a suitable adjoint system. We shall refer to it as a relaxed inequality due to the presence of an extra term on the right-hand side (as

(4)

0 =t0 t1 t2 tM−1 tM =T t1

2 t3

2 tM−1

2 tM+1

2 P = (tn)n∈J0,M

K

D= (tn+1 2)n∈

J0,MK

Figure 1: Discretization of the time variable and its notation.

compared to (1.4)). This inequality, in turn, allows to obtain by duality a controllability result where a small target is reached and whose size goes to zero exponentially as4t0, more precisely, we achieve

|yM|L2(Ω)Cobs

pφ(4t)|y0|L2(Ω),

where Cobs is a positive constant uniform with respect to 4tand 4t7→ φ(4t) is a suitable function of the discretization parameter 4t. For this reason, in the spirit of [6], we shall speak of aφ(4t)-null controllability result.

The main goal and novelty of our approach are twofold. By deriving directly a Carleman estimate for the time-discrete operator, we can deduce controllability results for more general systems (e.g., equations with time-dependent coefficients, right-hand side terms, etc) since we are no longer restricted to spectral techniques (as in [7,5,22]). Moreover, our methodology can be readily adapted to derive the analogous counterpart of well-known controllability results in the continuous case, commonly relying on Carleman inequalities, such as the cases of semilinear systems, coupled equations, or non-standard problems as the insensitizing control.

1.3 Notations and functional framework

Before introducing our main results, we establish the framework of the discrete setting we shall work with. The notation introduced here allow us to use a formalism as close as possible to the continuous case and, in this way, most of the computations will be carried out in a very intuitive manner.

From the discretization points (1.5), we will denote by P := {tn:nJ1, MK} the (primal) set of points excluding the first one and we writeP:=P ∪ {t0}. To handle in an efficient way computations related to the approximation of the time derivatives, we will naturally work on another (dual) set of points lying at the middle points of P. More precisely, we defineD:={tn+1

2 :nJ0, M1K}. It will be convenient to consider also an extra point tM+1

2 which lies outside the interval [0, T] (see Figure1) and to writeD:=D ∪ {TM+1

2}. Observe that bothPandDhave a total number ofM elements.

We denote byRP andRDthe sets of real-valued discrete functions defined onP andD. IfuPRP (resp. uDRD), we denote byun(resp. un+12) its value corresponding totn(resp.tn+1

2). ForuPRP we define the time-discrete integral

Z T 0

uP:=

M

X

n=1

4t un, (1.8)

and, analogously, foruDRDwe define

Z T

0

uD:=

M−1

X

n=0

4t un+12. (1.9)

Remark 1.1. To ease the notation and thanks to the introduction of two different integral symbols, in what follows we shall writeuindistinctly to refer to functionsuP oruD.

Let{X,| · |X}be a real Banach space. We denote byXP andXDthe sets of vector-valued functions defined on P and D, respectively. Using the definitions (1.8) and (1.9) for the discrete integrals, we denote by LpP(0, T;X) (resp. LpD(0, T;X)), 1 p <∞, the space XP (resp. XD) endowed with the norm

kukLp

P(0,T;X):=

Z T 0

|u|pX 1/p

resp. kukLp

D(0,T;X):=

Z T

0

|u|pX 1/p!

.

(5)

We also define the spaceLP(0, T;X) (resp. LD(0, T;X)) by means of the norm kukLP(0,T;X):= sup

tn∈P

|un|X

resp. kukLD(0,T;X):= sup

tn+ 1 2

∈D

|un+12|X

.

In the case wherep= 2 andX is replaced by a Hilbert space{H,(·,·)H},HP (resp. HD) becomes a Hilbert space for the norm induced by the inner product

ZT 0

(u, v)H:=

M

X

n=1

4t(un, vn)H resp. Z T

0

(u, v)H :=

M−1

X

n=0

4t(un+12, vn+12)H

! . Particularly, ifH=L2(Ω) we shall use the notation

Z Z

Q

uv:=

Z T 0

(u, v)L2(Ω)

resp. Z Z

Q

uv:= — ZT

0

(u, v)L2(Ω)

.

