This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Boundary controllability of the Korteweg-de Vries equation: The Neumann case

Roberto de A. Capistrano Filho Departamento de Matemática, Universidade Federal de Pernambuco (UFPE), 50740-545, Recife (PE), Brazil, www.ufpe.br/capistranofilho ([email protected]).    Jandeilson Santos da Silva ) Departamento de Matemática, Universidade Federal de Pernambuco (UFPE), 50740-545, Recife (PE), Brazil ([email protected]
Abstract

This article gives a necessary first step to understanding the critical set phenomenon for the Korteweg-de Vries (KdV) equation posed on interval [0,L][0,L] considering the Neumann boundary conditions with only one control input. We showed that the KdV equation is controllable in the critical case, i.e., when the spatial domain LL belongs to the set c\mathcal{R}_{c}, where c1c\neq-1 and

c:={2π3(c+1)m2+ml+m2;m,l}{mπc+1;m},\mathcal{R}_{c}:=\left\{\frac{2\pi}{\sqrt{3(c+1)}}\sqrt{m^{2}+ml+m^{2}};\ m,l\in\mathbb{N}^{*}\right\}\cup\left\{\frac{m\pi}{\sqrt{c+1}};\ m\in\mathbb{N}^{*}\right\},

the KdV equation is exactly controllable in L2(0,L)L^{2}(0,L). The result is achieved using the return method together with a fixed point argument.

keywords:
Korteweg–de Vries equation, exact boundary controllability, Neumann boundary conditions, Dirichlet boundary conditions, critical set
{AMS}

Primary, 35Q53, 93B05; Secondary, 37K10

1 Introduction

We had known when we formulated the waves as a free boundary problem of the incompressible, irrotational Euler equation in an appropriate non-dimensional form, there exist two non-dimensional parameters δ:=hλ\delta:=\frac{h}{\lambda} and ε:=ah\varepsilon:=\frac{a}{h}, where the water depth, the wavelength and the amplitude of the free surface are parameterized as h,λh,\lambda and aa, respectively. See, for instance, [2, 3, 1, 4, 23, 26] and references therein for a rigorous justification. Moreover, another non-dimensional parameter μ\mu appears, the Bond number, to measure the importance of gravitational forces compared to surface tension forces also appears. Considering the physical condition δ1\delta\ll 1 we can characterize the waves, called long waves or shallow water waves. In particular, considering the relations between ε\varepsilon and δ\delta, we can have the following regime:

  • Korteweg-de Vries (KdV): ε=δ21\varepsilon=\delta^{2}\ll 1 and μ13\mu\neq\frac{1}{3}. Under this regime, Korteweg and de Vries [20]111This equation was first introduced by Boussinesq [6], and Korteweg and de Vries rediscovered it twenty years later. derived the following well-known equation as a central equation among other dispersive or shallow water wave models called the KdV equation from the equations for capillary-gravity waves:

    ±2ηt+3ηηx+(13μ)ηxxx=0.\pm 2\eta_{t}+3\eta\eta_{x}+\left(\frac{1}{3}-\mu\right)\eta_{xxx}=0.

Today, it is well known that this equation has an important phenomenon that directly affects the control problem related to them, the so-called critical length phenomenon. Let us briefly present the control problem, which makes the phenomenon of critical lengths emerge. The control problem was presented in a pioneering work of Rosier [24] that studied the following system

{ut+ux+uux+uxxx=0, in (0,L)×(0,T),u(0,t)=0, u(L,t)=0, ux(L,t)=g(t), in (0,T),u(x,0)=u0(x), in (0,L),\left\{\begin{array}[c]{lll}u_{t}+u_{x}+uu_{x}+u_{xxx}=0,&&\text{ in }(0,L)\times(0,T),\\ u(0,t)=0,\text{ }u(L,t)=0,\text{ }u_{x}(L,t)=g(t),&&\text{ in }(0,T),\\ u(x,0)=u_{0}(x),&&\text{ in }(0,L),\end{array}\right. (1.1)

where the boundary value function g(t)g(t) is considered as a control input. Precisely, the author showed the following control problem for the system (1.1), giving the origin of the critical length phenomenon for the KdV equation.

Question 𝒜\mathcal{A}: Given T>0T>0 and u0,uTL2(0,L)u_{0},u_{T}\in L^{2}(0,L), can one find an appropriate control input g(t)L2(0,T)g(t)\in L^{2}(0,T) such that the corresponding solution u(x,t)u(x,t) of the system (1.1) satisfies

u(x,0)=u0(x)andu(x,T)=uT(x)?u(x,0)=u_{0}(x)\quad\text{and}\quad u(x,T)=u_{T}(x)? (1.2)

Rosier answered the previous question in [24]. He proved that considering L𝒩L\notin\mathcal{N}, where

𝒩:={2π3k2+kl+l2:k,l},\mathcal{N}:=\left\{\frac{2\pi}{\sqrt{3}}\sqrt{k^{2}+kl+l^{2}}\,:k,\,l\,\in\mathbb{N}^{\ast}\right\},

the associated linear system (1.1)

{ut+ux+uxxx=0, in (0,L)×(0,T),u(0,t)=0, u(L,t)=0, ux(L,t)=g(t), in (0,T),u(x,0)=u0(x) in (0,L),\left\{\begin{array}[c]{lll}u_{t}+u_{x}+u_{xxx}=0,&&\text{ in }(0,L)\times(0,T),\\ u(0,t)=0,\text{ }u(L,t)=0,\text{ }u_{x}(L,t)=g(t),&&\text{ in }(0,T),\\ u(x,0)=u_{0}(x)&&\text{ in }(0,L),\end{array}\right. (1.3)

is controllable; roughly speaking, if L𝒩L\in\mathcal{N} system (1.3) is not controllable, that is, there exists a finite-dimensional subspace of L2(0,L)L^{2}(0,L), denoted by =(L)\mathcal{M}=\mathcal{M}(L), which is unreachable from 0 for the linear system. More precisely, for every nonzero state ψ\psi\in\mathcal{M}, gL2(0,T)g\in L^{2}(0,T) and uC([0,T];L2(0,L))L2(0,T;H1(0,L))u\in C([0,T];L^{2}(0,L))\cap L^{2}(0,T;H^{1}(0,L)) satisfying (1.3) and u(,0)=0u(\cdot,0)=0, one has u(,T)ψu(\cdot,T)\neq\psi.

Definition 1.1.

A spatial domain (0,L)(0,L) is called critical for the system (1.3) if its domain length LL belongs to 𝒩\mathcal{N}.

Following the work of Rosier [24], the boundary control system of the KdV equation posed on the finite interval (0,L)(0,L) with various control inputs has been intensively studied (cf. [7, 8, 10, 11, 15, 17, 18, 19] and see [9, 25] for more complete reviews). Thus, this work gives a necessary step to understanding this phenomenon for the KdV equation with Neumann boundary conditions, completing, in some sense, the results shown in [7].

1.1 Problem set

In this article, we study a class of distributed parameter control systems described by the Korteweg–de Vries (KdV) equation posed on a bounded domain (0,L)(0,L) with the Neumann boundary conditions

{ut+ux+uxxx+uux=0, in (0,L)×(0,T),uxx(0,t)=uxx(L,t)=0, in (0,T),ux(L,t)=h(t), in (0,T),u(x,0)=u0(x), in (0,L),\left\{\begin{array}[c]{lll}u_{t}+u_{x}+u_{xxx}+uu_{x}=0,&&\text{ in }(0,L)\times(0,T),\\ u_{xx}(0,t)=u_{xx}(L,t)=0,&&\text{ in }(0,T),\\ u_{x}(L,t)=h(t),&&\text{ in }(0,T),\\ u(x,0)=u_{0}(x),&&\text{ in }(0,L),\end{array}\right. (1.4)

where h(t)h(t) will be considered as a control input. Recently, the first author dealt with the control problem related to the system (1.4). Precisely, was proved that their solutions are exactly controllable in a neighborhood of cc if the length LL of the spatial domain (0,L)(0,L) does not belong to the set

c:={2π3(c+1)m2+ml+m2;m,l}{mπc+1;m},\mathcal{R}_{c}:=\left\{\frac{2\pi}{\sqrt{3(c+1)}}\sqrt{m^{2}+ml+m^{2}};\ m,l\in\mathbb{N}^{*}\right\}\cup\left\{\frac{m\pi}{\sqrt{c+1}};\ m\in\mathbb{N}^{*}\right\},

that is, the relation (1.2) holds for the solution of the system (1.4). The result can be read as follows.

Theorem 1.2 (Caicedo, Capistrano–Filho, Zhang [7]).

Let T>0T>0, c1c\neq-1 and LcL\notin\mathcal{R}_{c}. There exists δ>0\delta>0 such that for any u0,uTL2(0,L)u_{0},u_{T}\in L^{2}(0,L) with

u0cL2(0,L)<δ and uTcL2(0,L)<δ\|u_{0}-c\|_{L^{2}(0,L)}<\delta\text{ \ and \ }\|u_{T}-c\|_{L^{2}(0,L)}<\delta

one can find hL2(0,T)h\in L^{2}(0,T) such that the system (1.4) admits a unique solution

u𝒵T=C([0,T];L2(0,L))L2(0,T;H1(0,L))u\in\mathcal{Z}_{T}=C\left([0,T];L^{2}(0,L)\right)\cap L^{2}\left(0,T;H^{1}(0,L)\right)

satisfying (1.2).

As in [24], the first step is to obtain a control result for the linear system, namely,

{vt+(1+c)vx+vxxx=0, in (0,L)×(0,T),vxx(0,t)=vxx(L,t)=0, in (0,T),vx(L,t)=h(t), in (0,T),v(x,0)=v0(x), in (0,L).\left\{\begin{array}[c]{lll}v_{t}+(1+c)v_{x}+v_{xxx}=0,&\text{ \ in \ }(0,L)\times(0,T),\\ v_{xx}(0,t)=v_{xx}(L,t)=0,&\text{ \ in \ }(0,T),\\ v_{x}(L,t)=h(t),&\text{ \ in \ }(0,T),\\ v(x,0)=v_{0}(x),&\text{ \ in \ }(0,L).\end{array}\right. (1.5)

Precisely the authors in [7] proved the following result.

Theorem 1.3 (Caicedo, Capistrano–Filho, Zhang [7]).

For c1c\neq-1, the linear system (1.5) is exactly controllable in the space L2(0,L)L^{2}(0,L) if and only if LcL\notin\mathcal{R}_{c}. Otherwise, that is, if c=1c=-1, the system (1.5) is not exactly controllable in the space L2(0,L)L^{2}(0,L) for any L>0L>0.

With this in hand, they extend the result for the nonlinear system (1.4) using fixed point argument achieving Theorem 1.2 whenever LcL\notin\mathcal{R}_{c}. It is important to point out that fixing kk\in\mathbb{N}^{*} and considering m=l=km=l=k we have L=2kπL=2k\pi, when c=0c=0, and, from Theorem 1.3, it follows that (1.5) are not exactly controllable. Additionally, as before mentioned, we do not know if the system (1.4) is exactly controllable. So, in this context, the natural questions appear:

Question \mathcal{B}: Given T>0T>0, LcL\in\mathcal{R}_{c} and y0,yTL2(0,L)y_{0},y_{T}\in L^{2}(0,L) close enough to cc, can one find an appropriate control input hL2(0,T)h\in L^{2}(0,T) such that the solution yy of the system (1.9), corresponding to hh and y0y_{0}, satisfies y(,T)=yTy(\cdot,T)=y_{T}?

Question 𝒞\mathcal{C}: Given T>0T>0, LcL\in\mathcal{R}_{c} and y0,yTL2(0,L)y_{0},y_{T}\in L^{2}(0,L). The system (1.9) is exactly controllable in the critical length LL, that is when LcL\in\mathcal{R}_{c}?

1.2 Main results

Let us consider the following nonlinear control system

{yt+(1+c)yx+yxxx+yyx=0, in (0,L)×(0,T),yxx(0,t)=0,yx(L,t)=h(t),yxx(L,t)=0, in (0,T),y(x,0)=y0(x), in (0,L),\displaystyle\left\{\begin{array}[]{lll}y_{t}+(1+c)y_{x}+y_{xxx}+yy_{x}=0,&\text{ in }(0,L)\times(0,T),\\ y_{xx}(0,t)=0,\ y_{x}(L,t)=h(t),\ y_{xx}(L,t)=0,&\text{ in }(0,T),\\ y(x,0)=y_{0}(x),&\text{ in }(0,L),\end{array}\right. (1.9)

where hh will be as a control input and yy is the state. Theorem 1.2 says that when LcL\notin\mathcal{R}_{c}, system (1.9) is locally controllable around cc but we do not know if the same holds when LcL\in\mathcal{R}_{c}. The main result in this work provides an affirmative answer to the Question 𝒞\mathcal{C}. Precisely, we have the following:

Theorem 1.4.

Let T>0T>0, c=0c=0 and L0L\in\mathcal{R}_{0}. Then, system (1.9) is exactly controllable around the origin 0 in L2(0,L)L^{2}(0,L), that is, there exists δ>0\delta>0 such that, for very y0,yTL2(0,L)y_{0},y_{T}\in L^{2}(0,L) with

y0L2(0,L),yTL2(0,L)<δ\displaystyle\|y_{0}\|_{L^{2}(0,L)},\|y_{T}\|_{L^{2}(0,L)}<\delta

it is possible to find hL2(0,T)h\in L^{2}(0,T) such that the corresponding solution of (1.9) satisfying y(,0)=y0y(\cdot,0)=y_{0} and y(,T)=yTy(\cdot,T)=y_{T}.

This previous result can be generalized for any cc as follows:

Theorem 1.5.

Let T>0T>0 and LcL\in\mathcal{R}_{c}. The system (1.9) is exactly controllable around cc in L2(0,L)L^{2}(0,L) in the sense of Theorem 1.4.

To prove the previous results we need an auxiliary property that ensures that for cc near enough to 0 (small perturbations of 0), the system (1.4) is exactly controllable in a neighborhood of cc in c)\mathcal{R}_{c}), or precisely, for dd close enough to cc one has LdL\notin\mathcal{R}_{d} so that, the system (1.5) corresponding to dd is exactly controllable, answering the Question \mathcal{B}. In other words, the set of critical lengths is sensitive to small disturbances in equilibrium cc, and the result can be read as follows.

