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

Necessary optimality conditions of a reaction-diffusion SIR model with ABC fractional derivatives

Abstract.

The main aim of the present work is to study and analyze a reaction-diffusion fractional version of the SIR epidemic mathematical model by means of the non-local and non-singular ABC fractional derivative operator with complete memory effects. Existence and uniqueness of solution for the proposed fractional model is proved. Existence of an optimal control is also established. Then, necessary optimality conditions are derived. As a consequence, a characterization of the optimal control is given. Lastly, numerical results are given with the aim to show the effectiveness of the proposed control strategy, which provides significant results using the AB fractional derivative operator in the Caputo sense, comparing it with the classical integer one. The results show the importance of choosing very well the fractional characterization of the order of the operators.

Key words and phrases:
Epidemic model; Optimality conditions; Reaction-diffusion equations; Atangana–Baleanu–Caputo fractional derivatives; Numerical simulations
1991 Mathematics Subject Classification:
34A08, 49K20; 35K57, 47H10
Corresponding author: M. R. Sidi Ammi

Moulay Rchid Sidi Ammia∗, Mostafa Tahiria and Delfim F. M. Torresb

aDepartment of Mathematics, AMNEA Group, Laboratory MAIS,

Faculty of Sciences and Techniques, Moulay Ismail University of Meknes, Morocco

bCenter for Research and Development in Mathematics and Applications (CIDMA),

Department of Mathematics, University of Aveiro, 3810-193 Aveiro, Portugal.


1. Introduction

Fractional derivatives give rise to theoretical models that allow a significant improvement in the fitting of real data when compared with analogous classical models [3]. For real data of Florida Department of Health from September 2011 to July 2014, some authors conclude that the absolute error between the solutions obtained statistically and that of fractional models are smaller than those obtained by models of integer derivatives [24]. In the fractional calculus literature, systems using fractional derivatives give a more realistic behavior [26, 25, 23]. There exists many definitions of fractional derivative [23]. Among the more well-known fractional derivatives, we can cite the Riemann–Liouville one. It is not always suitable for modeling physical systems, because the Riemann–Liouville derivative of a constant is not zero, and the initial conditions of associated Cauchy problems are expressed by fractional derivatives. Caputo fractional derivatives offers another alternative, where the derivative of a constant is null and initial conditions are expressed as in the classical case of integer order derivatives [25, 23, 13]. However, the kernel of this derivative has a singularity. Fractional derivatives that possess a non-singular kernel have aroused more interest from the scientific community. This is due to the non-singular memory of the Mittag–Leffler function and also to the non-obedience of the algebraic criteria of associativity and commutativity. The ABC fractional derivative is sometimes preferable for modeling physical dynamical systems, giving a good description of the phenomena of heterogeneity and diffusion at different scales [1, 5, 6].

Fractional calculus plays an important role in many areas of science and engineering. It also finds application in optimal control problems. The principle of mathematical theory of control is to determine a state and a control for a dynamic system during a specified period to optimize a given objective [27]. Fractional optimal control problems have been formulated and studied as fractional problems of the calculus of variations. Some authors have shown that fractional differential equations are more accurate than integer-order differential equations, and that fractional controllers work better than integer-order controllers [7, 20, 21, 28]. In [30], Yuan et al. have studied problems of fractional optimal control via left and right fractional derivatives of Caputo. A numerical technique for the solution of a class of fractional optimal control problems, in terms of both Riemann–Liouville and Caputo fractional derivatives, is presented in [8]. Authors in [9, 11] present a pseudo-state-space fractional optimal control problem formulation. Fixed and free final-time fractional optimal control problems are considered in [10, 12]. Guo [16] formulates a second-order necessary optimality condition for fractional optimal control problems in the sense of Caputo. Optimal control of a fractional-order HIV-immune system, in terms of Caputo fractional derivatives, is discussed in [14]. In [22], authors proposed a fractional-order optimal control model for malaria infection in terms of the Caputo fractional derivative. Optimal control of fractional diffusion equations has also been studied by several authors. For instance, in [2], Agrawal considers two problems, the simplest fractional variation problem and a fractional variational problem in Lagrange form. For both problems, the author developed Euler–Lagrange type necessary conditions, which must be satisfied for the given functional to have an extremum. In [26], authors prove necessary optimality conditions of a nonlocal thermistor problem with ABC fractional time derivatives.

