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

Bilinear optimal stabilization of a non-homogeneous Fokker-Planck equation

K. Ammari dUR Analysis and Control of PDE’s
Faculty of Sciences of Monastir, University of Monastir
5019 Monastir, Tunisia
[email protected]
   M. Ouzahra Laboratory MMPA, Department of mathematics &\& informatics,
ENS. University of Sidi Mohamed Ben Abdellah
Fes, Morocco
[email protected]
   S. Yahyaoui Laboratory MMPA, Department of mathematics &\& informatics,
ENS. University of Sidi Mohamed Ben Abdellah
Fes, Morocco
[email protected]
Abstract

In this work, we study the bilinear optimal stabilization of a non-homogeneous Fokker-Planck equation. We first study the problem of optimal control in a finite-time interval and then focus on the case of the infinite time horizon. We further show that the obtained optimal control guarantees the strong stability of the system at hand. An illustrating numerical example is given.

Index Terms:
Quadratic cost, optimal control, feedback stabilization, bilinear systems, Fokker-Planck equation
publicationid: pubid: 978-1-6654-8724-5/22/$31.00 ©2022 IEEE

I Introduction

In this paper we consider the following non-homogeneous bilinear system:

{y˙(t,x)=Δy(t,x)+i=1Nui(t)(yxi(t,x)+bi),(t,x)(0,T)×Ωy(t,x)=0,(t,x)(0,T)×Ωy(0,x)=y0L2(Ω),xΩ\begin{cases}\dot{y}(t,x)=\Delta y(t,x)+\sum_{i=1}^{N}u_{i}(t)\left(\frac{\partial y}{\partial x_{i}}(t,x)+b_{i}\right),(t,x)\in(0,T)\times\Omega\\ y(t,x)=0,\,\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (t,x)\in(0,T)\times\partial\Omega\\ y(0,x)=y_{0}\in L^{2}(\Omega),\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ x\in\Omega\end{cases} (1)

where 0<T+0<T\leq+\infty, Ω\Omega is an open and bounded domain of N\mathbb{R}^{N}, N{1,2,3}N\in\{1,2,3\} and biH02(Ω):=H01(Ω)H2(Ω)b_{i}\in H_{0}^{2}(\Omega):=H_{0}^{1}(\Omega)\cap H^{2}(\Omega). Here, ui(t)u_{i}(t) design the controls and y(t)y(t) the corresponding mild solution of the system (1).

In term of applications, equation like (1) may for instance describe the situation where some physical quantities (particles, energy,…) are transferred inside a system due to diffusion and convection processes, and the control can be seen as the velocity field that the quantity is moving with. For our special case, system (1) describes a deterministic Fokker–Planck equation for the time-dependent probability density P(A,t){P\left(A,t\right)} of a stochastic variable AA of the Langevin equation, which enables us to study many types of fluctuations in physical and biological systems (see e.g. [19]).

The goal of this paper is to study the problem of stability by an optimal control in the infinite time-horizon for the non-homogeneous bilinear system (1). The main difficulty in solving a quadratic optimal control for general bilinear systems is the non-convexity of the cost function. In the case of a bounded control operator, the question of bilinear optimal control problem has been widely studied in the literature (see [6, 11, 15, 20, 27, 28]). However, the modeling may give rise to the unboundedness aspect of the operator of control of the obtained bilinear model (see [1, 3, 7, 14, 16]), which is the case of equation (1) where the control is acting in the coefficient of the divergence term. In [1], the authors have considered the homogeneous version of the system (1) (i.e. bi=0,i=1,..,Nb_{i}=0,\;i=1,..,N), for which they characterized the optimal control for a finite time horizon. Moreover, the author in [3, 14] has studied the same problem in the presence of a time and state-dependent perturbation for finite time horizon.

The paper is organized as follows: In Section II we give a preliminary. In Section III, we first solve the optimal control problem in a finite time-horizon for the system (1), and then proceed to the case of infinite time horizon in which context we give a stabilization result by optimal control. Finally, in Section 4, we present a numerical example.

II Setting of the problem and some a priori estimates

Let us consider the following spaces: H=L2(Ω),V=H01(Ω),H=L^{2}(\Omega),\;V=H_{0}^{1}(\Omega), V=H1(Ω)V^{*}=H^{-1}(\Omega) and U=L2(0,T;N)U=L^{2}(0,T;\mathbb{R}^{N}), and let us introduce the following operators :

  • A:VV,yΔyA:V\rightarrow V^{*},\ y\mapsto\Delta y which is a linear continuous operator,

  • the linear continuous operator B:VVB:V\rightarrow V^{*}, is defined by By=1Ω.y=i=1N,yxiBy=1_{\Omega}.\nabla y=\sum_{i=1}^{N},\frac{\partial y}{\partial x_{i}}, here 1Ω1_{\Omega} is the vector (1,,1)(1,\cdots,1).

For all u:=(ui)1iNUu:=(u_{i})_{1\leq i\leq N}\in U and yVy\in V, we have

u.(By+b):=u.(y+b)=i=1Nyi(yxi+bi)u.(By+b):=u.(\nabla y+b)=\sum_{i=1}^{N}y_{i}\,\left(\frac{\partial y}{\partial x_{i}}+b_{i}\right)\cdot

Thus the system (1) can be rewritten in the form

{y˙(t)=Ay(t)+u(t).(By(t)+b)y(0)=y0H.\begin{cases}\dot{y}(t)=Ay(t)+u(t).(By(t)+b)\\ y(0)=y_{0}\in H.\end{cases} (2)

The quadratic cost function JJ to be minimized is defined by

J(u)=0Ty(t)H2𝑑t+r20Tu(t)N2𝑑t,J(u)=\int_{0}^{T}\|y(t)\|_{H}^{2}dt+\frac{r}{2}\int_{0}^{T}\|u(t)\|_{\mathbb{R}^{N}}^{2}dt, (3)

where r>0r>0, uUu\in U and y(t)y(t) is the respective solution to system (2) .