Theorem 1.6.

Let T>0T>0, c1c\neq-1 and LcL\in\mathcal{R}_{c}. There exists εc>0\varepsilon_{c}>0 such that, for every d(cεc,c+εc)\{c}d\in(c-\varepsilon_{c},c+\varepsilon_{c})\backslash\{c\}, d1d\neq-1, we have LdL\notin\mathcal{R}_{d}. Consequently, the linear system (1.5), with c=dc=d, is exactly controllable; and the nonlinear system (1.9) is exactly controllable around the steady state dd in L2(0,L)L^{2}(0,L), that is, there exists δd>0\delta_{d}>0 such that, for any y0,yTL2(0,L)y_{0},y_{T}\in L^{2}(0,L) with

y0dL2(0,L)<δd and yTdL2(0,L)<δd,\displaystyle\|y_{0}-d\|_{L^{2}(0,L)}<\delta_{d}\text{ \ and \ }\|y_{T}-d\|_{L^{2}(0,L)}<\delta_{d},

one can find hL2(0,T)h\in L^{2}(0,T) such that the system (1.9) admits a unique solution y𝒵Ty\in\mathcal{Z}_{T} satisfying y(,0)=y0y(\cdot,0)=y_{0} and y(,T)=yTy(\cdot,T)=y_{T}.

1.3 Heuristic and paper’s outline

The proof of the Theorem 1.6 is based on the topological properties of real numbers together with the Theorem 1.2. Moreover, with this in hand, both results stated in the previous paragraph (Theorems 1.4 and 1.5) rely on the so-called return method together with the fixed point argument.

It is important to point out that the return method was introduced by J.M. Coron in [12] (see also [13]) and has been used by several authors to prove control results in the critical lengths for the KdV-type equation (see, for instance, [8, 10, 11, 15]). This method consists of building particular trajectories of the system (1.9) starting and ending at some equilibrium such that the linearization of the system around these trajectories has good properties. Here, we use a combination of this method with a fixed point argument, successfully applied in [16]. We mention that this method can be applied together with quasi-static deformations and power series expansion, we infer to the reader the nice book of Coron [14] about more details of the method.

Concerning the construction of solutions to the Theorems 1.4 and 1.5 we follow the following procedure: In the first time, we construct a solution that starts from y0y_{0} and reaches at time T/3T/3 a state which is in some sense close to dd (which is yet to be defined). Then we construct a solution (close to the state solution dd ), which starts at time 2T/32T/3 from the previous state. In the last step, we bring the latter state to 0 via a function y2y_{2}, as we can see in Figure 1 below. For details of this construction see the characterization of the function yy in (3.52).

Refer to caption
Figure 1: Solutions driving states close to 0 to constants and vice versa.

We finish this introduction with an outline of this work which consists of four parts, including the introduction. Section 2 gives an overview of the well-posedness of the system (1.9). Section 3 is devoted to proving carefully the controllability of the system (1.9) when LcL\in\mathcal{R}_{c}. Precisely, in the first part of Section 3, we deal with the proof of Theorem 1.6. In the second part, we prove the construction of the function yy mentioned before, and finally, in the third part of Section 3, we use these previous results to achieve Theorem 1.4.

2 Overview of the well-posedness theory

In this section, we review the well-posedness theory for the KdV equation. The results presented here can be found in [5, 7, 21]. For that, consider L>0L>0 and T0,T1T_{0},T_{1}\in\mathbb{R} with T0<T1T_{0}<T_{1}. We define the space

𝒵T0,T1:=C([T0,T1];L2(0,L))L2([T0,T1];H1(0,L))\displaystyle\mathcal{Z}_{T_{0},T_{1}}:=C\left([T_{0},T_{1}];L^{2}(0,L)\right)\cap L^{2}\left([T_{0},T_{1}];H^{1}(0,L)\right)

which is a Banach space with the following norm

y𝒵T0,T1:=maxt[T0,T1]y(,t)L2(0,L)+(T0T1y(,t)H1(0,L)2𝑑t)1/2.\displaystyle\|y\|_{\mathcal{Z}_{T_{0},T_{1}}}:=\max_{t\in[T_{0},T_{1}]}\|y(\cdot,t)\|_{L^{2}(0,L)}+\left(\int_{T_{0}}^{T_{1}}\|y(\cdot,t)\|_{H^{1}(0,L)}^{2}dt\right)^{1/2}.

For any T>0T>0 we denote 𝒵0,T\mathcal{Z}_{0,T} simply by 𝒵T\mathcal{Z}_{T}.

Additionally, let T>0T>0 be given and consider the space

T:=H13(0,T)×L2(0,T)×H13(0,T)\displaystyle\mathcal{H}_{T}:=H^{-\frac{1}{3}}(0,T)\times L^{2}(0,T)\times H^{-\frac{1}{3}}(0,T)

with a norm

(h1,h2,h3)T:=h1H13(0,T)+h2L2(0,T)+h3H13(0,T).\|(h_{1},h_{2},h_{3})\|_{\mathcal{H}_{T}}:=\|h_{1}\|_{H^{-\frac{1}{3}}(0,T)}+\|h_{2}\|_{L^{2}(0,T)}+\|h_{3}\|_{H^{-\frac{1}{3}}(0,T)}.

The next proposition, showed in [7, Proposition 2.5], provides the well-posedness to the following system

{ut+uxxx=f, in (0,L)×(0,T),uxx(0,t)=h1(t),ux(L,t)=h2(t),uxx(L,t)=h3(t), in (0,T),u(x,0)=u0, in (0,L).\displaystyle\left\{\begin{array}[]{ll}u_{t}+u_{xxx}=f,&\ \text{ in }(0,L)\times(0,T),\\ u_{xx}(0,t)=h_{1}(t),\ u_{x}(L,t)=h_{2}(t),\ u_{xx}(L,t)=h_{3}(t),&\ \text{ in }(0,T),\\ u(x,0)=u_{0},&\ \text{ in }(0,L).\end{array}\right. (2.13)
Proposition 2.1 (Caicedo, Capistrano-Filho, Zhang [7]).

For any v0L2(0,L)v_{0}\in L^{2}(0,L), 𝐡=(h1,h2,h3)T{\bf h}=(h_{1},h_{2},h_{3})\in\mathcal{H}_{T} and fL1(0,T,L2(0,L))f\in L^{1}(0,T,L^{2}(0,L)), the IBVP (2.13) admits a unique mild solution u𝒵Tu\in\mathcal{Z}_{T}, which satisfies

xjuLx(0,L;H(1j)/3(0,T)),j=0,1,2.\displaystyle\partial_{x}^{j}u\in L^{\infty}_{x}(0,L;H^{(1-j)/3}(0,T)),\ j=0,1,2.

Moreover, there exists C1>0C_{1}>0 such that

u𝒵T+j=02xjuLx(0,L;H(1j)/3(0,T))C1(u0L2(0,L)+𝐡T+fL1(0,T;L2(0,L))).\displaystyle\|u\|_{\mathcal{Z}_{T}}+\sum_{j=0}^{2}\left\|\partial_{x}^{j}u\right\|_{L^{\infty}_{x}(0,L;H^{(1-j)/3}(0,T))}\leq C_{1}\left(\|u_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right).
Remark 2.2.

We highlight that the constant C1C_{1} in the above result depends on TT. However, if θ(0,T]\theta\in(0,T] then, the estimates in the Proposition 2.1 hold with the same constant C1C_{1} corresponding to TT.

In fact, let u0L2(0,L)u_{0}\in L^{2}(0,L), 𝐡θ,fL1(0,θ;L2(0,L)){\bf h}\in\mathcal{H}_{\theta},\ f\in L^{1}(0,\theta;L^{2}(0,L)) and u𝒵θu\in\mathcal{Z}_{\theta} the solution of (2.13) corresponding to these datas. We extend 𝐡{\bf h} and ff to [0,T][0,T] (we will also denote these extensions by 𝐡{\bf h} and ff) putting

𝐡=0 in (θ,T]andf=0 in (θ,T].{\bf h}=0\text{ in }(\theta,T]\quad\text{and}\quad f=0\text{ in }(\theta,T].

Now, denote by u~𝒵T\tilde{u}\in\mathcal{Z}_{T} the corresponding solution of (2.13). Then, from Proposition 2.1, we have u~|[0,θ]=u\tilde{u}\big{|}_{[0,\theta]}=u and

u𝒵θu~𝒵T\displaystyle\|u\|_{\mathcal{Z}_{\theta}}\leq\|\tilde{u}\|_{\mathcal{Z}_{T}} C1(u0L2(0,L)+𝐡T+fL1(0,T;L2(0,L)))\displaystyle\leq C_{1}\left(\|u_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right)
=C1(u0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L))).\displaystyle=C_{1}\left(\|u_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right).

Using the Proposition 2.1 we can get properties for the linear problem

{yt+(ay)x+yxxx=f, in (0,L)×(0,T),yxx(0,t)=h1(t),yx(L,t)=h2(t),yxx(L,t)=h3(t), in (0,T),y(x,0)=y0(x), in (0,L),\displaystyle\left\{\begin{array}[]{ll}y_{t}+(ay)_{x}+y_{xxx}=f,&\ \text{ in }(0,L)\times(0,T),\\ y_{xx}(0,t)=h_{1}(t),\ y_{x}(L,t)=h_{2}(t),\ y_{xx}(L,t)=h_{3}(t),&\ \text{ in }(0,T),\\ y(x,0)=y_{0}(x),&\ \text{ in }(0,L),\end{array}\right. (2.17)

where a𝒵Ta\in\mathcal{Z}_{T} is given. To do this, the following lemma will be very useful and was proved in [22, Lemma 3].

Lemma 2.3 (Kramer, Zhang [22]).

There exists a constant C>0C>0 such that

0Tuvx(,t)L2(0,L)𝑑tC(T12+T13)u𝒵Tv𝒵T,\displaystyle\int_{0}^{T}\|uv_{x}(\cdot,t)\|_{L^{2}(0,L)}dt\leq C\left(T^{\frac{1}{2}}+T^{\frac{1}{3}}\right)\|u\|_{\mathcal{Z}_{T}}\|v\|_{\mathcal{Z}_{T}},

for every u,v𝒵Tu,v\in\mathcal{Z}_{T}.

With these previous results in hand, the following proposition gives us the well-posedness to the general system (2.17), which will be used several times so, for the sake of completeness, we will give the proof.

Proposition 2.4.

For any y0L2(0,L)y_{0}\in L^{2}(0,L), 𝐡=(h1,h2,h3)T{\bf h}=(h_{1},h_{2},h_{3})\in\mathcal{H}_{T} and fL1(0,T,L2(0,L))f\in L^{1}(0,T,L^{2}(0,L)), the IBVP (2.17) admits a unique mild solution y𝒵Ty\in\mathcal{Z}_{T}, which satisfies

xjyLx(0,L;H(1j)/3(0,T)),j=0,1,2,\displaystyle\partial_{x}^{j}y\in L^{\infty}_{x}(0,L;H^{(1-j)/3}(0,T)),\ j=0,1,2,

and

y𝒵TC2(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L))),\displaystyle\|y\|_{\mathcal{Z}_{T}}\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right),

for some positive constant C2C_{2} which depends only on TT and a𝒵T\|a\|_{\mathcal{Z}_{T}}. In addition, the solution yy possesses the following sharp trace estimates

j=02xjyLx(0,L;H(1j)/3(0,T))C2(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L))).\displaystyle\sum_{j=0}^{2}\left\|\partial_{x}^{j}y\right\|_{L^{\infty}_{x}(0,L;H^{(1-j)/3}(0,T))}\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right).

In particular, the map (y0,𝐡,f)y(y_{0},{\bf h},f)\mapsto y is Lipschitz continuous.

Proof 2.5.

Let y0L2(0,L)y_{0}\in L^{2}(0,L), 𝐡T{\bf h}\in\mathcal{H}_{T} and fL1(0,T;L2(0,L))f\in L^{1}(0,T;L^{2}(0,L)) be given. Consider θ\theta satisfying 0<θT0<\theta\leq T and define the map Γ:𝒵θ𝒵θ\Gamma:\mathcal{Z}_{\theta}\rightarrow\mathcal{Z}_{\theta} in the next way: for y𝒵θy\in\mathcal{Z}_{\theta}, put Γy\Gamma y being the solution of

{ut+uxxx=f(ay)x, in (0,L)×(0,θ),uxx(0,t)=h1(t),ux(L,t)=h2(t),uxx(L,t)=h3(t), in (0,θ),u(x,0)=y0(x), in (0,L).\displaystyle\left\{\begin{array}[]{ll}u_{t}+u_{xxx}=f-(ay)_{x},&\ \text{ in }(0,L)\times(0,\theta),\\ u_{xx}(0,t)=h_{1}(t),\ u_{x}(L,t)=h_{2}(t),\ u_{xx}(L,t)=h_{3}(t),&\ \text{ in }(0,\theta),\\ u(x,0)=y_{0}(x),&\ \text{ in }(0,L).\end{array}\right.

Consider the set

B={y𝒵θ;y𝒵θr},\displaystyle B=\left\{y\in\mathcal{Z}_{\theta};\ \|y\|_{\mathcal{Z}_{\theta}}\leq r\right\},

with r>0r>0 to be determined later. From Proposition 2.1 and Lemma 2.3 we have for any yBy\in B the following estimate

Γy𝒵θ\displaystyle\|\Gamma y\|_{\mathcal{Z}_{\theta}}\leq C1(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L))+(ay)xL1(0,θ;L2(0,L)))\displaystyle C_{1}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}+\|(ay)_{x}\|_{L^{1}(0,\theta;L^{2}(0,L))}\right) (2.18)
\displaystyle\leq C1(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L)))\displaystyle C_{1}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right)
+2C1C(θ12+θ13)a𝒵θy𝒵θ\displaystyle+2C_{1}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|a\|_{\mathcal{Z}_{\theta}}\|y\|_{\mathcal{Z}_{\theta}}
\displaystyle\leq C1(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L)))\displaystyle C_{1}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right)
+2C1C(θ12+θ13)a𝒵Ty𝒵θ.\displaystyle+2C_{1}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|a\|_{\mathcal{Z}_{T}}\|y\|_{\mathcal{Z}_{\theta}}.