Several infectious diseases confer permanent immunity against reinfection. This type of diseases can be modeled by the SIRSIR model. The total population (N)(N) is divided into three compartments with N=S+I+RN=S+I+R, where SS is the number of susceptible (those able to contract the disease), II is the number of infectious individuals (those capable of transmitting the disease), and RR is the number of individuals recovered (those who have recovered and become immune). Vaccines are extremely important and have been proved to be most effective and cost-efficient method of preventing infectious diseases, such as measles, polio, diphtheria, tetanus, pertussis, tuberculosis, etc. The study of fractional calculus with a non-singular kernel is gaining more and more attention. Compared with classical fractional calculus with a singular kernel, non-singular kernel models can describe reality more accurately, which has been shown recently in a variety of fields such as physics, chemistry, biology, economics, control, porous media, aerodynamics and so on. For example, extensive treatment and various applications of fractional calculus with non-singular kernel has been discussed in the works of Atangana and Baleanu [5], and Djida et al. [15]. It has been demonstrated that fractional order differential equations (FODEs) with non-singular kernels give rise to dynamic system models that are more accurately.

In this work, we consider an optimal control problem for the reaction-diffusion SIR system with Atangana–Baleanu fractional derivative in the Caputo sense (the ABC operator). Our aim is to study the effect of non-local memory and vaccination strategies on the cost, needed to control the spread of infectious diseases. Our results generalize to the ABC fractional setting previous studies of classical control theory presented in [17]. The considered model there does not explain the influence of a complete memory of the system. For that, we extend such nonlinear system of first order differential equations to a fractional-order one in the ABC sense. We have further improved the cost and effectiveness of proposed control strategy during a given period of time.

This paper is organized as follows. Some important definitions related to the ABC fractional derivative operator and its properties are presented in Section 2, while the underlying fractional reaction-diffusion SIR mathematical model is formulated in Section 3. This led to the necessity of proving existence and uniqueness of solution to the proposed fractional model as well as existence of an optimal control. These results are extensively discussed in Sections 4 and 5. Section 6 is devoted to necessary optimality conditions. Interesting numerical tests, showing the importance of choosing very well the fractional characterization α\alpha, are given in Section 7. Finally, conclusions of the present study are widely discussed in Section 8.

2. Preliminary results

We now recall some properties on the Mittag–Leffler function and the definition of ABC fractional time derivative. First, we define the two-parameter Mittag–Leffler function Eα,ξ(z)E_{\alpha,\xi}(z), as the family of entire functions of zz given by

Eα,ξ(z)=k=0zkΓ(kα+ξ),z,E_{\alpha,\xi}(z)=\sum_{k=0}^{\infty}\frac{z^{k}}{\Gamma(k\alpha+\xi)},\quad z\in\mathbb{C},

where Γ()\Gamma(\cdot) denotes the Gamma function

Γ(z)=0tz1et𝑑t,Re(z)>0.\Gamma(z)=\int_{0}^{\infty}t^{z-1}e^{-t}dt,\quad Re(z)>0.

Observe that the Mittag–Leffler function is a generalization of the exponential function: E1,1(z)=ezE_{1,1}(z)=e^{z}. For more information about the definition of fractional derivative in the sense of Atangana–Baleanu, the reader can see [5, 6].

Definition 2.1.