Then, the optimal control problem may be stated as follows

{minJ(u)uU\left\{\begin{array}[]{ll}\min J(u)\\ u\in U\\ \end{array}\right. (4)

For the wellposedness of the system (2), let us consider the following system

{y˙(t)=Ay(t)+u(t).By(t)+f(t)y(0)=y0\begin{cases}\dot{y}(t)=Ay(t)+u(t).By(t)+f(t)\\ y(0)=y_{0}\end{cases} (5)

where fL2(0,T;V)f\in L^{2}(0,T;V^{*}), and let us introduce the following functional space

W(0,T)={ϕL2(0,T,V)/ϕ˙L2(0,T;V)}W(0,T)=\{\phi\in L^{2}(0,T,V)\ \ /\ \ \dot{\phi}\in L^{2}(0,T;V^{*})\}\cdot

Now, we recall the following existence result with some a priory estimates (see [1, 3, 14]).

Lemma 1

For all uUu\in U, there exists a unique solution yy of the system (5), which is such that

yW(0,T)L(0,T;H)y\in W(0,T)\cap L^{\infty}(0,T;H)\cdot

Moreover, the following estimates hold

yL2(0,T;V)12y0H+fL2(0,T;V)\|y\|_{L^{2}(0,T;V)}\leq\frac{1}{\sqrt{2}}\|y_{0}\|_{H}+\|f\|_{L^{2}(0,T;V^{*})} (6)
yL(0,T;H)y0H+2fL2(0,T;V)\|y\|_{L^{\infty}(0,T;H)}\leq\|y_{0}\|_{H}+\sqrt{2}\|f\|_{L^{2}(0,T;V^{*})} (7)
y˙L(0,T;V)\displaystyle\|\dot{y}\|_{L^{(}0,T;V^{*})}\leq (12y0H+fL2(0,T;V))(α+2uU)\displaystyle\left(\frac{1}{\sqrt{2}}\|y_{0}\|_{H}+\|f\|_{L^{2}(0,T;V^{*})}\right)\left(\alpha+\sqrt{2}\|u\|_{U}\right) (8)
+fL2(0,T;V)\displaystyle\ +\|f\|_{L^{2}(0,T;V^{*})}

where α\alpha is such that AzVαzV\|Az\|_{V^{*}}\leq\alpha\|z\|_{V}, for all zVz\in V.

III Characterization of the optimal control

III-A Existence of an optimal control

Theorem 2

For any y0Hy_{0}\in H, the problem (4) has at least one solution.

Proof 1

Since the set {J(u)/uU}\{J(u)/u\in U\} is not empty and is bounded from below, it admits a lower bound JJ^{*}.
Let (un)n(u_{n})_{n\in\mathbb{N}} be a minimizing sequence such that J(un)JJ(u_{n})\rightarrow J^{*}.
Then the sequence (un)(u_{n}) is bounded, so it admits a sub-sequence denoted by (un)(u_{n}) as well, which weakly converges to uUu^{*}\in U.
Let (yn)(y_{n}) be the sequence of solutions of (2) corresponding to (un)(u_{n}). According to Lemma 1, the sequences yn(0)H,ynL2(0,T;V),ynL(0,T;V),y˙nL2(0,T;V),\begin{aligned} \|y_{n}(0)\|_{H},\|y_{n}\|_{L^{2}(0,T;V)},\ \|y_{n}\|_{L^{\infty}(0,T;V)},\ \|\dot{y}_{n}\|_{L^{2}(0,T;V^{*})},\end{aligned} AynL2(0,T;V)\begin{aligned} \|Ay_{n}\|_{L^{2}(0,T;V^{*})}\end{aligned} and un.(Byn+b)L2(0,T;V)\|u_{n}.(By_{n}+b)\|_{L^{2}(0,T;V^{*})} are bounded, so (yn)(y_{n}) admits a sub-sequence, also denoted by (yn)(y_{n}), such that

ynyweaklyinL2(0,T;V),y_{n}\rightharpoonup y^{*}\ weakly\ in\ L^{2}(0,T;V),
ynyweaklyinL(0,T;H),y_{n}\rightharpoonup y^{*}\ weakly*\ \ in\ L^{\infty}(0,T;H),
y˙ny˙weaklyinL2(0,T;V).\dot{y}_{n}\rightharpoonup\dot{y}^{*}\ weakly\ in\ L^{2}(0,T;V^{*}).

