Null-controllability of cascade reaction-diffusion systems with odd coupling terms
Abstract
In this paper, we consider a nonlinear system of two parabolic equations, with a distributed control in the first equation and an odd coupling term in the second one. We prove that the nonlinear system is small-time locally null-controllable. The main difficulty is that the linearized system is not null-controllable. To overcome this obstacle, we extend in a nonlinear setting the strategy introduced in [LB19] that consists in constructing odd controls for the linear heat equation. The proof relies on three main steps. First, we obtain from the classical parabolic Carleman estimate, conjugated with maximal regularity results, a weighted observability inequality for the nonhomogeneous heat equation. Secondly, we perform a duality argument, close to the well-known Hilbert Uniqueness Method in a reflexive Banach setting, to prove that the heat equation perturbed by a source term is null-controllable thanks to odd controls. Finally, the nonlinearity is handled with a Schauder fixed-point argument.
Keywords: Null-controllability, parabolic system, nonlinear coupling, Carleman estimate
2020 Mathematics Subject Classification. 35K45, 35K58, 93B05, 93C10
1 Introduction
Let be a positive time, , be a bounded, connected, open subset of of class corresponding to the spatial domain and be a nonempty open subset such that . In what follows, we use the notation for the characteristic function of .
The null-controllability of the heat equation described below was first obtained by Fattorini and Russell [FR71] for and by Lebeau, Robbiano [LR95] and Fursikov, Imanuvilov [FI96] for . More precisely for any , there exists such that the solution of the system
(1.1) |
satisfies . These results were then extended to a large number of other parabolic systems, linear or nonlinear. For instance, the null-controllability of linear coupled parabolic systems has been a challenging issue for the control community in the last two decades. In that direction, we can quote, among the large literature devoted to this problem, [AKBDGB09], where Ammar-Khodja, Benabadallah, Dupaix, Gonzalez-Burgos exhibit sharp conditions for the null-controllability of systems of the form
(1.2) |
Here, at time , is the state, is the control, with is the diffusion matrix, is the coupling matrix and represents the distribution of controls. One objective is to reduce the number of controls (and in particular to have ) by using the coupling matrices and . Let us also quote the survey [AKBGBdT11] for other results and open problems in that direction.
In this article, we consider the following controlled semi-linear reaction-diffusion system
(1.3) |
where , and . In (1.3), at time , is the state while is the control. We are interested in the null-controllability of (1.3), that is find a control , supported in , that steers the state to zero at time , i.e. . Note that (1.3) is a so-called “cascade system” because the first equation is decoupled from the second equation. For such a system, the basic idea is to use the nonlinear coupling term , as an indirect control term, that acts on the second component .
1.1 Main results
Our control results on (1.3) are written in the framework of weak solutions. More precisely, we define the Banach space
(1.4) |
and we consider solutions of (1.3) such that . The precise definition of the weak solutions of (1.3) is given in 2.7 and a corresponding well-posedness result is stated in 2.8 for controls with
(1.5) |
Our first main result can be stated as follows.
Theorem 1.1.
Here and in all that follows, we use the notation if there exists a constant such that we have the inequality . In the whole paper, we use as a generic positive constant that does not depend on the other terms of the inequality. The value of the constant may change from one appearance to another. Our constants may depend on the geometry (, ), on the time and on the dimension . If we want to emphasize the dependence on a quantity , we write .
As we will see, the smallness conditions on the initial data i.e. (1.7) and on the control i.e. (1.8) are sufficient conditions to guarantee the well-posedness of the system (1.3), see 2.8 below.
Before continuing, let us make some comments related to 1.1.
-
•
The sufficient condition (1.6) ensuring the local null-controllability of (1.3) is actually necessary. Indeed, if then the second equation of (1.3) is decoupled from the first equation so cannot be driven to at time . Moreover, if is even, the strong maximum principle shows that we can not control : assume for instance that , then
(1.10) and thus , with satisfies
with and we can apply the standard strong maximum principle (see, for instance, [Eva10, Theorem 12, p.397]): if and then for all , in .
-
•
The linear case
is already treated in [dT00] by de Teresa. To obtain such a result, the author shows a Carleman estimate and deduce from it an observability inequality for the adjoint system.
-
•
For the semi-linear case, the main idea is to linearize the system in order to use the previous result. However, if , in the linearized system around the trajectory , we can see that the second equation is decoupled from the first one and thus can not be controlled; the linearized system is thus not null-controllable.