For a given function gH1(a,T)g\in H^{1}(a,T), T>aT>a, the Atangana–Baleanu fractional derivative in Caputo sense, shortly called the ABC fractional derivative of gg of order α(0,1)\alpha\in(0,1) with base point aa, is defined at a point t(a,T)t\in(a,T) by

abcaDtαg(t)=B(α)1αatg(τ)Eα[γ(tτ)α]dτ,^{abc}_{a}D_{t}^{\alpha}g(t)=\frac{B(\alpha)}{1-\alpha}\int_{a}^{t}g^{\prime}(\tau)E_{\alpha}[-\gamma(t-\tau)^{\alpha}]d\tau, (1)

where γ=α1α\gamma=\frac{\alpha}{1-\alpha}, Eα,1=EαE_{\alpha,1}=E_{\alpha} stands for the Mittag–Leffler function, and B(α)=(1α)+αΓ(α)B(\alpha)=(1-\alpha)+\frac{\alpha}{\Gamma(\alpha)}. Furthermore, the Atangana–Baleanu fractional integral of order α(0,1)\alpha\in(0,1) with base point aa is defined as

abaItαg(t)=1αB(α)g(t)+αB(α)Γ(α)atg(τ)(tτ)α1dτ.^{ab}_{a}I_{t}^{\alpha}g(t)=\frac{1-\alpha}{B(\alpha)}g(t)+\frac{\alpha}{B(\alpha)\Gamma(\alpha)}\int_{a}^{t}g(\tau)(t-\tau)^{\alpha-1}d\tau. (2)
Remark 1.

The usual ordinary derivative t\partial_{t} is obtained by letting α1\alpha\rightarrow 1 in (1). If α=0,1\alpha=0,1 in (2), then we get the initial function and the classical integral, respectively.

Definition 2.2 (See [1]).

For a given function gH1(a,T)g\in H^{1}(a,T), T>aT>a, the backward Atangana–Baleanu fractional derivative in Caputo sense of gg of order α(0,1)\alpha\in(0,1) with base point TT, is defined at a point t(a,T)t\in(a,T) by

abcTDtαg(t)=B(α)1αtTg(τ)Eα[γ(tτ)α]dτ.^{abc}_{T}D_{t}^{\alpha}g(t)=-\frac{B(\alpha)}{1-\alpha}\int_{t}^{T}g^{\prime}(\tau)E_{\alpha}[-\gamma(t-\tau)^{\alpha}]d\tau. (3)

3. Model formulation

The SIR model is one of the simplest compartmental models. It was first used by Kermack and McKendrick in 1927. It has subsequently been applied to a variety of diseases, especially airborne childhood diseases with lifelong immunity upon recovery, such as measles, mumps, rubella, and pertussis. We assume that the populations are in a spatially homogeneous environment and their densities depend on space, reflecting the spacial spread of the disease. Then, the model will be formulated as a system of reaction-diffusion equations. In this section, we formulate an optimal control of a nonlocal fractional SIR epidemic model with parabolic equations and boundary conditions. We consider that the movement of the population depends on the time tt and the space xx. Furthermore, all susceptible vaccinates are transferred directly to the recovered class.

Let QT=[0,T]×ΩQ_{T}=[0,T]\times\Omega and ΣT=[0,T]×Ω\Sigma_{T}=[0,T]\times\partial\Omega, where Ω\Omega is a fixed and bounded domain in 2\mathbb{R}^{2} with smooth boundary Ω\partial\Omega, and [0,T][0,T] is a finite interval. The dynamic of the ABC fractional SIR system with control is given by