Remark 1.2. For short and in accordance with the notation used in the continuous case, we will denote the spaces L2P(0, T;L2(Ω)) as L2P(Q) and we use LP(Q) to indicate the space LP(0, T;L(Ω)). The same notation holds for functions defined on the dual grid D.

To manipulate time-discrete functions, we define translation operators for indices t+ : XP XD andt:XPXDas follows:

(t+u)n+12 :=un+1, (tu)n+12 :=un, nJ0, M1K.

With this, we can define a difference operatorDt as the map fromXP intoXDgiven by (Dtu)n+12 :=un+1un

4t = 1

4t t+t u

n+12

, nJ0, M1K.

In the same manner, we can define the translation operators ¯t+:XDXP and ¯t:XDXP as follows:

t+u)n:=un+12, tu)n=un−12, nJ1, MK, (1.10) as well as a difference operatorDt (mappingXDintoXP) given by

(Dtu)n:=un+12 un−12

4t =

1

4t ¯t+¯t u

n

, nJ1, MK.

These definitions, together with the integral symbols (1.8) and (1.9), allow us to obtain a series of results for handling in quite natural fashion the application of the derivativesDtandDtto functions either continuously defined or discrete. For convenience, we have summarized in Appendix Bthe main tools and estimates used along this document. As an example, for functionsu[L2(Ω)]P andv[L2(Ω)]D, we have the following useful formula

Z Z

Q

(Dtu)v=−(u0, v12)L2(Ω)+ (uM, vM+12)L2(Ω) Z Z

Q

(Dtv)u, (1.11)

which resembles classical integration by parts. Expressions like (1.11) allow us to present and perform computations intuitively, facilitating also the presentation and reading of this paper.

1.4 Statement of the main results

1.4.1 Carleman estimate

Let us introduce several weight functions that will be useful in the remainder of this paper. We introduce a special function whose existence is guaranteed by the following result [12, Lemma 1.1].

(6)

Lemma 1.3. LetB0⊂⊂be a nonempty open subset. Then, there existsψC2(Ω)such that (ψ(x)>0 allxΩ, ψ|∂Ω= 0,

|∇ψ|>0 for allxΩ\B0. LetK >kψkC(Ω) and set

ϕ(x) =eλψ(x)eλK<0, φ(x) =eλψ(x), (1.12) and

θ(t) = 1

(t+δT)(T+δTt). (1.13)

for some 0< δ <1/2. The parameterδis introduced to avoid singularities at timet= 0 andt=T (see Remark1.5below for further comments).

We state our first result, a uniform Carleman estimate for the time-discrete backward parabolic operator formally defined on the dual grid as follows

(LDq)n:=−(Dtq)n∆(¯tq)n, nJ1, MK, (1.14) for anyq(H2(Ω))D. The result is the following.

Theorem 1.4. LetB0be a nonempty open set ofand a functionψprovided by Lemma1.3and define ϕ according to (1.12). LetB another open subset of such that B0 ⊂⊂ B. For the parameterλ 1 sufficiently large, there existC,τ01,ε0>0, depending onB,B0,T andλsuch that

τ−1 Z Z

Q

¯t(e2τ θϕθ−1)

|Dtq|2+

∆(¯tq)

2 +τ

Z Z

Q

¯t(e2τ θϕθ)|∇(¯tq)|2+τ3 Z Z

Q

¯t(e2τ θϕθ3)(¯tq)2

C Z Z

Q

te2τ θϕ)|LDq|2+τ3 Z Z

B×(0,T)

¯t(e2τ θϕθ3)(¯tq)2

!