In addition to this, the linear operator

𝔸:L2(0,T;V)L2(0,T;V)\mathbb{A}:L^{2}(0,T;V)\rightarrow L^{2}(0,T;V^{*})
y𝔸yy\mapsto\mathbb{A}y

is continuous, from which it follows that

𝔸yn𝔸yweaklyinL2(0,T;V)\mathbb{A}y_{n}\rightharpoonup\mathbb{A}y^{*}\ weakly\ in\ L^{2}(0,T;V^{*})\cdot

Then since the embedding W(0,T)L2(0,T;H)W(0,T)\rightarrow L^{2}(0,T;H) is compact, (yn)(y_{n}) admits a sub-sequence, still denoted by (yn)(y_{n}), for which we have

ynystronglyinL2(0,T;H).y_{n}\rightarrow y^{*}\ strongly\ in\ L^{2}(0,T;H). (9)

Taking into account that the operator 𝔹:L2(0,T;H)L2(0,T;V)\mathbb{B}:L^{2}(0,T;H)\rightarrow L^{2}(0,T;V^{*}) is linear and continuous, we deduce that

un.(𝔹yn+b)u.(𝔹y+b)weaklyinL2(0,T;V).u_{n}.(\mathbb{B}y_{n}+b)\rightharpoonup u^{*}.(\mathbb{B}y^{*}+b)\ \ weakly\ in\ L^{2}(0,T;V^{*}).

Now, by taking the limit we deduce that

{y˙(t)=Ay(t)+u(t).(By(t)+b)y(0)=y0\begin{cases}\dot{y^{*}}(t)=Ay^{*}(t)+u^{*}(t).(By^{*}(t)+b)\\ y^{*}(0)=y_{0}\end{cases}

In other words, yy^{*} is the solution of the system (2) corresponding to control u=uu=u^{*}.
Using that the norm L2(H)\|\cdot\|_{L^{2}(H)} is lower semi-continuous, it follows from the strong convergence of the sequence yny_{n} to yy^{*} in L2(0,T;H)L^{2}(0,T;H) that

0Ty(t)H2𝑑tlim infn+0Tyn(t)H2𝑑t.\int_{0}^{T}\|y^{*}(t)\|_{H}^{2}dt\leq\liminf_{n\rightarrow+\infty}\int_{0}^{T}\|y_{n}(t)\|_{H}^{2}dt. (10)

Since R:u0Tu(t)U2𝑑tR:u\mapsto\int_{0}^{T}\|u(t)\|_{U}^{2}dt is convex and lower semi-continuous with respect to the weak topology, we have (see Corollary III.8 of [13])

R(u)lim infn+R(un)R(u^{*})\leq\liminf_{n\rightarrow+\infty}R(u_{n})\cdot (11)

Combining the formulas (10) and (11) we deduce that

J(u)\displaystyle J(u^{*}) =0Ty(t)H2𝑑t+r20Tu(t)N2𝑑t\displaystyle=\int_{0}^{T}\|y^{*}(t)\|_{H}^{2}dt+\frac{r}{2}\int_{0}^{T}\|u^{*}(t)\|_{\mathbb{R}^{N}}^{2}dt
lim infn+0Tyn(t)H2𝑑t+r2lim infn+0Tun(t)N2𝑑t\displaystyle\leq\liminf_{n\rightarrow+\infty}\int_{0}^{T}\|y_{n}(t)\|_{H}^{2}dt+\frac{r}{2}\liminf_{n\rightarrow+\infty}\int_{0}^{T}\|u_{n}(t)\|_{\mathbb{R}^{N}}^{2}dt
lim infn+J(un)\displaystyle\leq\liminf_{n\rightarrow+\infty}J(u_{n})
J\displaystyle\leq J^{*}\cdot

We conclude that J(u)=JJ(u^{*})=J^{*}, and so uu^{*} is a solution of the problem (4).

III-B Expression of the optimal control for finite time-horizon

In this subsection, we will provide informations about the optimal control.

Theorem 3

For all T>0T>0, the problem (4) admits a solution uu^{*} which is given by:

ui(t)=1rϕ(t),y(t)xi+biV,V,i=1,,N,u_{i}^{*}(t)=-\dfrac{1}{r}\langle\phi(t),\frac{\partial y^{*}(t)}{\partial x_{i}}+b_{i}\rangle_{V^{*},V},\,\forall\,i=1,...,N,

where yy^{*} is the solution of the system (2) corresponding to uu^{*} and ϕ\phi is the solution of the following adjoint system

{ϕ˙(t)=Aϕ(t)+u(t).Bϕ(t)2y(t)ϕ(T)=0\left\{\begin{array}[]{ll}\dot{\phi}(t)=-A\phi(t)+u^{*}(t).B\phi(t)-2y^{*}(t)\\ \phi(T)=0\end{array}\right. (12)
Proof 2

First, let us show that the mapping

UC(0,T;H)U\rightarrow C(0,T;H)
uyuu\mapsto y_{u}

is Fréchet differentiable and that its derivative zhz_{h} at uUu\in U, for a given hUh\in U, is the unique solution of the following system

{zh˙(t)=Azh(t)+u(t).Bzh(t)+h(t).(Byu(t)+b)zh(0)=0\begin{cases}\dot{z_{h}}(t)=Az_{h}(t)+u(t).Bz_{h}(t)+h(t).(By_{u}(t)+b)\\ z_{h}(0)=0\end{cases} (13)

Let uUu\in U and let yuy_{u} be the corresponding solution of the system (2). We claim that the linear mapping hzhh\mapsto z_{h} is continuous. Indeed, using the estimate (7) for the system (13), we can find some M>0M>0 such that

zhL(0,T;H)2MhU.\|z_{h}\|_{L^{\infty}(0,T;H)}\leq\sqrt{2}M\|h\|_{U}.