Choosing

r=2C1(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L)))r=2C_{1}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right)

and θ\theta satisfying

2C1C(θ12+θ13)a𝒵T<12,\displaystyle 2C_{1}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|a\|_{\mathcal{Z}_{T}}<\frac{1}{2}, (2.19)

the inequality (2.18) give us

Γy𝒵θr2+r2=r,\displaystyle\|\Gamma y\|_{\mathcal{Z}_{\theta}}\leq\frac{r}{2}+\frac{r}{2}=r,

that is, Γ(B)B\Gamma(B)\subset B. Furthermore, ΓyΓw\Gamma y-\Gamma w solves

{ut+uxxx=[a(y+w)]x, in (0,L)×(0,θ),uxx(0,t)=ux(L,t)=uxx(L,t)=0, in (0,θ),u(x,0)=0, in (0,L).\displaystyle\left\{\begin{array}[]{ll}u_{t}+u_{xxx}=[a(-y+w)]_{x},&\ \text{ in }(0,L)\times(0,\theta),\\ u_{xx}(0,t)=u_{x}(L,t)=u_{xx}(L,t)=0,&\ \text{ in }(0,\theta),\\ u(x,0)=0,&\ \text{ in }(0,L).\end{array}\right.

So, from Proposition 2.1, Lemma 2.3 and inequality (2.19), we have

ΓyΓw𝒵θ\displaystyle\|\Gamma y-\Gamma w\|_{\mathcal{Z}_{\theta}} C1[a(yw)]xL1(0,θ;L2(0,L))\displaystyle\leq C_{1}\|[a(y-w)]_{x}\|_{L^{1}(0,\theta;L^{2}(0,L))}
2C1C(θ12+θ13)a𝒵θyw𝒵θ\displaystyle\leq 2C_{1}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|a\|_{\mathcal{Z}_{\theta}}\|y-w\|_{\mathcal{Z}_{\theta}}
2C1C(θ12+θ13)a𝒵Tyw𝒵θ\displaystyle\leq 2C_{1}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|a\|_{\mathcal{Z}_{T}}\|y-w\|_{\mathcal{Z}_{\theta}}
<12yw𝒵θ.\displaystyle<\frac{1}{2}\|y-w\|_{\mathcal{Z}_{\theta}}.

Thus, Γ:BB\Gamma:B\rightarrow B is a contraction so that, by Banach’s fixed point theorem, Γ\Gamma has a fixed point yBy\in B which is a solution to the problem (2.17) in [0,θ][0,\theta], corresponding to data (y0,𝐡,f)(y_{0},{\bf h},f). Additionally, inequalities (2.18) and (2.19) yields that

y𝒵θC1(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L)))+12y𝒵θ\displaystyle\|y\|_{\mathcal{Z}_{\theta}}\leq C_{1}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right)+\frac{1}{2}\|y\|_{\mathcal{Z}_{\theta}}

and, therefore,

y𝒵θ2C1(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L))).\displaystyle\|y\|_{\mathcal{Z}_{\theta}}\leq 2C_{1}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right).

Since θ\theta depends only on aa, with a standard continuation extension argument, the solution yy can be extended to interval [0,T][0,T] and the following estimate holds

y𝒵TC2(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L))),\displaystyle\|y\|_{\mathcal{Z}_{T}}\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right), (2.20)

for some suitable constant C2>0C_{2}>0 which only depends on TT and a𝒵T\|a\|_{\mathcal{Z}_{T}}. Therefore, it follows from (2.20) that the map (y0,𝐡,f)y(y_{0},{\bf h},f)\mapsto y is Lipschitz continuous and, as a consequence of this, we have the uniqueness of the solution yy in 𝒵T\mathcal{Z}_{T}. The sharp trace estimates follow as a consequence of the Proposition 2.1, showing the proposition.

Now, we will study the well-posedness of the nonlinear problem, namely:

{yt+(ay)x+yxxx+yyx=f, in (0,L)×(0,T),yxx(0,t)=h1(t),yx(L,t)=h2(t),yxx(L,t)=h3(t), in (0,T),y(x,0)=y0(x), in (0,L).\displaystyle\left\{\begin{array}[]{ll}y_{t}+(ay)_{x}+y_{xxx}+yy_{x}=f,&\ \text{ in }(0,L)\times(0,T),\\ y_{xx}(0,t)=h_{1}(t),\ y_{x}(L,t)=h_{2}(t),\ y_{xx}(L,t)=h_{3}(t),&\ \text{ in }(0,T),\\ y(x,0)=y_{0}(x),&\ \text{ in }(0,L).\end{array}\right. (2.24)

The result can be read as follows.

Proposition 2.6.

Let a𝒵Ta\in\mathcal{Z}_{T} be given. For every y0L2(0,L)y_{0}\in L^{2}(0,L), 𝐡=(h1,h2,h3)T{\bf h}=(h_{1},h_{2},h_{3})\in\mathcal{H}_{T} and fL1(0,T,L2(0,L))f\in L^{1}(0,T,L^{2}(0,L)) there exists a unique mild solution y𝒵Ty\in\mathcal{Z}_{T} of (2.24) which satisfies

y𝒵TC3(y0L2(0,L)+𝐡L2(0,T)+fL1(0,T;L2(0,L)))\|y\|_{\mathcal{Z}_{T}}\leq C_{3}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{L^{2}(0,T)}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right)

and

j=02xjyLx(0,L;H(1j)/3(0,T))\displaystyle\sum_{j=0}^{2}\left\|\partial_{x}^{j}y\right\|_{L^{\infty}_{x}(0,L;H^{(1-j)/3}(0,T))}\leq C3(y0L2(0,L)+hL2(0,T)+fL1(0,T;L2(0,L)))\displaystyle C_{3}\left(\|y_{0}\|_{L^{2}(0,L)}+\|h\|_{L^{2}(0,T)}~{}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right)
+C3(y0L2(0,L)+hL2(0,T)+fL1(0,T;L2(0,L)))2,\displaystyle+C_{3}\left(\|y_{0}\|_{L^{2}(0,L)}+\|h\|_{L^{2}(0,T)}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right)^{2},

for some constant C3>0C_{3}>0. Furthermore, the corresponding solution map SS is locally Lipschitz continuous, that is, given λ>0\lambda>0 there exist Lλ>0L_{\lambda}>0 such that, for every

(y0,𝐡,f0),(y1,𝐠,f1)L2(0,L)×T×L1(0,T;L2(0,L))(y_{0},{\bf h},f_{0}),(y_{1},{\bf g},f_{1})\in L^{2}(0,L)\times\mathcal{H}_{T}\times L^{1}(0,T;L^{2}(0,L))

with

y0L2(0,L)+𝐡T+f0L1(0,T;L2(0,L))<λ,\displaystyle\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f_{0}\|_{L^{1}(0,T;L^{2}(0,L))}<\lambda,
y1L2(0,L)+𝐠T+f1L1(0,T;L2(0,L))<λ,\displaystyle\|y_{1}\|_{L^{2}(0,L)}+\|{\bf g}\|_{\mathcal{H}_{T}}+\|f_{1}\|_{L^{1}(0,T;L^{2}(0,L))}<\lambda,

we have

S(y0,𝐡,f0)S(y1,𝐠,f1)𝒵TLλ(y0y1L2(0,L)+𝐡𝐠T+f0f1L1(0,T;L2(0,L))).\begin{split}\|S(y_{0},{\bf h},f_{0})-S(y_{1},{\bf g},f_{1})\|_{\mathcal{Z}_{T}}\leq&L_{\lambda}\left(\|y_{0}-y_{1}\|_{L^{2}(0,L)}\right.\\ &\left.+\|{\bf h}-{\bf g}\|_{\mathcal{H}_{T}}+\|f_{0}-f_{1}\|_{L^{1}(0,T;L^{2}(0,L))}\right).\end{split}
Proof 2.7.

We will proceed as follows: First, we will show the existence of a solution and obtain the desired estimates. Secondly, we will get an estimate that provides the uniqueness of the solution and guarantees that SS is locally Lipschitz continuous.

To do that, consider y0L2(0,L)y_{0}\in L^{2}(0,L), 𝐡T{\bf h}\in\mathcal{H}_{T} and fL1(0,T;L2(0,L))f\in L^{1}(0,T;L^{2}(0,L)). Let θ\theta satisfying 0<θT0<\theta\leq T and define the map Γ:𝒵θ𝒵θ\Gamma:\mathcal{Z}_{\theta}\rightarrow\mathcal{Z}_{\theta} in the following way: For y𝒵θy\in\mathcal{Z}_{\theta}, pick Γy\Gamma y as solution of the following problem

{ut+(au)x+uxxx=fyyx, in (0,L)×(0,θ),uxx(0,t)=h1(t),ux(L,t)=h2(t),uxx(L,t)=h3(t), in (0,θ),u(x,0)=y0(x), in (0,L).\displaystyle\left\{\begin{array}[]{ll}u_{t}+(au)_{x}+u_{xxx}=f-yy_{x},&\ \text{ in }(0,L)\times(0,\theta),\\ u_{xx}(0,t)=h_{1}(t),\ u_{x}(L,t)=h_{2}(t),\ u_{xx}(L,t)=h_{3}(t),&\ \text{ in }(0,\theta),\\ u(x,0)=y_{0}(x),&\ \text{ in }(0,L).\end{array}\right.

Consider the set

B={y𝒵θ;y𝒵θr},\displaystyle B=\left\{y\in\mathcal{Z}_{\theta};\ \|y\|_{\mathcal{Z}_{\theta}}\leq r\right\},

with r>0r>0 to be determined later. Using Proposition 2.4 and Lemma 2.3 we obtain, for any yBy\in B,

Γy𝒵θ\displaystyle\|\Gamma y\|_{\mathcal{Z}_{\theta}} C2(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L))+yyxL1(0,θ;L2(0,L)))\displaystyle\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}+\|yy_{x}\|_{L^{1}(0,\theta;L^{2}(0,L))}\right) (2.25)
C2(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L)))+C2C(θ12+θ13)y𝒵T2.\displaystyle\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right)+C_{2}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|y\|_{\mathcal{Z}_{T}}^{2}.

Choosing

r=4C2(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L)))\displaystyle r=4C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right)

and θ\theta satisfying

C2C(θ12+θ13)r<14,\displaystyle C_{2}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)r<\frac{1}{4}, (2.26)

inequality (2.25) give us

Γ𝒵T\displaystyle\|\Gamma\|_{\mathcal{Z}_{T}} r4+r4=r2<r,\displaystyle\leq\frac{r}{4}+\frac{r}{4}=\frac{r}{2}<r,

that is, Γ(B)B\Gamma(B)\subset B. Moreover given y,wBy,w\in B, ΓyΓw\Gamma y-\Gamma w solves

{ut+(au)x+uxxx=yyx+wwx, in (0,L)×(0,θ),uxx(0,t)=0,ux(L,t)=0,uxx(L,t)=0, in (0,θ),u(x,0)=0, in (0,L).\displaystyle\left\{\begin{array}[]{ll}u_{t}+(au)_{x}+u_{xxx}=-yy_{x}+ww_{x},&\ \text{ in }(0,L)\times(0,\theta),\\ u_{xx}(0,t)=0,\ u_{x}(L,t)=0,\ u_{xx}(L,t)=0,&\ \text{ in }(0,\theta),\\ u(x,0)=0,&\ \text{ in }(0,L).\end{array}\right.

Therefore, from Proposition 2.4,

ΓyΓw𝒵θC2yyxwwxL1(0,θ;L2(0,L)).\displaystyle\|\Gamma y-\Gamma w\|_{\mathcal{Z}_{\theta}}\leq C_{2}\|yy_{x}-ww_{x}\|_{L^{1}(0,\theta;L^{2}(0,L))}.

Note that

yyxwwx=12[(y+w)x(yw)+(y+w)(yw)x].\displaystyle yy_{x}-ww_{x}=\frac{1}{2}\big{[}(y+w)_{x}(y-w)+(y+w)(y-w)_{x}\big{]}.

Then, thanks to the Lemma 2.3, we get that

yyxwwxL1(0,θ;L2(0,L))\displaystyle\|yy_{x}-ww_{x}\|_{L^{1}(0,\theta;L^{2}(0,L))}
12[(y+w)x(yw)L1(0,θ;L2(0,L))+(y+w)(yw)xL1(0,θ;L2(0,L))]\displaystyle\leq\frac{1}{2}\big{[}\|(y+w)_{x}(y-w)\|_{L^{1}(0,\theta;L^{2}(0,L))}+\|(y+w)(y-w)_{x}\|_{L^{1}(0,\theta;L^{2}(0,L))}\big{]}
12[C(θ12+θ13)y+w𝒵θyw𝒵θ+C(θ12+θ13)y+w𝒵θyw𝒵θ]\displaystyle\leq\frac{1}{2}\left[C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|y+w\|_{\mathcal{Z}_{\theta}}\|y-w\|_{\mathcal{Z}_{\theta}}+C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|y+w\|_{\mathcal{Z}_{\theta}}\|y-w\|_{\mathcal{Z}_{\theta}}\right]
=C(θ12+θ13)y+w𝒵θyw𝒵θ\displaystyle=C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|y+w\|_{\mathcal{Z}_{\theta}}\|y-w\|_{\mathcal{Z}_{\theta}}
2C(θ12+θ13)ryw𝒵θ\displaystyle\leq 2C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)r\|y-w\|_{\mathcal{Z}_{\theta}}

and, using (2.26) yields

ΓyΓw𝒵θ2C2C(θ12+θ13)ryw𝒵θ<12yw𝒵θ.\displaystyle\|\Gamma y-\Gamma w\|_{\mathcal{Z}_{\theta}}\leq 2C_{2}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)r\|y-w\|_{\mathcal{Z}_{\theta}}<\frac{1}{2}\|y-w\|_{\mathcal{Z}_{\theta}}.