+C(4t)−1 Z

(eτ θϕq)12

2

+ Z

(eτ θϕq)M+12

2

+ Z

(eτ θϕ∇q)M+12

2

, (1.15)

for allτ τ0(T+T2), and for all4t >0and0< δ1/2satisfying the condition τ44t δ4min{T3, T6}−1

ε0, andq is any time discrete function in(H2(Ω)H01(Ω))D.

To prove the Carleman estimate, we proceed as close as possible to the continuous case and follow the procedure presented in [10]. During the proof, we will clarify the main differences and difficulties introduced by time discretization.

Remark 1.5. Some remarks are in order.

One can readily recognize from (1.15)the classical structure of a Carleman inequality in the contin- uous setting (cf. [10, Lemma 1.3]). The last three terms are, however, specific to the discrete case and arise during the proof. In fact, the presence of these terms is important: otherwise we could obtain a classical observability inequality (not relaxed) leading to a uniform controllability result (w.r.t4t) for (1.6), which would be a contradiction with the known results (see [22]).

Note that, despite the presence of a(4t)−1 factor, those three terms are actually exponentially small with respect to4tsinceϕ <0andθ is large (close to δT12) neart= 0 andt=T.

As compared to other discrete (in space) Carleman results (see [6]), the last term corresponding to the gradient of the functionqcannot be avoided. Actually, this term also appears during the proof of the estimate presented in [6, Theorem 1.3] but since the functions are discrete in the space variable, it is approximated as ∇qCh−1q, where his the mesh step size. In our case, it is not clear how to remove this term and how to prove estimate (1.15) without the last term remains open.

(7)

Even though it does not explicitly appear in our Carleman estimate since it is hidden in the definition (1.13) of θ, the parameterδ plays a key role in the proof. As mentioned before, the main interest is to avoid the singularity of the weight θ(t) at times t= 0andt=T (these singularities, which correspond to the case δ= 0, are systematically exploited in the continuous setting, see e.g. [12], but are rather difficult to handle in the discrete framework). Here, by takingδ >0, we enable two different things: one one hand, we can define continuously the weight outside the time interval[0, T], since functions on the dual mesh Dhave one extra point lying outside this interval (see Figure1).

On the other, it allow us to estimate the derivative ofθ (see LemmaB.4) and set a suitable change of variables (see eq. (2.1)), which is the starting point of the proof.

If one considers instead ofLDthe following forward-in-time operator ( ˜LDq)n:= (Dtq)n∆(¯t+q)n, nJ1, MK,

for anyq(H2(Ω)H01(Ω))D, then a similar Carleman estimate can be obtained: it is just needed to replace all the ¯toperators by¯t+ and the last term by

Z

(eτ θϕ∇q)12

2

.

Using the tools in Appendix B, Theorem 1.4 can be easily adapted to discrete parabolic operators acting on primal variables. For instance, we can consider the forward-in-time parabolic operator defined as follows

(LPy)n+12 = (Dty)n+12∆(t+y)n+12, nJ0, M1K,

for ally(H2(Ω)H01(Ω))P. Then, under the same conditions of Theorem1.4, we can prove the following estimate

τ−1 Z Z

Q

t+(e2τ θϕθ−1)

|Dty|2+

∆(t+y)

2 +τ

Z Z

Q

t+(e2τ θϕθ)|∇(t+y)|2+τ3 Z Z

Q

t+(e2τ θϕθ3)(t+y)2

C Z Z

Q

(t+e2τ θϕ)|LPy|2+τ3 Z Z

B×(0,T)

t+(e2τ θϕθ3)(t+y)2

!

+C(4t)−1 Z

(eτ θϕy)0

2

+ Z

(eτ θϕy)M

2

+ Z

(eτ θϕ∇y)0

2

, (1.16)

for any time discrete functiony(H2(Ω)H01(Ω))P.

As in the previous remark, we can adapt the result to the associated backward-in-time operator ( ˜LPy)n+12 =−(Dty)n+12 ∆(ty)n+12, nJ0, M1K.