Let us denote by yu+hy_{u+h} the solution of the system (2) corresponding to u+hu+h, and let zhz_{h} be the solution of the system (13) corresponding to hh. Taking z=yh+uyuzhz=y_{h+u}-y_{u}-z_{h}, we can see that zz is the solution of the following system

{z˙(t)=Az(t)+u(t).Bz(t)+h(t).B(yh+u(t)yu(t))z(0)=0\begin{cases}\dot{z}(t)=Az(t)+u(t).Bz(t)+h(t).B(y_{h+u}(t)-y_{u}(t))\\ z(0)=0\end{cases} (14)

So, according to (7) in Lemma 1, the following estimates hold for some K>0K>0

zL(0,T;H)\displaystyle\|z\|_{L^{\infty}(0,T;H)} 2h.B(yh+u(t)yu(t))L2(0,T;V)\displaystyle\leq\sqrt{2}\|h.B(y_{h+u}(t)-y_{u}(t))\|_{L^{2}(0,T;V^{*})} (15)
KhUyh+uyuL(0,T;H).\displaystyle\leq K\|h\|_{U}\|y_{h+u}-y_{u}\|_{L^{\infty}(0,T;H)}.

Let us set w=yh+uyuw=y_{h+u}-y_{u}. Then ww is the solution of the following system

{w˙(t)=Aw(t)+u(t).Bw(t)+h(t).(Byh+u(t)+b)w(0)=0\begin{cases}\dot{w}(t)=Aw(t)+u(t).Bw(t)+h(t).(By_{h+u}(t)+b)\\ w(0)=0\end{cases} (16)

Applying Lemma 1, the following estimates hold for some K1,K2>0K_{1},K_{2}>0

wL(0,T;H)\displaystyle\|w\|_{L^{\infty}(0,T;H)} 2h.(Byh+u+b)L2(0,T;V)\displaystyle\leq\sqrt{2}\|h.(By_{h+u}+b)\|_{L^{2}(0,T;V^{*})} (17)
hU(K1yh+uL(0,T;H)+K2).\displaystyle\leq\|h\|_{U}\left(K_{1}\|y_{h+u}\|_{L^{\infty}(0,T;H)}+K_{2}\right).

Then using (15) and (17) and taking into account that the mapping uyuu\mapsto y_{u} is continuous, we conclude that for some K3>0K_{3}>0, we have

zL(0,T;H)K3hU2,\|z\|_{L^{\infty}(0,T;H)}\leq K_{3}\|h\|_{U}^{2},

and hence the mapping uyuu\mapsto y_{u} is Fréchet differentiable from UU to C(0,T;H)C(0,T;H), and that the derivative at uUu\in U is given by the system (13).
Since the mappings yyL2(0,T;H)2y\mapsto\|y\|_{L^{2}(0,T;H)}^{2} and uuU2u\mapsto\|u\|_{U}^{2} are Fréchet differentiable, we deduce that uJ(u)u\mapsto J(u) is Fréchet differentiable as well, and we have

{DuJ.h=J(u),hUDuJ.h=0T2y(t),zh(t)H𝑑t+r0Tu(t),h(t)N𝑑t.\begin{cases}D_{u}J.h=\langle J^{\prime}(u),h\rangle_{U}\\ D_{u}J.h=\int_{0}^{T}\langle 2y(t),z_{h}(t)\rangle_{H}\,dt+r\int_{0}^{T}\langle u(t),h(t)\rangle_{\mathbb{R}^{N}}dt.\end{cases} (18)

The well-posedness of the system (12) is guaranteed by Lemma 1, after the following change of variables

{q(t)=ϕ(Tt)g(t)=2y(Tt)v(t)=u(Tt)q(0)=ϕ(T)=0.\begin{cases}q(t)=\phi(T-t)\\ g(t)=2y(T-t)\\ v(t)=u(T-t)\\ q(0)=\phi(T)=0.\end{cases} (19)

Indeed, this leads to the following equivalent Cauchy problem:

{q˙(t)=Aq(t)v(t).Bq(t)+g(t)q(0)=0.\begin{cases}\dot{q}(t)=Aq(t)-v(t).Bq(t)+g(t)\\ q(0)=0.\end{cases}

Let yy and ϕ\phi be the mild solution of the systems (2) and (12) respectively. Then we have

0T2y(t)\displaystyle\int_{0}^{T}\langle 2y(t) ,zh(t)Hdt=0Tϕ˙(t)Aϕ(t)+u(t).Bϕ(t),zh(t)V,Vdt\displaystyle,z_{h}(t)\rangle_{H}dt=\int_{0}^{T}\langle-\dot{\phi}(t)-A\phi(t)+u(t).B\phi(t),z_{h}(t)\rangle_{V^{*},V}dt
=0Tϕ˙(t),zh(t)V,V+ϕ(t),Azh(t)+u(t).Bzh(t)V,Vdt\displaystyle=-\int_{0}^{T}\langle\dot{\phi}(t),z_{h}(t)\rangle_{V^{*},V}+\langle\phi(t),Az_{h}(t)+u(t).Bz_{h}(t)\rangle_{V^{*},V}dt
=0Tϕ˙(t),zh(t)V,V+ϕ(t),zh˙(t)h(t).(By(t)+b)V,Vdt\displaystyle=-\int_{0}^{T}\langle\dot{\phi}(t),z_{h}(t)\rangle_{V^{*},V}+\langle\phi(t),\dot{z_{h}}(t)-h(t).(By(t)+b)\rangle_{V^{*},V}dt
=0Tϕn˙(t),zh(t)V,V+ϕ(t),zh˙(t)V,Vdt\displaystyle=-\int_{0}^{T}\langle\dot{\phi_{n}}(t),z_{h}(t)\rangle_{V^{*},V}+\langle\phi(t),\dot{z_{h}}(t)\rangle_{V^{*},V}dt
+0Tϕ(t),h(t).(By(t)+b)V,Vdt\displaystyle\ +\int_{0}^{T}\langle\phi(t),h(t).(By(t)+b)\rangle_{V^{*},V}dt
=(ϕ(T),zh(T)V,Vϕ(0),zh(0)V,V)\displaystyle=-\left(\langle\phi(T),z_{h}(T)\rangle_{V^{*},V}-\langle\phi(0),z_{h}(0)\rangle_{V^{*},V}\right)
+0Tϕ(t),h(t).(By(t)+b)V,Vdt.\displaystyle\ +\int_{0}^{T}\langle\phi(t),h(t).(By(t)+b)\rangle_{V^{*},V}dt.