-
•
To overcome this difficulty, Coron, Guerrero, Rosier [CGR10] use the return method in the case
More precisely, they construct a reference trajectory of (1.3) starting from , reaching and satisfying in . Then they linearize (1.3) around the reference trajectory and obtain for the second equation
(1.11) They can then use [dT00] to obtain that the null-controllability of the linearized system and then the local null-controllability of the nonlinear system (1.3) by a fixed-point argument.
- •
Strategy of the proof. We proceed in two steps: in the first step, we control the first equation of (1.3) in the time interval . Using the small-time local null-controllability of the semi-linear heat equation, there exists a control such that . Using the smallness assumptions, we can ensure that the second equation of (1.3) admits a solution on . In the second step, we control this second equation thanks to a fictitious odd control . More precisely, we can consider the control problem
(1.12) |
where and where has a compact support in , . We then need a control such that , satisfying and such that is regular. Such a control is given by our second main result (1.2) stated below. We can then set in
To simplify the work and without loss of generality, we assume in what follows that
The proof of 1.1 crucially relies on the construction of odd controls for the semi-linear heat equation that we present now. For , we thus consider the system
(1.13) |
The definition of the weak solutions for the above system and a corresponding well-posedness result are given in 2.4 and 2.5. Our second main result states as follows.
Theorem 1.2.
Assume that , , and . There exists such that for every initial data such that
(1.14) |
there exists a control satisfying
(1.15) |
(1.16) |
and such that the solution of (1.13) satisfies
(1.17) |
and
(1.18) |
As for 1.1, the smallness conditions (1.14) and (1.15) are sufficient to guarantee the well-posedness of the semi-linear heat equation (1.13), see 2.5 below.
Before continuing, let us make some comments related to 1.2.
- •
-
•
For , that is the linear case, the result of 1.2 is still true and has already been established by the first author, see [LB19, Proposition 3.7]. One can even obtain a (small-time) global null-controllability result with odd controls due to the linear setting. Note that here, we extend the result of [LB19] in the case of a linear heat equation with a source term, see Section 3.3.
Strategy of the proof. First, we use a classical Carleman estimate for the nonhomogeneous heat equation to obtain a weighted observability inequality stated in 3.3. From this result and after that, we need to take care about the weights appearing in the norm of the adjoint system they have to be “comparable”. We then deduce from this result a weighted observability inequality, see 3.4 below with an arbitrary large . As a consequence, a null-controllability result is obtained for the heat equation with a source term and with odd controls. Let us remark that taking large enough allows us to do only one bootstrap argument for getting the desired odd behavior for the control, see 3.6 below. This is different from [LB19, Theorem 4.4 and Proposition 4.9] where two such arguments are used for obtaining the null-controllability of the heat equation with odd controls. Another bootstrap argument is then required in order to deal with the nonlinearity in the fixed-point argument, see 3.8 below. Finally, a Schauder fixed-point argument, see Section 3.5, is performed to obtain 1.2. We can remark that here due to our method for constructing the control, in this fixed point argument, the corresponding nonlinear mapping is -Hölder continuous with . In particular, a Banach fixed point argument does not seem to apply.
1.2 Outline of the paper
The outline of the paper is as follows. In Section 2, we recall some standard facts about well-posedness, regularity results for linear and nonlinear heat equations in various functional settings. We notably prove that (1.13) and (1.3) are locally well-posed, see 2.5 and 2.8 below. Section 3 and Section 4 are devoted to the proofs of the main results, i.e. 1.1 and 1.2.
2 Well-posedness and regularity results for the heat equation
In this section, we give the notion of solutions that we consider in what follows. Then we recall standard well-posedness results for both linear and semi-linear heat equations in various functional settings we will use in what follows.
2.1 Functional spaces
In this article, we use in a crucial way a framework with . First, we introduce the standard notation for the dual exponent of defined by the relation
We also introduce the following functional spaces
(2.1) |
We have the following classical embedding results (see, for instance, [LSU68, Lemma 3.3, p.80]): for ,
(2.2) |
(2.3) |
We also have, see for instance [LSU68, Lemma 3.4, p.82],
(2.4) |
where denotes the fractional Sobolev spaces (see, for instance, [LSU68, p.70]). We recall that functions in admit a trace on if . If , we denote by the subspace of functions such that on . We also write if . From [DD12, Corollary 4.53, p.216], we have
and thus
(2.5) |
We finish with some other classical results on the spaces , for which we give a short proof for completeness.