D0abcStα(t,x)\displaystyle{}^{abc}_{0}D{{}_{t}^{\alpha}}S(t,x) =λ1ΔS(t,x)+μN(t,x)βS(t,x)I(t,x)dS(t,x)\displaystyle=\lambda_{1}\Delta S(t,x)+\mu N(t,x)-\beta S(t,x)I(t,x)-dS(t,x) (4)
u(t,x)S(t,x),\displaystyle\quad-u(t,x)S(t,x),
D0abcItα(t,x)\displaystyle{}^{abc}_{0}D{{}_{t}^{\alpha}}I(t,x) =λ2ΔI(t,x)+βS(t,x)I(t,x)(d+r)I(t,x),(t,x)QT,\displaystyle=\lambda_{2}\Delta I(t,x)+\beta S(t,x)I(t,x)-(d+r)I(t,x),\hskip 5.69046pt(t,x)\in Q_{T},
D0abcRtα(t,x)\displaystyle{}^{abc}_{0}D{{}_{t}^{\alpha}}R(t,x) =λ3ΔR(t,x)+rI(t,x)dR(t,x)+u(t,x)S(t,x),\displaystyle=\lambda_{3}\Delta R(t,x)+rI(t,x)-dR(t,x)+u(t,x)S(t,x),

with the homogeneous Neumann boundary conditions

S(t,x)ν=I(t,x)ν=R(t,x)ν=0,(t,x)ΣT,\frac{\partial S(t,x)}{\partial\nu}=\frac{\partial I(t,x)}{\partial\nu}=\frac{\partial R(t,x)}{\partial\nu}=0,\quad(t,x)\in\Sigma_{T}, (5)

and the following initial conditions of the three populations, which are considered positive for biological reasons:

S(0,x)=S0,I(0,x)=I0 and R(0,x)=R0,xΩ.S(0,x)=S_{0},\quad I(0,x)=I_{0}\quad\text{ and }\quad R(0,x)=R_{0},\quad x\in\Omega. (6)

The positive constants μ\mu, rr and dd are respectively the birth rate, the recovery rate of the infective individuals and the natural death rate. Susceptible individuals acquire infection by the contact with individuals in the class II at a rate βSI\beta SI, where β\beta is the infection coefficient. Positive constants λ1\lambda_{1}, λ2\lambda_{2}, λ3\lambda_{3} denote the diffusion coefficients for the susceptible, infected and recovered individuals. The control uu describes the effect of vaccination. It is assumed that vaccination transforms susceptible individuals to recovered ones and confers them immunity. The notation Δ=2x2+2y2\Delta=\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}} represents the usual Laplacian operator in two-dimensional space; ν\nu is the outward unit normal vector on the boundary with ν=ν\frac{\partial}{\partial\nu}=\nu; and \nabla is the normal derivative on Ω\partial\Omega. The no-flux homogeneous Neumann boundary conditions imply that model (4) is self-contained and there is a dynamic across the boundary, but there is no emigration.

Since the vaccination is limited and represents an economic burden, one important issue and goal is to know how much we should spend in vaccination to reduce the number of infections and, at the same time, save the cost of vaccination program. This can be mathematically interpreted by optimizing the following objective functional:

J(S,I,R,u)=I(t,x)L2(QT)2+I(T,)L2(Ω)2+θu(t,x)L2(QT)2,J(S,I,R,u)=\|I(t,x)\|_{L^{2}(Q_{T})}^{2}+\|I(T,\cdot)\|_{L^{2}(\Omega)}^{2}+\theta\|u(t,x)\|_{L^{2}(Q_{T})}^{2}, (7)

where θ\theta is a weight constant for the vaccination control uu, which belongs to the set

Uad={uL(QT);uL(QT)<1 and u>0}U_{ad}=\{u\in L^{\infty}(Q_{T});\|u\|_{L^{\infty}(Q_{T})}<1\quad\mbox{ and }\quad u>0\} (8)

of admissible controls. Let y=(y1,y2,y3)=(S,I,R)y=(y_{1},y_{2},y_{3})=(S,I,R), y0=(y10,y20,y30)=(S0,I0,R0)y^{0}=(y_{1}^{0},y_{2}^{0},y_{3}^{0})=(S_{0},I_{0},R_{0}), λ=(λ1,λ2,λ3)\lambda=(\lambda_{1},\lambda_{2},\lambda_{3}), L2(Ω)=(L2(Ω))3\mathrm{L}^{2}(\Omega)=(L^{2}(\Omega))^{3} and AA be the linear diffusion operator defined by