Since ϕ(T)=0\phi(T)=0 and zh(0)=0z_{h}(0)=0, we conclude that

0T2y(t),zh(t)H𝑑t\displaystyle\int_{0}^{T}\langle 2y(t),z_{h}(t)\rangle_{H}dt =0Tϕ(t),h(t).(By(t)+b)V,Vdt\displaystyle=\int_{0}^{T}\langle\phi(t),h(t).(By(t)+b)\rangle_{V^{*},V}dt (20)
=0T(By(t)+b))ϕ(t),h(t)Ndt.\displaystyle=\int_{0}^{T}\langle(By(t)+b))^{*}\phi(t),h(t)\rangle_{\mathbb{R}^{N}}dt.

Combining the formulas (18) and (20) we deduce that

J(u)(t),h(t)N=(By(t)+b)ϕ(t)+ru(t),h(t)N\langle J^{\prime}(u)(t),h(t)\rangle_{\mathbb{R}^{N}}=\langle(By(t)+b)^{*}\phi(t)+ru(t),h(t)\rangle_{\mathbb{R}^{N}}\cdot (21)

Hence the solution of the problem (4) satisfies

ui(t)=1rϕ(t),y(t)xi+biV,V,i=1,..Nu_{i}^{*}(t)=-\dfrac{1}{r}\langle\phi(t),\frac{\partial y(t)}{\partial x_{i}}+b_{i}\rangle_{V^{*},V},\ \ i=1,..N\cdot

This achieve this proof.

III-C Optimal control and strong stabilization

Let us consider the following quadratic cost function JJ:

J(u)=0+y(t)H2𝑑t+r20+u(t)N2𝑑t,J(u)=\int_{0}^{+\infty}\|y(t)\|_{H}^{2}dt+\frac{r}{2}\int_{0}^{+\infty}\|u(t)\|_{\mathbb{R}^{N}}^{2}dt,

where r>0,uU=L2(0,+;N)r>0,\ u\in U=L^{2}(0,+\infty;\mathbb{R}^{N}) and yy is the corresponding mild solution of the system (2).
The optimal control problem may be stated as follows

{minJ(u)uUad={uL2(0,+;N)/J(u)<+}\left\{\begin{array}[]{ll}\min J(u)\\ u\in U_{ad}=\{u\in L^{2}(0,+\infty;\mathbb{R}^{N})\ \ /\ J(u)<+\infty\}\end{array}\right. (22)

Our goal in this part is to give a solution of the problem (22). For this end, we consider the sequence of controls (un)(u_{n}) solutions of the problem (4) on [0,Tn][0,T_{n}] for an increasing sequence TnT_{n} such that Tn+T_{n}\rightarrow+\infty. Let us denote by yny_{n} the solution on [0,Tn][0,T_{n}] of the system (2), and by ϕn\phi_{n} the solution of the adjoint system (12).

We have the following result.

Theorem 4

Let us consider the control u=(ui)u^{*}=(u_{i}^{*}) defined by:

ui(t)=1rϕ(t),y(t)xi+bV,Vfori=1Nu_{i}^{*}(t)=-\dfrac{1}{r}\langle\phi(t),\frac{\partial y^{*}(t)}{\partial x_{i}}+b\rangle_{V^{*},V}\ \ \ for\ \ i=1...N (23)

where ϕ\phi is a weak limit value of (ϕn)(\phi_{n}) in L2(0,+;V)L^{2}(0,+\infty;V) and yy^{*} is the corresponding solution of the system 2. Then

  • uu^{*} is a solution of the problem (22)

  • uu^{*} guarantees the strong stabilization of the system (2).

Proof 3

Let us first observe that UadU_{ad}\neq\emptyset, as here the solution of the system (2) corresponding to u=0u=0 is exponentially stable.
Let JnJ_{n} be the functional (3) in [0,Tn],[0,T_{n}], and let us define the following sequence of globally defined controls:

vn(t)={un(t),iftTn0,ift>Tn.v_{n}(t)=\left\{\begin{array}[]{ll}u_{n}(t),\ \ \ \hbox{if}\ \ \ t\leq T_{n}\\ 0,\ \ \ \hbox{if}\ \ \ t>T_{n}.\end{array}\right.

Since unL2(0,Tn;N),u_{n}\in L^{2}(0,T_{n};\mathbb{R}^{N}), it follows that vnL2(0,+;N)v_{n}\in L^{2}(0,+\infty;\mathbb{R}^{N}). Let us consider the mapping :

R:vr20+v(t)N2𝑑t.R:\ v\mapsto\frac{r}{2}\displaystyle\int_{0}^{+\infty}\|v(t)\|_{\mathbb{R}^{N}}^{2}dt.