Lemma 2.1.
The following statements hold.
-
1.
If , then is an algebra.
-
2.
For any , , if
(2.6) then the embedding
(2.7)
Proof.
For the first point, we consider . Then
We thus deduce that
For the second point, we can use (2.6) to consider such that
(2.8) |
We thus deduce from (2.2) that
and from the Hölder inequality, there exists such that
(2.9) |
From the Aubin-Lions lemma (see, for instance, [Sim87, Section 8, Corollary 4]), the embedding
Consequently, if is a bounded sequence of , it has a subsequence converging in and bounded in . From (2.9), this subsequence is converging in . ∎
2.2 Linear heat equation
Let us first consider the linear nonhomogenenous heat equation
(2.10) |
In this article, we need several definitions of solutions for (2.10):
Definition 2.2.
We recall the following implications
and the reverse implications are also true assuming that is regular enough. We also note that the definition of weak solution is meaningful due to the continuous embedding (see, for instance, [Eva10, Theorem 3, p.303])
(2.13) |
We also state standard results for the well-posedness of (2.10) (see, for instance [Eva10, Theorems 3 and 4, pp.378-379], [LSU68, Theorem 7.1, p.181] and [LSU68, Theorem 9.1, p.341]):
Theorem 2.3.
The following well-posedness results hold.
2.3 Semi-linear heat equation
For , , let us then consider the semi-linear heat equation
(2.17) |
The space is defined in (1.4). First we recall the definition of a weak solution for the system (2.17):
Definition 2.4.
Let us state the following well-posedness result for (2.17) for small data. This result is standard, but we recall the proof for completeness.
Theorem 2.5.
Proof.
First, we show that for any , there exists a unique weak solution to the heat equation
(2.22) |
In order to do this, we can write with
(2.23) |
Applying 2.3, the above systems admit respectively a unique solution and and with the hypotheses on , we deduce from (2.2) that . We conclude the existence and the uniqueness of a weak solution of (2.22) and we have the estimate
(2.24) |
We can thus define the following mapping
(2.25) |
where is the unique weak solution to (2.22) and if and satisfy (2.20) and if we consider
(2.26) |
then we deduce from (2.24) that for small enough, . We can also show a similar way that the restriction of on is a strict contraction. The Banach fixed point yields the existence of a unique fixed point and the corresponding solution of (2.22) is a weak solution of (2.17).
For the uniqueness, we consider two solutions of (2.17). Then, satisfies (in a weak sense)
(2.27) |
In particular, using that , we can write the standard energy estimate: for any ,
and we conclude with the Grönwall lemma. ∎
We now state some regularizing effects of (2.17).
Lemma 2.6.
Proof.
The above definition and properties can be extended to the parabolic system
(2.30) |
More precisely, we have the following definition and well-posedness results:
Definition 2.7.
3 Proof of 1.2
The goal of this part is to prove 1.2.
We first set
(3.1) |
and
(3.2) |
Using 2.6, that is taking the control in in order to benefit from the regularizing effect of the semi-linear heat equation (1.13), we see that it is sufficient to show the following result.
Theorem 3.1.
Assume , and . Let us consider satisfying
(3.3) |
with large enough so that
(3.4) |
and satisfying (2.6). There exist and such that for any initial data with
there exists a control and a strong solution of (1.13) such that
(3.5) |
together with the estimate
(3.6) |
In particular, satisfies (1.15) and (1.16) and satisfies (1.18).
3.1 Carleman estimate and observability inequality for the heat equation
The goal of this part is to deduce a weighted observability inequality for the heat equation from a Carleman estimate. We first recall a standard Carleman estimate for the heat equation that is due to Fursikov and Imanuvilov [FI96]. We start by introducing a nonempty domain such that on . By using [FI96], see also [TW09, Theorem 9.4.3], there exists satisfying
(3.7) |
We then define the following functions:
(3.8) |
We can now state the Carleman estimate for the heat equation, see [FCG06, Lemma 1.3].
Theorem 3.2.
There exist such that for any , , with on ,
(3.9) |
From the above result, one can obtain a similar estimate with weights depending only on time. We recall that and are defined in (3.1) and (3.2). We have that and
(3.10) |
Moreover, we have the following instrumental estimates
(3.11) |
With the above notation, we can state the following corollary of 3.2.
Corollary 3.3.