Hence, Γ:BB\Gamma:B\rightarrow B is a contraction so that, by Banach’s fixed point theorem, Γ\Gamma has a fixed point yBy\in B which is a solution to the problem (2.24) in [0,θ][0,\theta], corresponding to data (y0,𝐡,f)(y_{0},{\bf h},f).

Furthermore, due to the inequalities (2.25) and (2.26), we have

y𝒵θC2(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L)))+14y𝒵θ,\displaystyle\|y\|_{\mathcal{Z}_{\theta}}\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right)+\frac{1}{4}\|y\|_{\mathcal{Z}_{\theta}},

and, therefore

y𝒵θ43C2(y0L2(0,L)+𝐡θ+fL1(0,θ;L2(0,L))).\displaystyle\|y\|_{\mathcal{Z}_{\theta}}\leq\frac{4}{3}C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{\theta}}+\|f\|_{L^{1}(0,\theta;L^{2}(0,L))}\right).

Since θ\theta depends only on aa, with a standard continuation extension argument, the solution yy can be extended to interval [0,T][0,T] and the following estimate holds

y𝒵TC~3(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L))),\displaystyle\|y\|_{\mathcal{Z}_{T}}\leq\tilde{C}_{3}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right), (2.27)

for some suitable constant C~3>0\tilde{C}_{3}>0 which only depends on TT and aa. Now, using one more time the Proposition 2.6 together with Lemma 2.3 we obtain

j=02xjyLx(0,L;H(1j)/3(0,T))\displaystyle\sum_{j=0}^{2}\left\|\partial_{x}^{j}y\right\|_{L^{\infty}_{x}(0,L;H^{(1-j)/3}(0,T))}
C2(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L))+yyxL1(0,T;L2(0,L)))\displaystyle\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}+\|yy_{x}\|_{L^{1}(0,T;L^{2}(0,L))}\right)
C2(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L)))+C2C(θ12+θ13)y𝒵T2.\displaystyle\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right)+C_{2}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\|y\|_{\mathcal{Z}_{T}}^{2}.

By (2.27) it follows that

j=02xjyLx(0,L;H(1j)/3(0,T))\displaystyle\sum_{j=0}^{2}\left\|\partial_{x}^{j}y\right\|_{L^{\infty}_{x}(0,L;H^{(1-j)/3}(0,T))}
C2(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L)))\displaystyle\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right)
+C2C(θ12+θ13)C~2(y0L2(0,L)+𝐡T+fL1(0,T;L2(0,L)))2.\displaystyle+C_{2}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\tilde{C}^{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f\|_{L^{1}(0,T;L^{2}(0,L))}\right)^{2}.

Choosing

C3:=max{C~3,C2,C2C(θ12+θ13)C~2}\displaystyle C_{3}:=\max\left\{\tilde{C}_{3},C_{2},C_{2}C\left(\theta^{\frac{1}{2}}+\theta^{\frac{1}{3}}\right)\tilde{C}^{2}\right\}

we obtain the desired estimates.

Now, consider

(y0,𝐡,f0),(y1,𝐠,f1)L2(0,L)×T×L1(0,T;L2(0,L))(y_{0},{\bf h},f_{0}),(y_{1},{\bf g},f_{1})\in L^{2}(0,L)\times\mathcal{H}_{T}\times L^{1}(0,T;L^{2}(0,L))

with

y0L2(0,L)+𝐡T+f0L1(0,T;L2(0,L))<λ,\displaystyle\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f_{0}\|_{L^{1}(0,T;L^{2}(0,L))}<\lambda,
y1L2(0,L)+𝐠T+f1L1(0,T;L2(0,L))<λ.\displaystyle\|y_{1}\|_{L^{2}(0,L)}+\|{\bf g}\|_{\mathcal{H}_{T}}+\|f_{1}\|_{L^{1}(0,T;L^{2}(0,L))}<\lambda.

Write 𝐡=(h1,h2,h3){\bf h}=(h_{1},h_{2},h_{3}) and 𝐠=(g1,g2,g3){\bf g}=(g_{1},g_{2},g_{3}). Let yy and uu be solutions of

{yt+(ay)x+yxxx+yyx=f0, in (0,L)×(0,T),yxx(0,t)=h1(t),yx(L,t)=h2(t),yxx(L,t)=h3(t), in (0,T),y(x,0)=y0(x) in (0,L),\displaystyle\left\{\begin{array}[]{ll}y_{t}+(ay)_{x}+y_{xxx}+yy_{x}=f_{0},&\ \text{ in }(0,L)\times(0,T),\\ y_{xx}(0,t)=h_{1}(t),\ y_{x}(L,t)=h_{2}(t),\ y_{xx}(L,t)=h_{3}(t),&\ \text{ in }(0,T),\\ y(x,0)=y_{0}(x)&\ \text{ in }(0,L),\end{array}\right.

and

{ut+(au)x+uxxx+uux=f1, in (0,L)×(0,T),uxx(0,t)=g1(t),ux(L,t)=g2(t),uxx(L,t)=g3(t), in (0,T),u(x,0)=y1(x) in (0,L),\displaystyle\left\{\begin{array}[]{ll}u_{t}+(au)_{x}+u_{xxx}+uu_{x}=f_{1},&\ \text{ in }(0,L)\times(0,T),\\ u_{xx}(0,t)=g_{1}(t),\ u_{x}(L,t)=g_{2}(t),\ u_{xx}(L,t)=g_{3}(t),&\ \text{ in }(0,T),\\ u(x,0)=y_{1}(x)&\ \text{ in }(0,L),\end{array}\right.

respectively. Then, w=yuw=y-u solves the problem

{wt+[(a+12(y+u))w]x+wxxx=f0f1, in (0,L)×(0,T,)wxx(0,)=h1g1,wx(L,)=h2g2,wxx(L,)=h3g3, in (0,T),w(x,0)=y0(x)y1(x), in (0,L).\displaystyle\left\{\begin{array}[]{ll}w_{t}+\big{[}\left(a+\frac{1}{2}(y+u)\right)w\big{]}_{x}+w_{xxx}=f_{0}-f_{1},&\ \text{ in }(0,L)\times(0,T,)\\ w_{xx}(0,\cdot)=h_{1}-g_{1},\ w_{x}(L,\cdot)=h_{2}-g_{2},\ w_{xx}(L,\cdot)=h_{3}-g_{3},&\ \text{ in }(0,T),\\ w(x,0)=y_{0}(x)-y_{1}(x),&\ \text{ in }(0,L).\end{array}\right.

From Proposition 2.4 it follows that

yu𝒵TD(y0y1L2(0,L)+𝐡𝐠T+f0f1L1(0,T;L2(0,L))),\displaystyle\|y-u\|_{\mathcal{Z}_{T}}\leq D\left(\|y_{0}-y_{1}\|_{L^{2}(0,L)}+\|{\bf h}-{\bf g}\|_{\mathcal{H}_{T}}+\|f_{0}-f_{1}\|_{L^{1}(0,T;L^{2}(0,L))}\right), (2.28)

where DD is a constant which depends on TT and a+12(y+u)𝒵T\|a+\frac{1}{2}(y+u)\|_{\mathcal{Z}_{T}}. But, using (2.27) we have

a+12(y+u)𝒵T\displaystyle\|a+\frac{1}{2}(y+u)\|_{\mathcal{Z}_{T}}\leq a𝒵T+12(y𝒵T+u𝒵T)\displaystyle\|a\|_{\mathcal{Z}_{T}}+\frac{1}{2}\left(\|y\|_{\mathcal{Z}_{T}}+\|u\|_{\mathcal{Z}_{T}}\right)
\displaystyle\leq a𝒵T+C~32(y0L2(0,L)+𝐡T+f0L1(0,T;L2(0,L)))\displaystyle\|a\|_{\mathcal{Z}_{T}}+\frac{\tilde{C}_{3}}{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|{\bf h}\|_{\mathcal{H}_{T}}+\|f_{0}\|_{L^{1}(0,T;L^{2}(0,L))}\right)
+C~32(y1L2(0,L)+𝐠T+f1L1(0,T;L2(0,L)))\displaystyle+\frac{\tilde{C}_{3}}{2}\left(\|y_{1}\|_{L^{2}(0,L)}+\|{\bf g}\|_{\mathcal{H}_{T}}+\|f_{1}\|_{L^{1}(0,T;L^{2}(0,L))}\right)

and, consequently,

a+12(y+u)𝒵T\displaystyle\|a+\frac{1}{2}(y+u)\|_{\mathcal{Z}_{T}} a𝒵T+C~3λ.\displaystyle\leq\|a\|_{\mathcal{Z}_{T}}+\tilde{C}_{3}\lambda.

Hence, the constant DD can be chosen depending only on TT and a𝒵T\|a\|_{\mathcal{Z}_{T}} (and also on λ\lambda), so that, (2.28) gives us the uniqueness of the solution which turns the solution map SS well-defined. Moreover, from (2.28)

S(y0,𝐡,f0)S(y1,𝐠,f1)𝒵TD(y0y1L2(0,L)+𝐡𝐠T+f0f1L1(0,T;L2(0,L))),\begin{split}\|S(y_{0},{\bf h},f_{0})-S(y_{1},{\bf g},f_{1})\|_{\mathcal{Z}_{T}}\leq&D\left(\|y_{0}-y_{1}\|_{L^{2}(0,L)}\right.\\ &\left.+\|{\bf h}-{\bf g}\|_{\mathcal{H}_{T}}+\|f_{0}-f_{1}\|_{L^{1}(0,T;L^{2}(0,L))}\right),\end{split}

which concludes the proof.

Finally, the following Lemma, whose proof can be found in [24], will be useful in the next section.

Lemma 2.8 (Rosier [24]).

If y𝒵Ty\in\mathcal{Z}_{T} then yyxL1(0,T;L2(0,L))yy_{x}\in L^{1}(0,T;L^{2}(0,L)) and the map

𝒵TL1(0,T;L2(0,L))yyyx,\displaystyle\begin{array}[]{rcl}\mathcal{Z}_{T}&\longrightarrow&L^{1}(0,T;L^{2}(0,L))\\ y&\longmapsto&yy_{x},\end{array}

is continuous. More precisely, for every y,z𝒵Ty,z\in\mathcal{Z}_{T} we have that

yyxzzxL1(0,T;L2(0,L))C4(yL2(0,T;H1(0,L))+zL2(0,T;H1(0,L)))yzL2(0,T;H1(0,L)),\begin{split}\|yy_{x}-zz_{x}\|_{L^{1}(0,T;L^{2}(0,L))}\leq&C_{4}\left(\|y\|_{L^{2}(0,T;H^{1}(0,L))}\right.\\ &\left.+\|z\|_{L^{2}(0,T;H^{1}(0,L))}\right)\|y-z\|_{L^{2}(0,T;H^{1}(0,L))},\end{split}

where C4C_{4} is a positive constant that depends only on LL.

Remark 2.9.

We end this subsection with the following remarks.

  1. 1.

    For every a𝒵Ta\in\mathcal{Z}_{T}, the Proposition 2.4 give us the well-definition for the solution operator

    Λa:L2(0,L)×T×L1(0,T;L2(0,L))𝒵T\Lambda_{a}:L^{2}(0,L)\times\mathcal{H}_{T}\times L^{1}(0,T;L^{2}(0,L))\rightarrow\mathcal{Z}_{T}

    where, for each (y0,𝐡,f)L2(0,L)×T×L1(0,T;L2(0,L))(y_{0},{\bf h},f)\in L^{2}(0,L)\times\mathcal{H}_{T}\times L^{1}(0,T;L^{2}(0,L)), Λa(y0,𝐡,f)\Lambda_{a}(y_{0},{\bf h},f) is the corresponding solution to the problem (2.17). Furthermore, the Proposition 2.6 guarantees that Λa\Lambda_{a} is a bounded linear operator.

  2. 2.

    Of course, the constant C2>0C_{2}>0 in the Proposition 2.4 depends on a𝒵T\|a\|_{\mathcal{Z}_{T}}. However, given M>0M>0, the same constant C2C_{2} can be used for every a𝒵Ta\in\mathcal{Z}_{T} with a𝒵TM\|a\|_{\mathcal{Z}_{T}}\leq M.

3 Boundary controllability in the critical lenght

In this section, we study the controllability of the system

{yt+yx+yxxx+yyx=0, in (0,L)×(0,T),yxx(0,t)=0,yx(L,t)=h(t),yxx(L,t)=0, in (0,T),y(x,0)=y0(x) in (0,L),\displaystyle\left\{\begin{array}[]{ll}y_{t}+y_{x}+y_{xxx}+yy_{x}=0,&\ \text{ in }(0,L)\times(0,T),\\ y_{xx}(0,t)=0,\ y_{x}(L,t)=h(t),\ y_{xx}(L,t)=0,&\ \text{ in }(0,T),\\ y(x,0)=y_{0}(x)&\ \text{ in }(0,L),\end{array}\right. (3.32)

when LL is a critical length. We will use the return method together with the fixed point argument to ensure the controllability of the system (3.32) when LcL\in\mathcal{R}_{c}. Before presenting the proof of the main result, let us give a preliminary result which is important in our analysis.

3.1 An auxiliary result

Given c1c\neq-1, the set of critical lengths for the linearization of (3.32) around cc is

c:={2π3(c+1)m2+ml+l2;m,l}{mπc+1;m}.\displaystyle\mathcal{R}_{c}:=\left\{\frac{2\pi}{\sqrt{3(c+1)}}\sqrt{m^{2}+ml+l^{2}}\ ;\ m,l\in\mathbb{N}^{*}\right\}\cup\left\{\frac{m\pi}{\sqrt{c+1}};\ m\in\mathbb{N}^{*}\right\}.

As mentioned before, when LcL\in\mathcal{R}_{c}, the linear system

{yt+(1+c)yx+yxxx=0, in (0,L)×(0,T),yxx(0,t)=0,yx(L,t)=h(t),yxx(L,t)=0, in (0,T),y(x,0)=y0(x), in (0,L),\displaystyle\left\{\begin{array}[]{lll}y_{t}+(1+c)y_{x}+y_{xxx}=0,&\text{ in }(0,L)\times(0,T),\\ y_{xx}(0,t)=0,\ y_{x}(L,t)=h(t),\ y_{xx}(L,t)=0,&\text{ in }(0,T),\\ y(x,0)=y_{0}(x),&\text{ in }(0,L),\end{array}\right. (3.36)

is not exactly controllable. So, in this section, we first give the proof of Theorem 1.6.