Let vUadv\in U_{ad} be fixed. Since unu_{n} is a solution of the problem (4) in [0,Tn][0,T_{n}], it comes that

R(vn)=r20Tnun(t)N2𝑑tJn(un)Jn(v)J(v).R(v_{n})=\frac{r}{2}\int_{0}^{T_{n}}\|u_{n}(t)\|^{2}_{\mathbb{R}^{N}}dt\leq J_{n}(u_{n})\leq J_{n}(v)\leq J(v).

Thus R(vn)R(v_{n}) is bounded and so is vnv_{n}. We deduce that the sequence (vn)(v_{n}) admits a subsequence, still denoted by (vn)(v_{n}), which weakly converges to uLp(0,+)u^{*}\in L^{p}(0,+\infty).

Similarly to the proof of Theorem 2, we deduce that there exists a subsequence of (vn)(v_{n}), (which can be also denoted by (vn)(v_{n})) such that

yvnystronglyinL2(0,+;H),y_{v_{n}}\rightarrow y^{*}\ strongly\ in\ L^{2}(0,+\infty;H),

where yy^{*} is the mild solution of the system (2) corresponding to uu^{*} in infinite time-horizon (i.e. T=+T=+\infty). Then we conclude that

limn+0Tnyvn(t)H2𝑑t=0+y(t)H2𝑑t.\lim_{n\rightarrow+\infty}\int_{0}^{T_{n}}\|y_{v_{n}}(t)\|_{H}^{2}dt=\int_{0}^{+\infty}\|y^{*}(t)\|_{H}^{2}dt. (24)

The continuity of the mapping RR implies the lower semi-continuity w.r.t to the weak topology (see Corollary III.8 in [13]). We deduce that

R(u)lim infn+R(vn)R(u^{*})\leq\liminf_{n\rightarrow+\infty}R(v_{n})\cdot (25)

Observing that

Jn(un)=0+yvn1(0,Tn)(t)H2𝑑t+r20+vn(t)N𝑑t,J_{n}(u_{n})=\int_{0}^{+\infty}\|y_{v_{n}}1_{(0,T_{n})}(t)\|_{H}^{2}dt+\frac{r}{2}\int_{0}^{+\infty}\|v_{n}(t)\|_{\mathbb{R}^{N}}dt,

we derive via (24) and (25)

J(u)lim infn+(Jn(un))J(u^{*})\leq\liminf_{n\rightarrow+\infty}(J_{n}(u_{n}))\cdot (26)

Let us show that the sequence (Jn(un))n(J_{n}(u_{n}))_{n\in\mathbb{N}} converges to J(u)J(u^{*}). For this end, we will show that the sequence (Jn(un))n(J_{n}(u_{n}))_{n\in\mathbb{N}} is increasing and upper bounded by J(u)J(u^{*}). We have

Jn(un)Jn(un+1)Jn+1(un+1)andJn(un)Jn(u)J(u),J_{n}(u_{n})\leq J_{n}(u_{n+1})\leq J_{{n+1}}(u_{n+1})\ \ and\ \ J_{n}(u_{n})\leq J_{n}(u^{*})\leq J(u^{*}),

from which it comes

limn+Jn(un)J(u)\lim_{n\rightarrow+\infty}J_{n}(u_{n})\leq J(u^{*})\cdot (27)

Combining (26) and (27), we conclude that

limn+Jn(un)=J(u)\lim_{n\rightarrow+\infty}J_{n}(u_{n})=J(u^{*})\cdot

Keeping in mind that unu_{n} is the solution of the problem (4) on [0,Tn][0,T_{n}], we conclude that:

Jn(un)J(v)\displaystyle J_{n}(u_{n})-J(v) =0Tn(un(t)N2+yn(t)H2)𝑑t\displaystyle=\int_{0}^{T_{n}}\left(\|u_{n}(t)\|_{\mathbb{R}^{N}}^{2}+\|y_{n}(t)\|_{H}^{2}\right)\,dt
0+(v(t)N2+yv(t)H2)𝑑t\displaystyle\ \ -\int_{0}^{+\infty}\left(\|v(t)\|_{\mathbb{R}^{N}}^{2}+\|y_{v}(t)\|_{H}^{2}\right)\,dt
=Jn(un)Jn(v)Tn+(v(t)N2+yv(t)H2)𝑑t\displaystyle=J_{n}(u_{n})-J_{n}(v)-\int_{T_{n}}^{+\infty}(\|v(t)\|_{\mathbb{R}^{N}}^{2}+\|y_{v}(t)\|_{H}^{2})dt
0\displaystyle\leq 0\cdot

Thus letting n+n\longrightarrow+\infty, we get

J(u)J(v)=limn+Jn(un)J(v)0.J(u^{*})-J(v)=\lim_{n\rightarrow+\infty}J_{n}(u_{n})-J(v)\leq 0.