Assume . Then, there exist with
(3.12) |
such that for any with on the following relation holds
(3.13) |
We want to highlight the fact that the dependence in space of the Carleman weights appearing in (3.9) has been removed in (3.13). Moreover, it is worth mentioning that the vanishing property at of the Carleman weights for the left-hand-side of (3.9) has been dropped. This is why one can make appeared the first left-hand-side of (3.13), that is the classical left-hand-side term for proving a observability inequality. The same remark applies for the first right-hand-side term of (3.9) to get the first right-hand-side term of (3.13). Finally, the fact that and are comparable is quantified in (3.12).
Proof of 3.3.
We consider and from 3.2. Then, we deduce from (3.7) and (3.8) that for any , ,
Therefore combining these estimates with (3.7) and (3.8) and taking , we deduce that
with
We now fix large enough, so that (3.12) holds. Applying (3.9), we obtain
(3.14) |
Using (3.1) and (3.2), the above relation yields
(3.15) |
Let us consider , in , in and . Then
(3.16) |
By using the maximal regularity of the heat equation in i.e. 2.3 with to (3.16) and the Sobolev embedding (2.4) we deduce
and thus by using that in and in , we obtain
Combining this last estimate with (3.15) and (3.10), we deduce the expected observability inequality (3.13). ∎
3.2 A weighted observability inequality
The goal of this part is to deduce from the weighted observability inequality in 3.3 a weighted observability inequality for , by applying maximal regularity results for the heat equation. More precisely, we show the following result:
Proposition 3.4.
Assume and . Then, there exist with
(3.17) |
such that for any with on , the following relation holds
(3.18) |
The main difference between (3.18) and (3.13) is the framework. We want to highlight that of (3.12) has been transformed into of (3.17). Basically, the proof is as follows. By a bootstrap argument, we apply recursively maximal regularity results in , starting from together with Sobolev embeddings to obtain (3.18). During the induction process, becomes then , etc. to finally take the value .
Proof.
First, we apply 3.3 to obtain satisfying (3.12) and such that (3.13) holds for any with on . We then set so that for any ,
(3.19) |
In particular, if we consider then by (3.12) and (3.11)
so that
(3.20) |
We can apply the maximal regularity result in , i.e. 2.3 with to (3.19), and use (3.20) and the observability inequality (3.13) to deduce
(3.21) |
We then use the Sobolev embedding (2.2) to deduce
(3.22) |
with defined by
Then, we consider so that from (3.11),
and with (3.22) and (3.21), we deduce
(3.23) |
Now we apply 2.3 to
(3.24) |
with , and using (3.23), we obtain
(3.25) |
If , then using and , we deduce from the above relation the desired observability inequality (3.18) with . Else, we have and we can repeat the argument, that is we use the Sobolev embedding (2.2) to deduce
(3.26) |
with defined by
Taking , and proceeding as above, applying 2.3 with and using (3.26) and (3.25), we find
We can proceed by induction and since decrease by at each step, after a finite number of steps, we obtain and we deduce the desired observability inequality (3.18). ∎
3.3 Controllability of the heat equation with a source term in
We use the above observability results to show, by a duality argument, the controllability of a linear system associated with (1.13):
(3.27) |
In order to control the above system, we fix and we consider and as in 3.4. Then, we introduce
(3.28) |
and we define the following norm for ,
(3.29) |
The fact that it is a norm is a consequence of the weighted observability inequality (3.18). We denote by the completion of with respect to the norm .
First, we have the following result that roughly states that a function belongs to some suitable weighted spaces.
Lemma 3.5.
Assume . Then, for any ,
(3.30) |
Proof.
Using , (3.17) and (3.11), we have
(3.31) |
Now, if , then
(3.32) |
Combining the observability inequality (3.18) and (3.31), we deduce
Applying the maximal regularity result 2.3 on (3.32) and using the above relation, we deduce the second estimate in (3.30). For the first estimate, we use (3.10) to obtain that and this allows us to conclude the proof. ∎
We now introduce some functional spaces: for and , we set
(3.33) |
(3.34) |
endowed with the following norm
(3.35) |
Let us consider, for any and , the functional defined as follows:
(3.36) |
Using the observability inequality (3.18), we can check that is a strictly convex and coercive functional on . In particular, admits a unique minimum . We can thus define, for and , the following maps
(3.37) |
Proposition 3.6.