Proof 3.1.

(Proof of Theorem 1.6.) We will split the proof into two cases.

First case: L=2πj2+jk+k23(c+1)L=2\pi\frac{\sqrt{j^{2}+jk+k^{2}}}{\sqrt{3(c+1)}} for some (j,k)×(j,k)\in\mathbb{N}^{*}\times\mathbb{N}^{*}.

Let d1d\neq-1 and suppose that LdL\in\mathcal{R}_{d}. Then

L=2πm2+ml+l23(d+1);m,lL=2\pi\frac{\sqrt{m^{2}+ml+l^{2}}}{\sqrt{3(d+1)}};\ m,l\in\mathbb{N}^{*} (3.37)

or

L=mπd+1;m.L=\frac{m\pi}{\sqrt{d+1}};\ m\in\mathbb{N}^{*}. (3.38)

If (3.37) is the case, we have that

2πj2+jk+k23(c+1)=2πm2+ml+l23(d+1),2\pi\frac{\sqrt{j^{2}+jk+k^{2}}}{\sqrt{3(c+1)}}=2\pi\frac{\sqrt{m^{2}+ml+l^{2}}}{\sqrt{3(d+1)}},

which implies

d=(c+1)(m2+ml+l2)j2+jk+k21d=\frac{(c+1)(m^{2}+ml+l^{2})}{j^{2}+jk+k^{2}}-1

with m,lm,l\in\mathbb{N}^{*}.

Otherwise, if (3.38) holds, so

2πj2+jk+k23(c+1)=mπd+12\pi\frac{\sqrt{j^{2}+jk+k^{2}}}{\sqrt{3(c+1)}}=\frac{m\pi}{\sqrt{d+1}}

giving that

d=3m2(c+1)4(j2+jk+k2)1d=\frac{3m^{2}(c+1)}{4(j^{2}+jk+k^{2})}-1

where mm\in\mathbb{N}^{*}. Therefore, if LdL\in\mathcal{R}_{d} then we necessarily have d𝒜11d\in\mathcal{A}_{1}\cup\mathcal{B}_{1}. Here,

𝒜1:={(c+1)(m2+ml+l2)j2+jk+k21,m,l}\mathcal{A}_{1}:=\left\{\frac{(c+1)(m^{2}+ml+l^{2})}{j^{2}+jk+k^{2}}-1,\ m,l\in\mathbb{N}^{*}\right\}

and

1:={3m2(c+1)4(j2+jk+k2)1;m}.\mathcal{B}_{1}:=\left\{\frac{3m^{2}(c+1)}{4(j^{2}+jk+k^{2})}-1;\ m\in\mathbb{N}^{*}\right\}.

We are now in a position to prove that 𝒜11\mathcal{A}_{1}\cup\mathcal{B}_{1} is discrete. To do that, consider x,y𝒜1x,y\in\mathcal{A}_{1} such that xyx\neq y in the form

x=(c+1)(m12+m1l1+l12)j2+jk+k21x=\frac{(c+1)(m_{1}^{2}+m_{1}l_{1}+l_{1}^{2})}{j^{2}+jk+k^{2}}-1

and

y=(c+1)(m22+m2l2+l22)j2+jk+k21,y=\frac{(c+1)(m_{2}^{2}+m_{2}l_{2}+l_{2}^{2})}{j^{2}+jk+k^{2}}-1,

with m1,l1,m2,l2m_{1},l_{1},m_{2},l_{2}\in\mathbb{N}^{*}. Note that

xy=c+1j2+jk+k2[(m12+m1l1+l12)(m22+m2l2+l22)].x-y=\frac{c+1}{j^{2}+jk+k^{2}}\left[(m_{1}^{2}+m_{1}l_{1}+l_{1}^{2})-(m_{2}^{2}+m_{2}l_{2}+l_{2}^{2})\right].

Since xyx\neq y we have

(m12+m1l1+l12)(m22+m2l2+l22).(m_{1}^{2}+m_{1}l_{1}+l_{1}^{2})\neq(m_{2}^{2}+m_{2}l_{2}+l_{2}^{2}).

Thus

|(m12+m1l1+l12)(m22+m2l2+l22)|1\left|(m_{1}^{2}+m_{1}l_{1}+l_{1}^{2})-(m_{2}^{2}+m_{2}l_{2}+l_{2}^{2})\right|\geq 1

so that

d(x,y)c+1j2+jk+k2,x,y𝒜1,xy.d(x,y)\geq\frac{c+1}{j^{2}+jk+k^{2}},\ \forall x,y\in\mathcal{A}_{1},\ x\neq y.

Analogously, we get

d(x,y)3(c+1)4(j2+jk+k2),x,y1,xy.d(x,y)\geq\frac{3(c+1)}{4(j^{2}+jk+k^{2})},\ \forall\ x,y\in\mathcal{B}_{1},x\neq y.

Now, let x𝒜1x\in\mathcal{A}_{1} and y1y\in\mathcal{B}_{1} with xyx\neq y, as follow:

x=(c+1)(m2+ml+l2)j2+jk+k21x=\frac{(c+1)(m^{2}+ml+l^{2})}{j^{2}+jk+k^{2}}-1

and

y=3p2(c+1)4(j2+jk+k2)1,y=\frac{3p^{2}(c+1)}{4(j^{2}+jk+k^{2})}-1,

where m,l,pm,l,p\in\mathbb{N}^{*}. Observe that

xy=(c+1)(m2+ml+l2)j2+jk+k23p2(c+1)4(j2+jk+k2)=4(c+1)(m2+ml+l2)4(j2+jk+k2)3p2(c+1)4(j2+jk+k2)=c+14(j2+jk+k2)[4(m2+ml+l2)3p2].\begin{split}x-y=&\frac{(c+1)(m^{2}+ml+l^{2})}{j^{2}+jk+k^{2}}-\frac{3p^{2}(c+1)}{4(j^{2}+jk+k^{2})}\\ =&\frac{4(c+1)(m^{2}+ml+l^{2})}{4(j^{2}+jk+k^{2})}-\frac{3p^{2}(c+1)}{4(j^{2}+jk+k^{2})}\\ =&\frac{c+1}{4(j^{2}+jk+k^{2})}\left[4(m^{2}+ml+l^{2})-3p^{2}\right].\end{split}

Since xyx\neq y we have that 4(m2+ml+l2)4(m^{2}+ml+l^{2}) and 3p23p^{2} are distinct natural numbers so

|4(m2+ml+l2)3p2|1\left|4(m^{2}+ml+l^{2})-3p^{2}\right|\geq 1

and, consequently, we have

d(x,y)c+14(j2+jk+k2).d(x,y)\geq\frac{c+1}{4(j^{2}+jk+k^{2})}.

From all the above we conclude that

d(x,y)c+14(j2+jk+k2),x,y𝒜11,xy,d(x,y)\geq\frac{c+1}{4(j^{2}+jk+k^{2})},\ \forall\ x,y\in\mathcal{A}_{1}\cup\mathcal{B}_{1},\ x\neq y,

which implies that 𝒜11\mathcal{A}_{1}\cup\mathcal{B}_{1} is discrete.

Note that c𝒜11c\in\mathcal{A}_{1}\cup\mathcal{B}_{1} since

c=[(c+1)(j2+jk+k2)j2+jk+k21]𝒜1.c=\left[\frac{(c+1)(j^{2}+jk+k^{2})}{j^{2}+jk+k^{2}}-1\right]\in\mathcal{A}_{1}.

As any point in 𝒜11\mathcal{A}_{1}\cup\mathcal{B}_{1} is isolated, there exists ϵ1>0\epsilon_{1}>0 such that

(cϵ1,c+ϵ1)[𝒜11]={c}.(c-\epsilon_{1},c+\epsilon_{1})\cap\left[\mathcal{A}_{1}\cup\mathcal{B}_{1}\right]=\{c\}.

Therefore for d(cϵ1,c+ϵ1)\{c}d\in(c-\epsilon_{1},c+\epsilon_{1})\backslash\{c\} we have that d𝒜11d\notin\mathcal{A}_{1}\cup\mathcal{B}_{1} and, therefore, LdL\notin\mathcal{R}_{d}.

Second case: L=kπc+1L=\frac{k\pi}{\sqrt{c+1}} for some kk\in\mathbb{N}^{*}.

Let d1d\neq-1 and suppose that LdL\in\mathcal{R}_{d}, that is, (3.37) or (3.38) holds. If (3.37) holds, then

kπc+1=2πm2+ml+l23(d+1)\frac{k\pi}{\sqrt{c+1}}=2\pi\frac{\sqrt{m^{2}+ml+l^{2}}}{\sqrt{3(d+1)}}

giving that

d=(c+1)(m2+ml+l2)3k21,d=\frac{(c+1)(m^{2}+ml+l^{2})}{3k^{2}}-1,

with m,lm,l\in\mathbb{N}^{*}. If (3.38) is the case, thus

kπc+1=mπd+1\frac{k\pi}{\sqrt{c+1}}=\frac{m\pi}{\sqrt{d+1}}

so that

d=(c+1)m2k21.d=\frac{(c+1)m^{2}}{k^{2}}-1.

Therefore, if LdL\in\mathcal{R}_{d} then we have necessarily d𝒜22d\in\mathcal{A}_{2}\cup\mathcal{B}_{2} where

𝒜2:={(c+1)(m2+ml+l2)3k21;m,l}\mathcal{A}_{2}:=\left\{\frac{(c+1)(m^{2}+ml+l^{2})}{3k^{2}}-1;\ m,l\in\mathbb{N}^{*}\right\}

and

2:={(c+1)m2k21;m}.\mathcal{B}_{2}:=\left\{\frac{(c+1)m^{2}}{k^{2}}-1;\ m\in\mathbb{N}^{*}\right\}.

As done before, taking x,y𝒜2x,y\in\mathcal{A}_{2} with xyx\neq y, such that

x=(c+1)(m12+m1l1+l12)3k21x=\frac{(c+1)(m_{1}^{2}+m_{1}l_{1}+l_{1}^{2})}{3k^{2}}-1

and

y=(c+1)(m22+m2l2+l22)3k21,y=\frac{(c+1)(m_{2}^{2}+m_{2}l_{2}+l_{2}^{2})}{3k^{2}}-1,

where m1,l1,m2,l2m_{1},l_{1},m_{2},l_{2}\in\mathbb{N}^{*},yields that

xy=c+13k2[(m12+m1l1+l12)(m22+m2l2+l22)]x-y=\frac{c+1}{3k^{2}}\left[(m_{1}^{2}+m_{1}l_{1}+l_{1}^{2})-(m_{2}^{2}+m_{2}l_{2}+l_{2}^{2})\right]

and, as xyx\neq y we have

|(m12+m1l1+l12)(m22+m2l2+l22)|1.\left|(m_{1}^{2}+m_{1}l_{1}+l_{1}^{2})-(m_{2}^{2}+m_{2}l_{2}+l_{2}^{2})\right|\geq 1.

Consequently,

d(x,y)c+13k2,x,y𝒜2,xy.d(x,y)\geq\frac{c+1}{3k^{2}},\forall x,y\in\mathcal{A}_{2},\ x\neq y.

In an analogous way,

d(x,y)c+1k2,x,y2,xyd(x,y)\geq\frac{c+1}{k^{2}},\ \forall x,y\in\mathcal{B}_{2},\ x\neq y

and

d(x,y)c+13k2,x𝒜2,y2,xy.d(x,y)\geq\frac{c+1}{3k^{2}},\ \forall x\in\mathcal{A}_{2},\ y\in\mathcal{B}_{2},\ x\neq y.

Therefore,

d(x,y)c+13k2,x,y𝒜22,xy.d(x,y)\geq\frac{c+1}{3k^{2}},\ \forall x,y\in\mathcal{A}_{2}\cup\mathcal{B}_{2},\ x\neq y.

Proceeding as in the first case, we conclude that there exists ϵ2>0\epsilon_{2}>0 such that, for every d(cϵ2,c+ϵ2)\{c}d\in(c-\epsilon_{2},c+\epsilon_{2})\backslash\{c\} we have d𝒜22d\notin\mathcal{A}_{2}\cup\mathcal{B}_{2} so that LdL\notin\mathcal{R}_{d}. Considering ϵc:=min{ϵ1,ϵ2}\epsilon_{c}:=\min\{\epsilon_{1},\epsilon_{2}\}, thanks thanks to the [7, Proposition 3.6] and Theorem 1.2, the proof is completed.

3.2 Construction of the trajectories

In this subsection, for simplicity, we will consider c=0c=0 in c\mathcal{R}_{c}. Let T>0T>0 and note that by Theorem 1.6 there exist ϵ0>0\epsilon_{0}>0 such that for every d(0,ϵ0)d\in(0,\epsilon_{0}), the system (3.36) is exactly controllable around dd. Henceforth, we use C,C1,C2C,C_{1},C_{2} and C3C_{3} to denote the positive constants given in Proposition 2.6, corresponding to TT. One can see that, for every τ[0,T]\tau\in[0,T], the same constants can be used to apply the corresponding results for 𝒵τ\|\cdot\|_{\mathcal{Z}_{\tau}} so, for simplicity, we will consider now on τ=T3\tau=\frac{T}{3}.

The first result of this section ensures that we can construct solutions for the system (3.32) which starts close to 0 (left-hand side of the spatial domain) and achieves some non-null equilibrium in a certain time T/3T/3. The result is shown by a fixed-point argument.

Proposition 3.2.

There exist δ1>0\delta_{1}>0 such that, for every d(0,δ1)d\in(0,\delta_{1}) and y0L2(0,L)y_{0}\in L^{2}(0,L) with y0L2(0,L)<δ1,\|y_{0}\|_{L^{2}(0,L)}<\delta_{1}, there exists h1L2(0,T/3)h_{1}\in L^{2}(0,T/3) such that, the solution of (3.32) for t[0,T/3]t\in[0,T/3], satisfies

y(,T/3)=d.y(\cdot,T/3)=d.
Proof 3.3.