This shows that uu^{*} is a solution of the problem (22). Let ϕn\phi_{n} be the solution of the adjoint system (12) corresponding to unu_{n}. By the change of variables given by (19), qnq_{n} is solution of the following system

{q˙n(t)=Aqn(t)vn(t).Bqn(t)+gn(t)qn(0)=0\begin{cases}\dot{q}_{n}(t)=Aq_{n}(t)-v_{n}(t).Bq_{n}(t)+g_{n}(t)\\ q_{n}(0)=0\par\end{cases}

So, by the estimate (6) in Lemma 1, we have ϕn(Tn.)L2(0,T;V)=qnL2(0,T;V)gnL2(0,T;H)\begin{aligned} \|\phi_{n}(T_{n}-.)\|_{L^{2}(0,T;V)}=\|q_{n}\|_{L^{2}(0,T;V)}\leq\|g_{n}\|_{L^{2}(0,T;H)}\end{aligned} and gnL2(0,T;H)=yn(Tn.)L2(0,T;V)\begin{aligned} \|g_{n}\|_{L^{2}(0,T;H)}=\|y_{n}(T_{n}-.)\|_{L^{2}(0,T;V)}\cdot\end{aligned} Then the boundedness of yny_{n} implies that of ϕn\phi_{n} in L2(0,+;V)L^{2}(0,+\infty;V). So, we can deduce that the sequence (ϕn)(\phi_{n}) admits a subsequence, still denoted by (ϕn)(\phi_{n}), which weakly converges to ϕL2(0,+;V)\phi\in L^{2}(0,+\infty;V).
Using the fact that unuu_{n}\rightharpoonup u^{*} in UU, ynyy_{n}\rightarrow y^{*} in L2(0,+;V)L^{2}(0,+\infty;V) and ϕnϕ\phi_{n}\rightharpoonup\phi in L2(0,+;V)L^{2}(0,+\infty;V), we conclude by Theorem 3 that

ui(t)=1rϕ(t),y(t)xi+bV,V,i=1,..,N.u_{i}^{*}(t)=-\dfrac{1}{r}\langle\phi(t),\frac{\partial y^{*}(t)}{\partial x_{i}}+b\rangle_{V^{*},V},\forall\,i=1,..,N.

Let us now show that this controls lead to a strongly stable system in closed loop. For all 0<s<t<+,0<s<t<+\infty, we have

|y(t)2y(s)2|\displaystyle|\|y^{*}(t)\|^{2}-\|y^{*}(s)\|^{2}| =|st2y˙(r),y(r)V,V𝑑r|\displaystyle=|\int_{s}^{t}2\langle\dot{y}^{*}(r),y^{*}(r)\rangle_{V^{*},V}dr|
2y˙L2(s,t;V)(sty(r)V2𝑑r)12.\displaystyle\leq 2\|\dot{y}^{*}\|_{L^{2}(s,t;V^{*})}\left(\int_{s}^{t}\|y^{*}(r)\|_{V}^{2}dr\right)^{\frac{1}{2}}.

Then, according to estimate (8), we have for some My0>0M_{y_{0}}>0

|y(t)2y(s)2|My0(sty(r)V2𝑑r)12.\left|\|y^{*}(t)\|^{2}-\|y^{*}(s)\|^{2}\right|\leq M_{y_{0}}\left(\int_{s}^{t}\|y^{*}(r)\|_{V}^{2}dr\right)^{\frac{1}{2}}.

Using the fact that 0+u(t)2𝑑t<+\int_{0}^{+\infty}\|u^{*}(t)\|^{2}dt<+\infty, we deduce via (6) that 0+y(t)V2<+\int_{0}^{+\infty}\|y^{*}(t)\|^{2}_{V}<+\infty. Then we conclude that

y(t)0ast+.\|y^{*}(t)\|\rightarrow 0\;\hbox{as}\;t\rightarrow+\infty.

IV A numerical example

Here, we will present simulations in which we show numerically the strong stability of the optimal trajectory yy^{*} and we further compare numerically the optimal control w.r.t some controls vv in terms of energy consumption.
Let us consider the following parameters: Ω=(0,1)\Omega=(0,1), b=5b=5, y0(x)=10x(1x)y_{0}(x)=10x(1-x), r=1r=1 and T=8T=8.
Then reporting the states norm of the system for both controls uu^{*} and v1=0v_{1}=0 in Figure 1, we can see that uu^{*} performs slightly better than the zero control. This tendency is confirmed in the table below regarding the states norm and the energy consumed by the system under the optimal control and constant controls.

Refer to caption
Figure 1: Time-evolution of y(t)\|y^{*}(t)\| under the optimal control uu^{*} (black line) and the zero control v1=0v_{1}=0 (blue line).
Time (t) 0.2 0.3 0.6
y(t)\|y^{*}(t)\| 15.810215.8*10^{-2} 4.791024.79*10^{-2} 1.21031.2*10^{-3}
yv1(t)\|y_{v_{1}}(t)\| 24.810224.8*10^{-2} 9.121029.12*10^{-2} 4.561034.56*10^{-3}
J(u)J(u^{*}) 14.6810214.68*10^{-2} 14.7710214.77*10^{-2} 14.7710214.77*10^{-2}
J(v1)J(v_{1}) 16.3810216.38*10^{-2} 16.6310216.63*10^{-2} 16.6710216.67*10^{-2}
J(v2)J(v_{2}) 4.194.19 4.214.21 4.224.22
Remark 5

Note that the stabilization problem of non-homogeneous distributed bilinear systems has been only considered for bounded control operator (see [2, 9, 10, 17]). Thus the existing results from the above literature are not applicable as here, the control operator BB is unbounded. Moreover, even in the homogeneous case (i.e b=0b=0), the existing results for unbounded operator BB (see [4, 8]) are not applicable as here, the operator BB is skew adjoint. In particular, the observation inequality is not verified. Now, if we formally consider the feedback control v(t)v(t) used in [4, 8], then we find v(t)=0v(t)=0 as v(t)v(t) involves the term By,y\langle By,y\rangle which is null when BB is skew-adjoint.