1.4.2 Controllability results: the linear case

By considering a standard implicit Euler scheme for the time variable, the time-discrete homogeneous heat equation with potentialaLP(Q) reads as follows

yn+1yn

4t ∆yn+1+an+1yn+1= 0, nJ0, M1K,

yn+1|∂Ω = 0, nJ0, M1K,

y0=y0.

(1.17)

With the notation introduced previously, we can rewrite system (1.17) as (

(Dty)n+12 ∆(t+y)n+12 + (t+ay)n+12 = 0, nJ0, M1K, y0=y0.

where, for convenience, we shall not explicitly write the homogeneous Dirichlet boundary conditions in such compact formulas since we will not deal with other boundary conditions in this paper.

(8)

Observe that the equation verified by y is written on the dual grid D. This motivates us to look for a time-discrete controlvwhich is naturally attached to this grid (cf. equation (1.6)). Thus, we will consider controlled systems of the form

((Dty)n+12∆(t+y)n+12 + (t+ay)n+12 =1ωvn+12, nJ0, M1K,

y0=y0. (1.18)

Following the well-known Hilbert Uniqueness Method (see Proposition3.3), we can build a control function by minimizing a quadratic functional defined for the solutions to the adjoint of (1.18), which in this case is given by

qn−12 qn+12

4t ∆qn−12 +anqn−12 = 0, nJ1, MK, qn−

1 2

|∂Ω = 0, nJ1, MK,

qM+12 =qT.

(1.19)

System (1.19) is the proper adjoint system of (1.18). With our notation, we can rewrite (1.19) in a more compact way, namely,

(−(Dtq)n∆(¯tq)n+antq)n= 0, nJ1, MK,

qM+12 =qT. (1.20)

Applying the Carleman inequality (1.15) to (1.20) and after a series of steps, we will deduce an observability inequality of the form

|q12|L2(Ω)Cobs Z Z

ω×(0,T)

|q|2+e

C2

(4t)1/4|qT|2H1 0(Ω)

!12

, (1.21)

for some positive constantsCobsandC2 only depending onT,ωandkak.

As mentioned before, this inequality has an extra term in the right-hand side as compared with the similar estimate in the continuous setting (1.4). This extra term is exponentially small which is actually an improvement compared to the similar inequality proved in [7].

With this inequality we are able to prove that there exists a discrete controlvL2D(0, T;L2(ω)) with kvkL2

D(0,T;L2(ω))C, uniformly with respect to4t, such that the associated solution of (1.18) satisfies

|yM|H−1(Ω)Cp

φ(4t)|y0|L2(Ω) (1.22)

where4t7→φ(4t) is any given function of the discretization parameter such that lim inf

4t→0

φ(4t)

e−C2/(4t)1/4 >0. (1.23)

Actually, this means that we reach a small targetyM whose size goes to zero as the time step4t0, at the prescribed rate p

φ(4t) with controls that remain uniformly bounded with respect to 4t. In practice, computing such a control can be done by a minimization algorithm for which one can choose the functionφin such a way that the errorp

φ(4t) on the target is comparable to the actual accuracy of the discretization scheme. We refer to [2] for a more complete discussion on that point.

The precise result is given by the following theorem.

Theorem 1.6. Let us considerT >0and a discretization parameter4tsufficiently small. Then, for any y0L2(Ω)and any functionφverifying (1.23), there exists a time-discrete controlvL2D(0, T;L2(ω)) such that

kvkL2

D(0,T;L2(ω))C|y0|L2(Ω),

and such that the associated solutionyto(1.17)verifies (1.22), for a positive constantC only depending onφ, T andkak as in (3.3).

(9)

Observe that Theorem1.6only yields a controllability result inH−1. This is due to the presence of the term|qT|H1

0 in the estimate (1.21) which in turn comes from the Carleman inequality (1.15) and is closely related to the selection of the weight function (1.12) (see Remarks1.5and2.1). Using classical parabolic regularity results, we can actually obtain a controllability result in aL2-setting.