Let δ1(0,ϵ0)\delta_{1}\in(0,\epsilon_{0}) be a number to be chosen later. Consider d(0,δ1)d\in(0,\delta_{1}) and y0L2(0,L)y_{0}\in L^{2}(0,L) satisfying

y0L2(0,L)<δ1.\|y_{0}\|_{L^{2}(0,L)}<\delta_{1}.

For ε(0,ϵ0)\varepsilon\in(0,\epsilon_{0}) such that

C(τ12+τ13)ε𝒵T<δ1 and C(τ12+τ13)ε𝒵T2<δ1,\displaystyle C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|\varepsilon\|_{\mathcal{Z}_{T}}<\delta_{1}\text{ \ \ and \ \ }C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|\varepsilon\|_{\mathcal{Z}_{T}}^{2}<\delta_{1}, (3.39)

where C>0C>0 is the positive constant given in Lemma 2.3, [7, Proposition 3.6] guarantees the existence of a bounded linear operator

Ψε:L2(0,L)×L2(0,L)L2(0,τ)\Psi^{\varepsilon}:L^{2}(0,L)\times L^{2}(0,L)\rightarrow L^{2}(0,\tau)

such that, for any u0,uτL2(0,L)u_{0},u_{\tau}\in L^{2}(0,L), the solution uu of

{ut+(1+ε)ux+uxxx=0, in (0,L)×(0,τ),uxx(0,t)=0,ux(L,t)=h(t),uxx(L,t)=0, in (0,τ),u(x,0)=u0(x), in (0,L),\displaystyle\left\{\begin{array}[]{lll}u_{t}+(1+\varepsilon)u_{x}+u_{xxx}=0,&\text{ in }(0,L)\times(0,\tau),\\ u_{xx}(0,t)=0,\ u_{x}(L,t)=h(t),\ u_{xx}(L,t)=0,&\text{ in }(0,\tau),\\ u(x,0)=u_{0}(x),&\text{ in }(0,L),\end{array}\right.

with the control h=Ψε(u0,uτ)h=\Psi^{\varepsilon}(u_{0},u_{\tau}) satisfies u(,τ)=uτu(\cdot,\tau)=u_{\tau}.

We will denote, for simplicity, the operator Λ1+ε\Lambda_{1+\varepsilon} (given in Proposition 2.4 and Remark 2.9 with a=1+εa=1+\varepsilon) by Λε\Lambda_{\varepsilon}. Observe that, if yy is solution for (3.32) for some control hh, then yy is a solution of

{yt+(1+ε)yx+yxxx=yyx+εyx, in (0,L)×(0,τ),yxx(0,t)=0,yx(L,t)=h(t),yxx(L,t)=0, in (0,τ),y(x,0)=y0(x) in (0,L),\displaystyle\left\{\begin{array}[]{ll}y_{t}+(1+\varepsilon)y_{x}+y_{xxx}=-yy_{x}+\varepsilon y_{x},&\ \text{ in }(0,L)\times(0,\tau),\\ y_{xx}(0,t)=0,\ y_{x}(L,t)=h(t),\ y_{xx}(L,t)=0,&\ \text{ in }(0,\tau),\\ y(x,0)=y_{0}(x)&\ \text{ in }(0,L),\end{array}\right.

that is,

y=Λε(y0,h,yyx+εyx)=Λε(y0,h,0)+Λε(0,0,yyx+εyx).\displaystyle y=\Lambda_{\varepsilon}(y_{0},h,-yy_{x}+\varepsilon y_{x})=\Lambda_{\varepsilon}(y_{0},h,0)+\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x}). (3.40)

Let y𝒵τy\in\mathcal{Z}_{\tau} and hyL2(0,τ)h_{y}\in L^{2}(0,\tau) given by

hy=Ψε(y0,dΛε(0,0,yyx+εyx)(,τ)).h_{y}=\Psi^{\varepsilon}\big{(}y_{0},d-\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})(\cdot,\tau)\big{)}.

Define the map Γ:𝒵τ𝒵τ\Gamma:\mathcal{Z}_{\tau}\rightarrow\mathcal{Z}_{\tau} by

Γy=Λε(y0,hy,0)+Λε(0,0,yyx+εyx).\displaystyle\Gamma y=\Lambda_{\varepsilon}(y_{0},h_{y},0)+\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x}).

Note that if Γ\Gamma has a fixed point yy, then, from the above construction, it follows that yy is a solution of (3.32) with the control hyh_{y}. Moreover, from (3.40) we have

y=Λε(y0,hy,0)+Λε(0,0,yyx+εyx)\displaystyle y=\Lambda_{\varepsilon}(y_{0},h_{y},0)+\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})

so, by definitions of Λε,hy\Lambda_{\varepsilon},\ h_{y} and Ψε\Psi^{\varepsilon}, we get that

y(,τ)\displaystyle y(\cdot,\tau) =Λε(y0,hy,0)(,τ)+Λε(0,0,yyx+εyx)(,τ)\displaystyle=\Lambda_{\varepsilon}(y_{0},h_{y},0)(\cdot,\tau)+\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})(\cdot,\tau)
=dΛε(0,0,yyx+εyx)(,τ)+Λε(0,0,yyx+εyx)(,τ)\displaystyle=d-\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})(\cdot,\tau)+\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})(\cdot,\tau)
=d,\displaystyle=d,

and our problem would be solved. So we will focus our efforts on showing that Γ\Gamma has a fixed point in a suitable metric space.

To do that, let BB the set

B={y𝒵τ;y𝒵τr},\displaystyle B=\left\{y\in\mathcal{Z}_{\tau};\ \|y\|_{\mathcal{Z}_{\tau}}\leq r\right\},

with r>0r>0 to be chosen later. By (3.40) and Proposition 2.4 (together with Remark 2.9) we have, for yBy\in B, that

Γy𝒵τΛε(y0,hy,0)𝒵τ+Λε(0,0,yyx+εyx)𝒵τC2(y0L2(0,L)+hyL2(0,τ)+yyx+εyxL1(0,τ;L2(0,L))).\begin{split}\|\Gamma y\|_{\mathcal{Z}_{\tau}}&\leq\|\Lambda_{\varepsilon}(y_{0},h_{y},0)\|_{\mathcal{Z}_{\tau}}+\|\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})\|_{\mathcal{Z}_{\tau}}\\ &\leq C_{2}\left(\|y_{0}\|_{L^{2}(0,L)}+\|h_{y}\|_{L^{2}(0,\tau)}+\|-yy_{x}+\varepsilon y_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}\right).\end{split}

From Lemma 2.3 and Young inequality we ensure that

yyxεyxL1(0,τ,L2(0,L))\displaystyle\|yy_{x}-\varepsilon y_{x}\|_{L^{1}(0,\tau,L^{2}(0,L))} =(yε)yxL1(0,τ,L2(0,L))\displaystyle=\|(y-\varepsilon)y_{x}\|_{L^{1}(0,\tau,L^{2}(0,L))}
C(τ12+τ13)yε𝒵τy𝒵τ\displaystyle\leq C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|y-\varepsilon\|_{\mathcal{Z}_{\tau}}\|y\|_{\mathcal{Z}_{\tau}}
C(τ12+τ13)12(yε𝒵τ2+y𝒵τ2)\displaystyle\leq C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\frac{1}{2}\left(\|y-\varepsilon\|_{\mathcal{Z}_{\tau}}^{2}+\|y\|_{\mathcal{Z}_{\tau}}^{2}\right)
C(τ12+τ13)12(y𝒵τ2+y𝒵τ2+ε𝒵τ2+ε𝒵τ2+y𝒵τ2).\displaystyle\leq C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\frac{1}{2}\left(\|y\|_{\mathcal{Z}_{\tau}}^{2}+\|y\|_{\mathcal{Z}_{\tau}}^{2}+\|\varepsilon\|_{\mathcal{Z}_{\tau}}^{2}+\|\varepsilon\|_{\mathcal{Z}_{\tau}}^{2}+\|y\|_{\mathcal{Z}_{\tau}}^{2}\right).

Thus,

yyxεyxL1(0,τ,L2(0,L))\displaystyle\|yy_{x}-\varepsilon y_{x}\|_{L^{1}(0,\tau,L^{2}(0,L))} C(τ12+τ13)32y𝒵τ2+C(τ12+τ13)ε𝒵τ2\displaystyle\leq C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\frac{3}{2}\|y\|_{\mathcal{Z}_{\tau}}^{2}+C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|\varepsilon\|_{\mathcal{Z}_{\tau}}^{2}

and, by (3.39) we obtain

yyxεyxL1(0,τ,L2(0,L))\displaystyle\|yy_{x}-\varepsilon y_{x}\|_{L^{1}(0,\tau,L^{2}(0,L))} C(τ12+τ13)32y𝒵τ2+δ1,\displaystyle\leq C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\frac{3}{2}\|y\|_{\mathcal{Z}_{\tau}}^{2}+\delta_{1},

that is,

yyxεyxL1(0,τ,L2(0,L))\displaystyle\|yy_{x}-\varepsilon y_{x}\|_{L^{1}(0,\tau,L^{2}(0,L))} C¯y𝒵τ2+δ1,\displaystyle\leq\overline{C}\|y\|_{\mathcal{Z}_{\tau}}^{2}+\delta_{1},

where

C¯:=3C2(τ12+τ13).\displaystyle\overline{C}:=\frac{3C}{2}\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right).

In this way

hyL2(0,τ)\displaystyle\|h_{y}\|_{L^{2}(0,\tau)} =Ψε(y0,dΛε(0,0,yyx+εyx)(,τ))L2(0,τ)\displaystyle=\|\Psi^{\varepsilon}\big{(}y_{0},d-\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})(\cdot,\tau)\big{)}\|_{L^{2}(0,\tau)}
Ψε(y0L2(0,L)+dL2(0,L)+Λε(0,0,yyx+εyx)(,τ)L2(0,L))\displaystyle\leq\|\Psi^{\varepsilon}\|\left(\|y_{0}\|_{L^{2}(0,L)}+\|d\|_{L^{2}(0,L)}+\|\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})(\cdot,\tau)\|_{L^{2}(0,L)}\right)
Ψεδ1+ΨεdL+ΨεΛε(0,0,yyx+εyx)𝒵τ\displaystyle\leq\|\Psi^{\varepsilon}\|\delta_{1}+\|\Psi^{\varepsilon}\|d\sqrt{L}+\|\Psi^{\varepsilon}\|\|\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})\|_{\mathcal{Z}_{\tau}}
Ψεδ1+Ψεδ1L+ΨεC2yyx+εyxL1(0,τ;L2(0,L))\displaystyle\leq\|\Psi^{\varepsilon}\|\delta_{1}+\|\Psi^{\varepsilon}\|\delta_{1}\sqrt{L}+\|\Psi^{\varepsilon}\|C_{2}\|-yy_{x}+\varepsilon y_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
Ψεδ1+Ψεδ1L+ΨεC2C¯y𝒵τ2+ΨεC2δ1\displaystyle\leq\|\Psi^{\varepsilon}\|\delta_{1}+\|\Psi^{\varepsilon}\|\delta_{1}\sqrt{L}+\|\Psi^{\varepsilon}\|C_{2}\overline{C}\|y\|_{\mathcal{Z}_{\tau}}^{2}+\|\Psi^{\varepsilon}\|C_{2}\delta_{1}
=(1+L+C2)Ψεδ1+C2C¯Ψεr2.\displaystyle=\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|\delta_{1}+C_{2}\overline{C}\|\Psi^{\varepsilon}\|r^{2}.

Therefore,

Γy𝒵τ\displaystyle\|\Gamma y\|_{\mathcal{Z}_{\tau}} C2δ1+C2[(1+L+C2)Ψεδ1+C2C¯Ψεr2]+C2(C¯r2+δ1)\displaystyle\leq C_{2}\delta_{1}+C_{2}\left[\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|\delta_{1}+C_{2}\overline{C}\|\Psi^{\varepsilon}\|r^{2}\right]+C_{2}\left(\overline{C}r^{2}+\delta_{1}\right)
=[2C2+C2(1+L+C2)Ψε]δ1+(C22Ψε+C2)C¯r2.\displaystyle=\left[2C_{2}+C_{2}\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|\right]\delta_{1}+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\overline{C}r^{2}.

Choosing

r=2[2C2+C2(1+L+C2)Ψε]δ1\displaystyle r=2\left[2C_{2}+C_{2}\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|\right]\delta_{1}

and δ1\delta_{1} small enough such that

(C22Ψε+C2)C¯r<12, 2(C22Ψε+C2)r<12,(C22Ψε+C2)δ1<12,\displaystyle\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\overline{C}r<\frac{1}{2},\ \ \ 2\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)r<\frac{1}{2},\ \ \ \left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\delta_{1}<\frac{1}{2}, (3.41)

yields that

Γ𝒵τr2+r2=rΓ(B)B.\displaystyle\|\Gamma\|_{\mathcal{Z}_{\tau}}\leq\frac{r}{2}+\frac{r}{2}=r\implies\Gamma(B)\subset B.

Additionally, observe that for y,wBy,w\in B, Proposition 2.4 give us

ΓyΓw𝒵τ=\displaystyle\|\Gamma y-\Gamma w\|_{\mathcal{Z}_{\tau}}= Λε(0,hyhw,0)+Λε(0,0,yyx+wwx+εyxεwx)𝒵τ\displaystyle\|\Lambda_{\varepsilon}(0,h_{y}-h_{w},0)+\Lambda_{\varepsilon}(0,0,-yy_{x}+ww_{x}+\varepsilon y_{x}-\varepsilon w_{x})\|_{\mathcal{Z}_{\tau}}
\displaystyle\leq C2hyhwL2(0,τ)+C2yyxwwxL1(0,τ;L2(0,L))\displaystyle C_{2}\|h_{y}-h_{w}\|_{L^{2}(0,\tau)}+C_{2}\|yy_{x}-ww_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
+C2ε(yxwx)L1(0,τ;L2(0,L)).\displaystyle+C_{2}\|\varepsilon(y_{x}-w_{x})\|_{L^{1}(0,\tau;L^{2}(0,L))}.