V Conclusion

In this work, we studied the quadratic optimal control problem for a class of non-homogeneous bilinear Fokker-Planck equation. Both finite and infinite horizon cases are considered. It is further showed that the infinite horizon optimal control leads to a stabilized state of the system in closed-loop. This study provided a stabilization result which does not require the observation assumption. The result of Theorem 5 is promising. Indeed one can be inspired by it to investigate the optimal stabilization of a general unbounded bilinear system.

References

  • [1] A. Addou and A. Benbrik.(2002) Existence and uniqueness of optimal control for a distributed parameter bilinear system, J. Dyn. Control Syst., 8, 141–152.
  • [2] M, Akkouchi, & A. Bounabat (2003). Weak stabilizability of a nonautonomous and non-linear system. Mathematica Pannonica, 55, 61.
  • [3] M. S. Aronna, & F. Tröltzsch (2021). First and second-order optimality conditions for the control of Fokker-Planck equations. ESAIM: Control, Optimisation and Calculus of Variations, 27, 15.
  • [4] R. E. Ayadi, M. Ouzahra, & A. Boutoulout. (2012) Strong stabilisation and decay estimate for unbounded bilinear systems. International journal of control, 85(10), 1497-1505.
  • [5] J. M. Ball, J. E. Marsden and M. Slemrod. (1982) Controllability for distributed bilinear systems, SIAM J. Control Optim., 20, 575–597.
  • [6] S. Banks and M. Yew. (1985) On a class of suboptimal controls for infinite-dimensional bilinear systems, Systems and Control Letters, 5, 327–333.
  • [7] L. Berrahmoune. (2009) A note on admissibility for unbounded bilinear control systems, Bulletin of the Belgian Mathematical Society-Simon Stevin, 16 , 193–204.
  • [8] L. Berrahmoune. (2010). Stabilization of unbounded bilinear control systems in Hilbert space. Journal of mathematical analysis and applications, 372(2), 645-655.
  • [9] H. Bounit, & H. Hammouri. (1999). Feedback stabilization for a class of distributed semilinear control systems. Nonlinear Analysis: Theory, Methods & Applications, 37(8), 953-969.
  • [10] H. Bounit. (2003) Comments on the feedback stabilization for bilinear control systems. Applied mathematics letters, 16(6), 847-851.
  • [11] M. E. Bradly and S. Lenhart. (1994) Bilinear optimal control of a Kirchhoff plate, Syst. Control Lett., 22, 27–38.
  • [12] A. Bensoussan, G. Da Prato, M.C. Delfour and S.K. Mitter. (2007)Representation and Control of Infinite Dimensional Systems, Birkhaüser, Boston.
  • [13] H. Brezis (1987) Analyse fonctionnelle : Théorie et applications, Paris, Masson.
  • [14] J. M. Clérin. (2009) Problèmes de contrôle optimal du type bilinéaire gouvernés par des équations aux dérivées partielles d’évolution, Doctoral dissertation.
  • [15] N. El Alami.(1986) Analyse et commande optimale des systèmes bilinéaires distribués, application aux procédés energétiques, Doctorat d’Etat-I.M.P. Perpignan, 1986.
  • [16] A. Fleig and R. Guglielmi. (2016) Bilinear optimal control of the Fokker-Planck equation, IFAC-PapersOnLine, 49, 254–259.
  • [17] Z. Hamidi, & M. Ouzahra. (2018). Partial stabilisation of non-homogeneous bilinear systems. International Journal of Control, 91(6), 1251-1258.
  • [18] R. E. Kalman. (1960) Contributions to the theory of optimal control, Bol. Soc. Mat. Mexicano, 2, 102–119.
  • [19] N. A. Krall, & A. W. Trivelpiece. (1973). Principles of plasma physics. American Journal of Physics, 41(12), 1380-1381.
  • [20] X. Li and J. Yong. (2012) Optimal control theory for infinite dimensional systems, Springer Science &\& Business Media.
  • [21] M. Ouzahra. (2007) Stabilization with Decay Estimate for a Class of Distributed Bilinear Systems, European Journal of Control, 5 509–515.
  • [22] A. Pazy. (1983) Semi-groups of Linear Operators and Applications to Partial Differential Equations, Springer-Verlag, New York.
  • [23] L. S. Pontryagin, V. G. Boltanski, R. V. Gamkrelidze and E. F. Mischenko (1962) Mathematical theory of optimal processes, Wiley, New York.
  • [24] J.P. Quinn. (1982) Stabilization of bilinear systems by quadratic feedback control, J. Math. Anal. Appl, 75 66–80.
  • [25] E. P. Ryan. (1984) Optimal feedback control of bilinear systems, Journal of Optimization Theory and Applications, 44 333–362.
  • [26] C. S. Chen. (2009) Quadratic optimal neural fuzzy control for synchronization of uncertain chaotic systems, Expert Systems with Applications, 36 11827–11835.
  • [27] S. Yahyaoui and M. Ouzahra. (2021) Quadratic optimal control and feedback stabilization of bilinear systems, Optimal Control Applications and Methods, https://doi.org/10.1002/oca.2704.
  • [28] E. Zerrik and N. El Boukhari. (2019) Regional optimal control for a class of semilinear systems with distributed controls, International Journal of Control, 92 896–907
  • [29] E. Zerrik and N. El Boukhari. (2018) Constrained optimal control for a class of semilinear infinite dimensional Systems, Journal of Dynamical and Control Systems, 24 65–81.