Assume and and let us consider and given by 3.4. For any and , let us set
(3.38) |
- 1.
-
2.
Odd behavior of the control. The control satisfies and
(3.41) together with the estimate
(3.42) -
3.
Regularity of the solution. Assume that and that on if . Then for any , together with the estimate
(3.43) In particular, .
The first point will be obtained from Euler-Lagrange equation. The odd behavior of the control, i.e. (3.42), remarking that is odd, comes from the identification of in (3.38), (3.37) and from a weighted estimate of . Finally, the regularity result on the solution comes from a maximal parabolic regularity result. Note that if , then and so that we do not need to impose the compatibility condition on .
Proof of 3.6.
We start by writing the Euler-Lagrange equation for at to obtain
(3.44) |
Taking in the above relation and using Young’s inequality and the observability inequality (3.18), we deduce (3.39).
Moreover, (3.44) and (3.38) imply
(3.45) |
that is is the very weak solution to (3.27) associated with the control , and in the sense of 2.2.
3.4 bound on the control and estimate of the nonlinearity
From now on, we assume and we assume that and are given by 3.4 with this . In particular they satisfy (3.17) which yields
First we have the following result on the control .
Lemma 3.7.
In the above result, has to be sufficiently large to get that is an algebra, and this enables us to get that is sufficiently smooth, because , as expected in (1.16).
Proof of 3.7.
Proposition 3.8.
The goal of the above result is to get an appropriate bound on the nonlinearity, this would be a first step in order to prove the local null-controllability of the semi-linear heat equation.
Proof.
We define as follows
(3.54) |
In both cases, we have and .
We deduce from (3.43) and the Sobolev embedding (2.2) that for any ,
(3.55) |
We then consider satisfying (3.48). We have in particular and we can write
Applying 2.3 on the above equation and using (3.49) and (3.55) with together with (3.10), (3.11), we deduce
(3.56) |
We can proceed by induction, using again (2.2), and since the corresponding sequence decreases by (see (3.54)) at each step, we obtain after a finite number of steps that for any satisfying (3.48), we have
(3.57) |
Using that satisfies (2.6) so that the Sobolev embedding (2.7) holds, we deduce that
(3.58) |
From (3.17) and (3.51), we have
so that we can take in (3.57), (3.58) and we deduce (3.53). ∎
3.5 A Schauder fixed-point argument
Let us consider the hypotheses of 3.8 and assume . Then, using the conclusion of 3.8, we can define the mapping
(3.59) |
where . Moreover, using (3.53), we deduce that if is small enough, then the closed set
(3.60) |
is invariant by .
Proposition 3.9.
The mapping defined above is continuous and is relatively compact into .
Proof.
Let us consider a sequence of . We write . Then we can use (3.52) to obtain that is bounded in . Applying 2.1, we deduce that, up to a subsequence,
for some . We deduce that is relatively compact into .
To show the continuity of , we consider and we write (see (3.37) and (3.38)) for ,
From the Euler-Lagrange equation (3.44) for and , we deduce
(3.61) |
In the above relation, we take in the above relation and we combine it with the observability inequality (3.18) and with the relation
to deduce
(3.62) |
Moreover, using that
we obtain from (3.10)
Thus, if satisfies (3.48), the above relation combined with (3.10), (3.4) that guarantees that is an algebra and 3.5 yields
Therefore, using (3.39) and (3.62), we find
(3.63) |
Note that satisfies the following system
Now, we follow the same proof as in 3.8 and we use that satisfies (3.48) to deduce from (3.63) that
(3.64) |
We then write
so that from Hölder’s inequality, we have
Combining this relation with the Sobolev embedding (2.2), (3.52), (3.64), we deduce that
(3.65) |
which implies the continuity of . ∎
Remark 3.10.
In the above proof, let us remark that we show that the mapping is -Hölder continuous with (see (3.65)). It is not clear if this mapping is Lipschitz continuous or if we can show that for small enough it is contractive. As a consequence, in the proof of 3.1, we do not apply the Banach fixed-point theorem (as it can be done with the method proposed in [LTT13]) and we use instead the Schauder fixed-point theorem.
We are now in a position to prove 3.1.
Proof of 3.1.