Since

hyhw\displaystyle h_{y}-h_{w} =Ψε(0,Λε(0,0,yyx+εyx)(,τ)+Λε(0,0,wwx+εwx)(,τ))\displaystyle=\Psi^{\varepsilon}\left(0,-\Lambda_{\varepsilon}(0,0,-yy_{x}+\varepsilon y_{x})(\cdot,\tau)+\Lambda_{\varepsilon}(0,0,-ww_{x}+\varepsilon w_{x})(\cdot,\tau)\right)
=Ψε(0,Λε(yyxεyxwwx+εwx)(,τ)),\displaystyle=\Psi^{\varepsilon}\left(0,\Lambda_{\varepsilon}(yy_{x}-\varepsilon y_{x}-ww_{x}+\varepsilon w_{x})(\cdot,\tau)\right),

we have again from Proposition 2.4 that

C2hyhwL2(0,τ)\displaystyle C_{2}\|h_{y}-h_{w}\|_{L^{2}(0,\tau)}\leq C22Ψεyyxεyxwwx+εwxL1(0,τ;L2(0,L))\displaystyle C_{2}^{2}\|\Psi^{\varepsilon}\|\|yy_{x}-\varepsilon y_{x}-ww_{x}+\varepsilon w_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
\displaystyle\leq C22ΨεyyxwwxL1(0,τ;L2(0,L))\displaystyle C_{2}^{2}\|\Psi^{\varepsilon}\|\|yy_{x}-ww_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
+C22ΨεεyxεwxL1(0,τ;L2(0,L)).\displaystyle+C_{2}^{2}\|\Psi^{\varepsilon}\|\|\varepsilon y_{x}-\varepsilon w_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}.

Putting these two previous inequalities together, we find that

ΓyΓw𝒵τ\displaystyle\|\Gamma y-\Gamma w\|_{\mathcal{Z}_{\tau}}\leq (C22Ψε+C2)yyxwwxL1(0,τ;L2(0,L))\displaystyle\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\|yy_{x}-ww_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
+(C22Ψε+C2)ε(yxwx)L1(0,τ;L2(0,L)).\displaystyle+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\|\varepsilon(y_{x}-w_{x})\|_{L^{1}(0,\tau;L^{2}(0,L))}.

From Lemmas 2.3 and 2.8, together with the choices (3.39) and (3.41), it follows that

ΓyΓw𝒵τ\displaystyle\|\Gamma y-\Gamma w\|_{\mathcal{Z}_{\tau}}\leq 2(C22Ψε+C2)ryw𝒵τ\displaystyle 2\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)r\|y-w\|_{\mathcal{Z}_{\tau}}
+(C22Ψε+C2)C(τ12+τ13)ε𝒵τyw𝒵τ\displaystyle+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|\varepsilon\|_{\mathcal{Z}_{\tau}}\|y-w\|_{\mathcal{Z}_{\tau}}
\displaystyle\leq [2(C22Ψε+C2)r+(C22Ψε+C2)δ1]yw𝒵τ\displaystyle\left[2\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)r+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\delta_{1}\right]\|y-w\|_{\mathcal{Z}_{\tau}}
\displaystyle\leq yw𝒵τ\displaystyle\|y-w\|_{\mathcal{Z}_{\tau}}

Therefore, Γ:BB\Gamma:B\rightarrow B is a contraction so that, by Banach’s fixed point theorem, Γ\Gamma has a fixed point yBy\in B, concluding the proof.

The second result of this section ensures the construction of solutions for the system (3.32) on [2T/3,T][2T/3,T] starting in one non-null equilibrium and ending near 0.

Proposition 3.4.

There exists δ2>0\delta_{2}>0 such that, for every d(0,δ2)d\in(0,\delta_{2}) and yTL2(0,L)y_{T}\in L^{2}(0,L) satisfying yTL2(0,L)<δ2,\|y_{T}\|_{L^{2}(0,L)}<\delta_{2}, there exists h2L2(2T/3,T)h_{2}\in L^{2}(2T/3,T) such that, the solution of (1.9) for t[2T/3,T]t\in[2T/3,T] satisfies

y(,2T/3)=d and y(,T/3)=yT.y(\cdot,2T/3)=d\text{ \ \ and \ \ }y(\cdot,T/3)=y_{T}.
Proof 3.5.

Let δ2(0,ϵ0)\delta_{2}\in(0,\epsilon_{0}) be a number to be chosen later. Consider d(0,δ2)d\in(0,\delta_{2}) and yTL2(0,L)y_{T}\in L^{2}(0,L) satisfying

yTL2(0,L)<δ2.\|y_{T}\|_{L^{2}(0,L)}<\delta_{2}.

Let ε(0,ϵ0)\varepsilon\in(0,\epsilon_{0}) be such that

C(τ12+τ13)ε𝒵T<δ2 and C(τ12+τ13)ε𝒵T2<δ2,\displaystyle C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|\varepsilon\|_{\mathcal{Z}_{T}}<\delta_{2}\text{ \ \ and \ \ }C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|\varepsilon\|_{\mathcal{Z}_{T}}^{2}<\delta_{2}, (3.42)

where C>0C>0 is the positive constant given in Lemma 2.3. If zz is a solution to the problem

{zt+zx+zxxx+zzx=0, in (0,L)×(0,τ),zxx(0,t)=0,zx(L,t)=h(t),zxx(L,t)=0 in (0,τ),z(x,0)=d, in (0,L),\displaystyle\left\{\begin{array}[]{ll}z_{t}+z_{x}+z_{xxx}+zz_{x}=0,&\ \text{ in }(0,L)\times(0,\tau),\\ z_{xx}(0,t)=0,\ z_{x}(L,t)=h(t),\ z_{xx}(L,t)=0&\ \text{ in }(0,\tau),\\ z(x,0)=d,&\ \text{ in }(0,L),\end{array}\right. (3.46)

then zz is a solution of

{zt+(1+ε)zx+zxxx=zzx+εzx, in (0,L)×(0,τ),zxx(0,t)=0,zx(L,t)=h(t),zxx(L,t)=0, in (0,τ),z(x,0)=d, in (0,L),\displaystyle\left\{\begin{array}[]{ll}z_{t}+(1+\varepsilon)z_{x}+z_{xxx}=-zz_{x}+\varepsilon z_{x},&\ \text{ in }(0,L)\times(0,\tau),\\ z_{xx}(0,t)=0,\ z_{x}(L,t)=h(t),\ z_{xx}(L,t)=0,&\ \text{ in }(0,\tau),\\ z(x,0)=d,&\ \text{ in }(0,L),\end{array}\right.

that is,

z=Λε(d,h,zzx+εzx)=Λε(d,h,0)+Λε(0,0,zzx+εzx).\displaystyle z=\Lambda_{\varepsilon}(d,h,-zz_{x}+\varepsilon z_{x})=\Lambda_{\varepsilon}(d,h,0)+\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x}). (3.47)

Given z𝒵τz\in\mathcal{Z}_{\tau}, let hzL2(0,τ)h_{z}\in L^{2}(0,\tau) defined by

hz=Ψε(d,yTΛε(0,0,zzx+εzx)(,τ)).h_{z}=\Psi^{\varepsilon}\big{(}d,y_{T}-\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,\tau)\big{)}.

Now, consider the map Γ:𝒵τ𝒵τ\Gamma:\mathcal{Z}_{\tau}\rightarrow\mathcal{Z}_{\tau} given by

Γz=Λε(d,hz,0)+Λε(0,0,zzx+εzx).\displaystyle\Gamma z=\Lambda_{\varepsilon}(d,h_{z},0)+\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x}).

Once again, if Γ\Gamma has a fixed point zz, from the above construction, it follows that zz is a solution of (3.46) with the control hzh_{z}. Moreover, from (3.47) we have

z=Λε(d,hz,0)+Λε(0,0,zzx+εzx)\displaystyle z=\Lambda_{\varepsilon}(d,h_{z},0)+\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})

so, by definitions of Λε,hz\Lambda_{\varepsilon},\ h_{z} and Ψε\Psi^{\varepsilon}, it follows that

z(,0)\displaystyle z(\cdot,0) =Λε(d,hz,0)(,0)+Λε(0,0,zzx+εzx)(,0)=d+0=d\displaystyle=\Lambda_{\varepsilon}(d,h_{z},0)(\cdot,0)+\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,0)=d+0=d

and

z(,τ)\displaystyle z(\cdot,\tau) =Λε(d,hz,0)(,τ)+Λε(0,0,zzx+εzx)(,τ)\displaystyle=\Lambda_{\varepsilon}(d,h_{z},0)(\cdot,\tau)+\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,\tau)
=yTΛε(0,0,zzx+εzx)(,τ)+Λε(0,0,zzx+εzx)(,τ)\displaystyle=y_{T}-\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,\tau)+\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,\tau)
=yT.\displaystyle=y_{T}.

Hence our issue would be solved defining y:[0,L]×[2T/3,T]y:[0,L]\times[2T/3,T]\rightarrow\mathbb{R} by

y(x,t)=z(x,t2T/3).y(x,t)=z(x,t-2T/3).

Now ow, our focus is to show that Γ\Gamma has a fixed point in a suitable metric space. To do that, consider the set BB given by

B={z𝒵τ;z𝒵τr},\displaystyle B=\left\{z\in\mathcal{Z}_{\tau};\ \|z\|_{\mathcal{Z}_{\tau}}\leq r\right\},

with r>0r>0 to be chosen later. By (3.47) and Proposition 2.4 (together with Remark 2.9) we have, for zBz\in B, that

Γz𝒵τΛε(d,hz,0)𝒵τ+Λε(0,0,zzx+εzx)𝒵τC2(dL2(0,L)+hzL2(0,τ)+zzx+εzxL1(0,τ;L2(0,L))).\begin{split}\|\Gamma z\|_{\mathcal{Z}_{\tau}}&\leq\|\Lambda_{\varepsilon}(d,h_{z},0)\|_{\mathcal{Z}_{\tau}}+\|\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})\|_{\mathcal{Z}_{\tau}}\\ &\leq C_{2}\left(\|d\|_{L^{2}(0,L)}+\|h_{z}\|_{L^{2}(0,\tau)}+\|-zz_{x}+\varepsilon z_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}\right).\end{split}

As in the proof of the Proposition 3.2,

zzxεzxL1(0,τ,L2(0,L))\displaystyle\|zz_{x}-\varepsilon z_{x}\|_{L^{1}(0,\tau,L^{2}(0,L))} C¯z𝒵τ2+δ2.\displaystyle\leq\overline{C}\|z\|_{\mathcal{Z}_{\tau}}^{2}+\delta_{2}.

Moreover,

hzL2(0,τ)\displaystyle\|h_{z}\|_{L^{2}(0,\tau)} =Ψε(d,yTΛε(0,0,zzx+εzx)(,τ))L2(0,τ)\displaystyle=\|\Psi^{\varepsilon}\big{(}d,y_{T}-\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,\tau)\big{)}\|_{L^{2}(0,\tau)}
Ψε(dL2(0,L)+yTL2(0,L)+Λε(0,0,zzx+εzx)(,τ)L2(0,L))\displaystyle\leq\|\Psi^{\varepsilon}\|\left(\|d\|_{L^{2}(0,L)}+\|y_{T}\|_{L^{2}(0,L)}+\|\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,\tau)\|_{L^{2}(0,L)}\right)
ΨεdL+Ψεδ2+ΨεΛε(0,0,zzx+εzx)𝒵τ\displaystyle\leq\|\Psi^{\varepsilon}\|d\sqrt{L}+\|\Psi^{\varepsilon}\|\delta_{2}+\|\Psi^{\varepsilon}\|\|\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})\|_{\mathcal{Z}_{\tau}}
Ψεδ2+Ψεδ2L+ΨεC2zzx+εzxL1(0,τ;L2(0,L))\displaystyle\leq\|\Psi^{\varepsilon}\|\delta_{2}+\|\Psi^{\varepsilon}\|\delta_{2}\sqrt{L}+\|\Psi^{\varepsilon}\|C_{2}\|-zz_{x}+\varepsilon z_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
Ψεδ2+Ψεδ2L+ΨεC2C¯z𝒵τ2+ΨεC2δ2\displaystyle\leq\|\Psi^{\varepsilon}\|\delta_{2}+\|\Psi^{\varepsilon}\|\delta_{2}\sqrt{L}+\|\Psi^{\varepsilon}\|C_{2}\overline{C}\|z\|_{\mathcal{Z}_{\tau}}^{2}+\|\Psi^{\varepsilon}\|C_{2}\delta_{2}
=(1+L+C2)Ψεδ2+C2C¯Ψεr2.\displaystyle=\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|\delta_{2}+C_{2}\overline{C}\|\Psi^{\varepsilon}\|r^{2}.

Therefore,

Γz𝒵τ\displaystyle\|\Gamma z\|_{\mathcal{Z}_{\tau}} C2δ2L+C2[(1+L+C2)Ψεδ2+C2C¯Ψεr2]+C2(C¯r2+δ2)\displaystyle\leq C_{2}\delta_{2}\sqrt{L}+C_{2}\left[\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|\delta_{2}+C_{2}\overline{C}\|\Psi^{\varepsilon}\|r^{2}\right]+C_{2}\left(\overline{C}r^{2}+\delta_{2}\right)
=[C2L+C2(1+L+C2)Ψε+C2]δ2+(C22Ψε+C2)C¯r2.\displaystyle=\left[C_{2}\sqrt{L}+C_{2}\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|+C_{2}\right]\delta_{2}+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\overline{C}r^{2}.