Theorem 1.7. Let us consider T > 0 and a discretization parameter 4t sufficiently small. Let 0 <

Ce2< C2. Then, for anyy0L2(Ω)and any functionφverifying lim inf

4t→0

φ(4t)

eCe2(4t)1/4 >0, (1.24)

there exists a time-discrete controlvsuch that kvkL2

D(0,T;L2(ω))C|y0|L2(Ω), and the associated solutionyto (1.18)verifies

|yM|L2(Ω)Cp

φ(4t)|y0|L2(Ω), (1.25)

where the positive constantC depends only onφ, T andkak as in (3.3).

The methodology to prove Theorem1.7is to split the time interval in two parts. In the first (large) subinterval, we choosevsuch thatysatisfies (1.22). In the second one, we setv= 0 and let evolve the uncontrolled system. Then, from an elliptic regularity result, we finally deduce (1.25).

1.4.3 Applications to the semilinear case

One of the main advantages of using Carleman estimates for proving controllability results is the pos- sibility of addressing nonlinear problems. Unlike [7] or [22], where spectral properties of the equations are needed (therefore restricted to the linear case), here we will study the controllability of the following implicit Euler scheme

yn+1yn

4t ∆yn+1+f(yn+1) =1ωvn+12, nJ0, M1K,

y|∂Ωn+1= 0, nJ0, M1K,

y0=y0.

(1.26)

wherefC1(R) is a globally Lipschitz function withf(0) = 0. The result is the following.

Theorem 1.8. Let us considerT >0and assume that4tis small enough. Then, for anyy0L2(Ω)and any function verifying (1.23), there exists a uniformly bounded time-discrete controlvL2D(0, T;L2(ω)) such that the associated solution yto(1.26) verifies(1.22).

The proof of Theorem1.8follows other well-known controllability results for nonlinear systems (see, for instance, [9,11,19]). First, we prove the existence of aφ(4t)-null control for a linearized version of (1.26) and then, after a careful analysis on the dependence of the constants appearing in (1.21) and using a fixed point argument, we deduce the result for the nonlinear case.

1.5 Outline

The rest of the paper is organized as follows. In section2we present the proof of Theorem1.4. We have divided the proof in several steps to ease the reading. Section3is divided in two parts: in the first one, by employing the Carleman estimate (1.15), we obtain a relaxed observability inequality and then we use it to deduce theφ(4t)-null controllability result stated in Theorem1.6. In the second one, we prove the L2-null controllability result presented in Theorem1.7. We devote Section4to prove the nonlinear result enunciated in Theorem1.8. Finally, in Section 5we present additional results and a brief discussion on the applicability of the Carleman estimate (1.15) for handling control problems for coupled systems.

Références

Documents relatifs

Dans le cadre de cette thèse, nous avons étudié le comportement mécanique, ou plus précisément le comportement à la fissuration et la plasticité, à température ambiante,

Then we use this Carleman estimate with a special choice of weight functions to obtain in section 3 a three balls theorem and doubling inequalities on solutions of (1.1).. Section 4

Section 4 is devoted to the proof the semi- discrete parabolic Carleman estimate in the case of a discontinuous diffu- sion cofficient for piecewise uniform meshes in

Carleman estimate for elliptic operators with coefficients with jumps at an interface in arbitrary dimension and application to the null controllability of linear parabolic

The Carleman estimate proven in Section 3 allows to give observability estimates that yield null controllability results for classes of semilinear heat equations... We have

From the relaxed observability estimate given above we obtain a h-null controllability result for the linear operator P M.. The equation is linearized yielding a bounded potential and

In this work, we proved a necessary and sufficient condition of approximate and null controllability for system (6) with distributed controls under the geometrical assumption

In this paper we investigate on a new strategy combining the logarithmic convexity (or frequency function) and the Carleman commutator to obtain an observation estimate at one time