From 3.9, if is small enough, then the mapping defined by (3.59) is continuous, where is the closed convex set defined by (3.60). Moreover, is relatively compact in and we can thus apply the Schauder fixed point theorem to deduce the existence of a fixed point . Setting and , we can apply 3.6, 3.7 and 3.8 and obtain that satisfies (1.16), that is the strong solution of (1.13) associating with and and that for any satisfying (3.48), , , , together with the estimates (3.6). ∎
4 Proof of 1.1
The goal of this part is to prove the local null-controllability of (1.3).
Proof of 1.1.
As explained in the introduction, the proof is divided into two steps.
Step 1: control of the first equation in .
First we apply 2.5: there exists small enough such that if
(4.1) |
the system
(4.2) |
admits a unique weak solution in the sense of 2.4. Now we apply 1.2 to
(4.3) |
There exists such that for any with
(4.4) |
there exists a control such that and
Assuming (1.7) with possibly smaller, we have that
satisfies (4.1) so that we have obtained at this step a control , such that (1.3) admits a weak solution in and . By using 2.6, satisfies
Step 2: control of the second equation in through a fictitious odd control.
References
- [AKBDGB09] Farid Ammar Khodja, Assia Benabdallah, Cédric Dupaix, and Manuel González-Burgos. A generalization of the Kalman rank condition for time-dependent coupled linear parabolic systems. Differ. Equ. Appl., 1(3):427–457, 2009.
- [AKBGBdT11] Farid Ammar-Khodja, Assia Benabdallah, Manuel González-Burgos, and Luz de Teresa. Recent results on the controllability of linear coupled parabolic problems: a survey. Math. Control Relat. Fields, 1(3):267–306, 2011.
- [AT02] Sebastian Anita and Daniel Tataru. Null controllability for the dissipative semilinear heat equation. Appl. Math. Optim., 46(2-3):97–105, 2002. Special issue dedicated to the memory of Jacques-Louis Lions.
- [CGR10] Jean-Michel Coron, Sergio Guerrero, and Lionel Rosier. Null controllability of a parabolic system with a cubic coupling term. SIAM J. Control Optim., 48(8):5629–5653, 2010.
- [DD12] Françoise Demengel and Gilbert Demengel. Functional spaces for the theory of elliptic partial differential equations. Universitext. Springer, London; EDP Sciences, Les Ulis, 2012. Translated from the 2007 French original by Reinie Erné.
- [dT00] Luz de Teresa. Insensitizing controls for a semilinear heat equation. Comm. Partial Differential Equations, 25(1-2):39–72, 2000.
- [Eva10] Lawrence C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, second edition, 2010.
- [FCG06] Enrique Fernández-Cara and Sergio Guerrero. Global Carleman inequalities for parabolic systems and applications to controllability. SIAM J. Control Optim., 45(4):1399–1446, 2006.
- [FI96] Andrei Fursikov and Oleg Imanuvilov. Controllability of evolution equations, volume 34 of Lecture Notes Series. Seoul National University, Research Institute of Mathematics, Global Analysis Research Center, Seoul, 1996.
- [FR71] Hector O. Fattorini and David L. Russell. Exact controllability theorems for linear parabolic equations in one space dimension. Arch. Rational Mech. Anal., 43:272–292, 1971.
- [Fri64] Avner Friedman. Partial differential equations of parabolic type. Prentice-Hall, Inc., Englewood Cliffs, N.J., 1964.
- [LB19] Kévin Le Balc’h. Null-controllability of two species reaction-diffusion system with nonlinear coupling: a new duality method. SIAM J. Control Optim., 57(4):2541–2573, 2019.
- [LR95] Gilles Lebeau and Luc Robbiano. Contrôle exact de l’équation de la chaleur. Comm. Partial Differential Equations, 20(1-2):335–356, 1995.
- [LSU68] Olga A. Ladyženskaja, Vsevolod A. Solonnikov, and Nina N. Uralceva. Linear and quasilinear equations of parabolic type. Translations of Mathematical Monographs, Vol. 23. American Mathematical Society, Providence, R.I., 1968. Translated from the Russian by S. Smith.
- [LTT13] Yuning Liu, Takéo Takahashi, and Marius Tucsnak. Single input controllability of a simplified fluid-structure interaction model. ESAIM Control Optim. Calc. Var., 19(1):20–42, 2013.
- [Sim87] Jacques Simon. Compact sets in the space . Ann. Mat. Pura Appl. (4), 146:65–96, 1987.
- [TW09] Marius Tucsnak and George Weiss. Observation and control for operator semigroups. Birkhäuser Advanced Texts: Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks]. Birkhäuser Verlag, Basel, 2009.