Choosing

r=2[C2L+C2(1+L+C2)Ψε+C2]δ2\displaystyle r=2\left[C_{2}\sqrt{L}+C_{2}\left(1+\sqrt{L}+C_{2}\right)\|\Psi^{\varepsilon}\|+C_{2}\right]\delta_{2}

and δ2\delta_{2} small enough such that

(C22Ψε+C2)C¯r<12, 2(C22Ψε+C2)r<12,(C22Ψε+C2)δ1<12,\displaystyle\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\overline{C}r<\frac{1}{2},\ \ \ 2\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)r<\frac{1}{2},\ \ \ \left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\delta_{1}<\frac{1}{2}, (3.48)

we get that

Γz𝒵τr2+r2=rΓ(B)B\displaystyle\|\Gamma z\|_{\mathcal{Z}_{\tau}}\leq\frac{r}{2}+\frac{r}{2}=r\implies\Gamma(B)\subset B

Furthermore, observe that for z,wBz,w\in B, Proposition 2.4 give us

ΓzΓw𝒵τ=\displaystyle\|\Gamma z-\Gamma w\|_{\mathcal{Z}_{\tau}}= Λε(0,hzhw,0)+Λε(0,0,zzx+wwx+εzxεwx)𝒵τ\displaystyle\|\Lambda_{\varepsilon}(0,h_{z}-h_{w},0)+\Lambda_{\varepsilon}(0,0,-zz_{x}+ww_{x}+\varepsilon z_{x}-\varepsilon w_{x})\|_{\mathcal{Z}_{\tau}}
\displaystyle\leq C2hzhwL2(0,τ)+C2zzxwwxL1(0,τ;L2(0,L))\displaystyle C_{2}\|h_{z}-h_{w}\|_{L^{2}(0,\tau)}+C_{2}\|zz_{x}-ww_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
+C2ε(zxwx)L1(0,τ;L2(0,L)).\displaystyle+C_{2}\|\varepsilon(z_{x}-w_{x})\|_{L^{1}(0,\tau;L^{2}(0,L))}.

Since

hzhw\displaystyle h_{z}-h_{w} =Ψε(0,Λε(0,0,zzx+εzx)(,τ)+Λε(0,0,wwx+εwx)(,τ))\displaystyle=\Psi^{\varepsilon}\left(0,-\Lambda_{\varepsilon}(0,0,-zz_{x}+\varepsilon z_{x})(\cdot,\tau)+\Lambda_{\varepsilon}(0,0,-ww_{x}+\varepsilon w_{x})(\cdot,\tau)\right)
=Ψε(0,Λε(zzxεzxwwx+εwx)(,τ)),\displaystyle=\Psi^{\varepsilon}\left(0,\Lambda_{\varepsilon}(zz_{x}-\varepsilon z_{x}-ww_{x}+\varepsilon w_{x})(\cdot,\tau)\right),

we have, again from Proposition 2.4, that

C2hzhwL2(0,τ)\displaystyle C_{2}\|h_{z}-h_{w}\|_{L^{2}(0,\tau)}\leq C22Ψεzzxεzxwwx+εwxL1(0,τ;L2(0,L))\displaystyle C_{2}^{2}\|\Psi^{\varepsilon}\|\|zz_{x}-\varepsilon z_{x}-ww_{x}+\varepsilon w_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
\displaystyle\leq C22ΨεzzxwwxL1(0,τ;L2(0,L))\displaystyle C_{2}^{2}\|\Psi^{\varepsilon}\|\|zz_{x}-ww_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
+C22ΨεεzxεwxL1(0,τ;L2(0,L)).\displaystyle+C_{2}^{2}\|\Psi^{\varepsilon}\|\|\varepsilon z_{x}-\varepsilon w_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}.

Then,

ΓzΓw𝒵τ\displaystyle\|\Gamma z-\Gamma w\|_{\mathcal{Z}_{\tau}}\leq (C22Ψε+C2)zzxwwxL1(0,τ;L2(0,L))\displaystyle\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\|zz_{x}-ww_{x}\|_{L^{1}(0,\tau;L^{2}(0,L))}
+(C22Ψε+C2)ε(zxwx)L1(0,τ;L2(0,L)).\displaystyle+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\|\varepsilon(z_{x}-w_{x})\|_{L^{1}(0,\tau;L^{2}(0,L))}.

From Lemmas 2.3 and 2.8, together with (3.42) and (3.48), it follows that

ΓzΓw𝒵τ\displaystyle\|\Gamma z-\Gamma w\|_{\mathcal{Z}_{\tau}}\leq 2(C22Ψε+C2)rzw𝒵τ\displaystyle 2\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)r\|z-w\|_{\mathcal{Z}_{\tau}}
+(C22Ψε+C2)C(τ12+τ13)ε𝒵τzw𝒵τ\displaystyle+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)C\left(\tau^{\frac{1}{2}}+\tau^{\frac{1}{3}}\right)\|\varepsilon\|_{\mathcal{Z}_{\tau}}\|z-w\|_{\mathcal{Z}_{\tau}}
\displaystyle\leq [2(C22Ψε+C2)r+(C22Ψε+C2)δ2]zw𝒵τ\displaystyle\left[2\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)r+\left(C_{2}^{2}\|\Psi^{\varepsilon}\|+C_{2}\right)\delta_{2}\right]\|z-w\|_{\mathcal{Z}_{\tau}}
\displaystyle\leq zw𝒵τ.\displaystyle\|z-w\|_{\mathcal{Z}_{\tau}}.

Therefore, Γ:BB\Gamma:B\rightarrow B is a contraction so that, by Banach’s fixed point theorem, Γ\Gamma has a fixed point zBz\in B which concludes our proof.

3.3 Controllability on c\mathcal{R}_{c}

We are in a position to prove Theorems 1.4 and 1.5. For the sake of simplicity, we will give the proof of the case L0L\in\mathcal{R}_{0} (Theorem 1.4), and the case LcL\in\mathcal{R}_{c} (Theorem 1.5) follows similarly.

Proof 3.6.

(Proof of Theorem 1.4.) Let δ1\delta_{1} and δ2\delta_{2} be positive real numbers given in Propositions 3.2 and 3.4, respectively. Define δ:=min{δ1,δ2}\delta:=\min\{\delta_{1},\delta_{2}\} and consider d(0,δ)d\in(0,\delta) and y0,yTL2(0,L)y_{0},y_{T}\in L^{2}(0,L) such that

y0L2(0,L),yTL2(0,L)<δ.\displaystyle\|y_{0}\|_{L^{2}(0,L)},\|y_{T}\|_{L^{2}(0,L)}<\delta.

From Proposition 3.2 there exists h1L2(0,T/3)h_{1}\in L^{2}(0,T/3) such that, the solution y1𝒵T/3y_{1}\in\mathcal{Z}_{T/3} of

{yt+yx+yxxx+yyx=0, in (0,L)×(0,T/3),yxx(0,t)=0,yx(L,t)=h1(t),yxx(L,t)=0, in (0,T/3),y(x,0)=y0(x), in (0,L),\displaystyle\left\{\begin{array}[]{ll}y_{t}+y_{x}+y_{xxx}+yy_{x}=0,&\ \text{ in }(0,L)\times(0,T/3),\\ y_{xx}(0,t)=0,\ y_{x}(L,t)=h_{1}(t),\ y_{xx}(L,t)=0,&\ \text{ in }(0,T/3),\\ y(x,0)=y_{0}(x),&\ \text{ in }(0,L),\end{array}\right.

satisfies

y1(x,T/3)=d.\displaystyle y_{1}(x,T/3)=d.

On the other hand, thanks to the Proposition 3.4, there exists h2L2(2T/3,T)h_{2}\in L^{2}(2T/3,T) such that, the solution y2𝒵2T/3,Ty_{2}\in\mathcal{Z}_{2T/3,T} of

{yt+yx+yxxx+yyx=0, in (0,L)×(2T/3,T),yxx(0,t)=0,yx(L,t)=h2(t),yxx(L,t)=0, in (2T/3,T),y(x,2T/3)=d, in (0,L),\displaystyle\left\{\begin{array}[]{ll}y_{t}+y_{x}+y_{xxx}+yy_{x}=0,&\ \text{ in }(0,L)\times(2T/3,T),\\ y_{xx}(0,t)=0,\ y_{x}(L,t)=h_{2}(t),\ y_{xx}(L,t)=0,&\ \text{ in }(2T/3,T),\\ y(x,2T/3)=d,&\ \text{ in }(0,L),\end{array}\right.

satisfies

y2(x,T)=yT.\displaystyle y_{2}(x,T)=y_{T}.

Defining y:[0,L]×[0,T]y:[0,L]\times[0,T]\rightarrow\mathbb{R} by

y={y1, in [0,T/3],d, in [T/3,2T/3],y2, in [2T/3,T],\displaystyle y=\left\{\begin{array}[]{ll}y_{1},&\text{ in }[0,T/3],\\ d,&\text{ in }[T/3,2T/3],\\ y_{2},&\text{ in }[2T/3,T],\end{array}\right. (3.52)

we have that y𝒵Ty\in\mathcal{Z}_{T} and yy is solution of (3.32) driving y0y_{0} to yTy_{T} at time TT, showing that the system (3.32) is exactly controllable, and the proof is completed.

Acknowledgment

Capistrano-Filho was supported by CAPES grant number 88881.311964/2018-01, COFECUB/CAPES grant number 8887.879175/2023-00, CNPq grants numbers 307808/2021-1 and 401003/2022-1. Da Silva acknowledges support from CNPq. This work is part of the Ph.D. thesis of da Silva at the Department of Mathematics of the Universidade Federal de Pernambuco.

References

  • [1] B. Alvarez-Samaniego and D. Lannes, Large time existence for 3D water-waves and asymptotics, Invent. Math. 171 (2008), no. 3, 485–541.
  • [2] J. L. Bona, M. Chen, and J.-C. Saut, Boussinesq equations and other systems for small-amplitude long waves in nonlinear dispersive media. I: Derivation and linear theory, J. Nonlinear. Sci. Vol. 12 (2002), 283–318.
  • [3] J. L. Bona, T. Colin, and D. Lannes, Long wave approximations for water waves, Arch. Ration. Mech. Anal. 178 (2005), no. 3, 373–410.
  • [4] J. L. Bona, D. Lannes, and J.-C. Saut, Asymptotic models for internal waves, J. Math. Pures Appl. (9) 89 (2008), no. 6, 538–566.
  • [5] J. J. Bona, S.M. Sun, and B.-Y. Zhang, A nonhomogeneous boundary-value-problem for the Korteweg-de Vries equation posed on a finite domain, Comm. Partial Differential Equations, 28 (2003), 1391–1438.
  • [6] J. M. Boussinesq, Thórie de l’intumescence liquide, applelée onde solitaire ou de, translation, se propageant dans un canal rectangulaire, C. R. Acad. Sci. Paris. 72 (1871) 755–759.
  • [7] M. A. Caicedo, R. A. Capistrano–Filho, and B.-Y Zhang, Neumann boundary controllability of the Korteweg–de Vries equation on a bounded domain, SIAM J. Control Optim., 55 (2017), 3503–3532.
  • [8] E. Cerpa, Exact controllability of a nonlinear Korteweg-de Vries equation on a critical spatial domain, SIAM J. Control Optim., 46 (2007), 877–899.
  • [9] E. Cerpa, Control of a Korteweg-de Vries equation: a tutorial, Math. Control Relat. Fields, 4 (2014), 45–99.
  • [10] E. Cerpa and E. Crépeau, Boundary controllability for the nonlinear Korteweg-de Vries equation on any critical domain, Ann. I.H. Poincaré, 26 (2009), 457–475.
  • [11] J.-M. Coron and E. Crépeau, Exact boundary controllability of a nonlinear KdV equation with critical lengths. Journal of the European Mathematical Society, no. 6 (2004), 367-398.
  • [12] J.-M. Coron, Global asymptotic stabilization for controllable systems without drift, Math. Control. Signals System, 5 (1992), 295–312.
  • [13] J.-M. Coron, Contrôlabilité exacte frontiére de l’équation d’Euler des fluids parfaits incompressibles bidimensionnels. C. R. Acad. Sci. Paris, 317 (1993), 271–276.
  • [14] J.-M. Coron, Control and Nonlinearity, Math. Surveys Monogr. 136, American Mathematical Society, Providence, 2007.
  • [15] E. Crépeau, Exact boundary controllability of the Korteweg–de Vries equation around a non-trivial stationary solution, International Journal of Control, 74:11 (2001), 1096–1106.
  • [16] O. Glass, Controllability and asymptotic stabilization of the Camassa-Holm equation, Journal of Differential Equations, 245 (2008), 1584–1615.
  • [17] O. Glass and S. Guerrero, Some exact controllability results for the linear KdV equation and uniform controllability in the zero-dispersion limit, Asymptot. Anal., 60 (2008), 61–100.
  • [18] O. Glass and S. Guerrero, Controllability of the Korteweg-de Vries equation from the right Dirichlet boundary condition, Systems Control Lett., 59 (2010), 390–395.
  • [19] J.-P. Guilleron, Null controllability of a linear KdV equation on an interval with special boundary conditions, Math. Control Signals Syst., 26 (2014), 375–401.
  • [20] D. J. Korteweg and G. de Vries, On the change of form of long waves advancing in a rectangular canal, and on a new type of long stationary waves, Phil. Mag. 39 (1895), 422–443.
  • [21] E. Kramer, I. Rivas, and B.-Y. Zhang, Well-posedness of a class of initial-boundary-problem for the Korteweg-de Vries equation on a bounded domain, ESAIM Control Optim. Calc. Var., 19 (2013), 358–384.
  • [22] E. Kramer and B.-Y. Zhang, Nonhomogeneous boundary value problems for the Korteweg-de Vries equation on a bounded domain, J. Syst. Sci. Complex, 23 (2010) 499–526.
  • [23] D. Lannes, The water waves problem. Mathematical analysis and asymptotics. Mathematical Surveys and Monographs, 188. American Mathematical Society, Providence, RI, 2013. xx+32.
  • [24] L. Rosier, Exact boundary controllability for the Korteweg-de Vries equation on a bounded domain, ESAIM Control Optim. Cal. Var. 2 (1997), 33–55.
  • [25] L. Rosier and B.-Y. Zhang, Control and stabilization of the Korteweg-de Vries equation: Recent progress, J. Syst Sci &\& Complexity, 22 (2009), 647–682.
  • [26] J.-C. Saut and B. Scheurer, Unique continuation for some evolution equations. J. Diff. Eqs. 66 (1987), 118–139.