A:D(A)L2(Ω)L2(Ω)A:D(A)\subset\mathrm{L}^{2}(\Omega)\rightarrow\mathrm{L}^{2}(\Omega)
Ay=λΔy=(λ1Δy1,λ2Δy2,λ3Δy3),yD(A),Ay=\lambda\Delta y=(\lambda_{1}\Delta y_{1},\lambda_{2}\Delta y_{2},\lambda_{3}\Delta y_{3}),\quad\forall y\in D(A),

where

D(A)={y=(y1,y2,y3)(H2(Ω))3,y1ν=y2ν=y3ν=0,a.e. xΩ}.D(A)=\left\{y=(y_{1},y_{2},y_{3})\in(H^{2}(\Omega))^{3},\frac{\partial y_{1}}{\partial\nu}=\frac{\partial y_{2}}{\partial\nu}=\frac{\partial y_{3}}{\partial\nu}=0,\quad\mbox{a.e. }x\in\partial\Omega\right\}.

We also set

f(y(t))=(f1(y(t)),f2(y(t)),f3(y(t)))f(y(t))=(f_{1}(y(t)),f_{2}(y(t)),f_{3}(y(t)))

with

{f1(y(t))=μ(y1+y2+y3)βy1y2dy1uy1,f2(y(t))=βy1y2(d+r)y2,t[0,T],f3(y(t))=ry2dy3+uy1.\begin{cases}f_{1}(y(t))=&\mu(y_{1}+y_{2}+y_{3})-\beta y_{1}y_{2}-dy_{1}-uy_{1},\\ f_{2}(y(t))=&\beta y_{1}y_{2}-(d+r)y_{2},\quad t\in[0,T],\\ f_{3}(y(t))=&ry_{2}-dy_{3}+uy_{1}.\end{cases}

The problem can be rewritten in a compact form as

{Dtα0abcy=Ay+f(y(t)),y(0)=y0,\begin{cases}{}^{abc}_{0}D_{t}^{\alpha}y=Ay+f(y(t)),\\ y(0)=y^{0},\end{cases}

where Dtα0abc{}^{abc}_{0}D_{t}^{\alpha} is the Atangana–Baleanu fractional derivative of order α(0,1)\alpha\in(0,1) in the sense of Caputo with respect to time tt. The symbol \triangle denotes the Laplacian with respect to the spacial variables, defined on H2(Ω)H01(Ω)H^{2}(\Omega)\bigcap H_{0}^{1}(\Omega).

4. Existence of solution

Existence of solution is proved in the weak sense.

Definition 4.1.

We say that yy is a weak solution to (4) if

Ω(0abcDtαy)vdx+λΩyvdx=Ωf(y)vdx\int_{\Omega}\,(^{abc}_{0}D_{t}^{\alpha}y)vdx+\lambda\int_{\Omega}\nabla y\nabla vdx=\int_{\Omega}f(y)vdx (9)

for all vH1(Ω)v\in H^{1}(\Omega).

Integrating by parts, involving the ABC fractional-time derivative (see [15]), and using a straightforward calculation, one obtains the following result.

Proposition 1.

Let y,vC(QT¯)y,v\in{C^{\infty}}(\overline{Q_{T}}). Then,

0TDtα0abcyv𝑑t=0TDtαTabcvy𝑑t+B(α)1αv(x,T)0TyEα,α[γ(Tt)α]𝑑tB(α)1αy(x,0)0TEα,α[γtα]v𝑑t.\int_{0}^{T}\,{}^{abc}_{0}D_{t}^{\alpha}y\cdot vdt=-\int_{0}^{T}\,{}^{abc}_{T}D_{t}^{\alpha}v\cdot ydt+\frac{B(\alpha)}{1-\alpha}v(x,T)\int_{0}^{T}yE_{\alpha,\alpha}[-\gamma(T-t)^{\alpha}]dt\\ -\frac{B(\alpha)}{1-\alpha}y(x,0)\int_{0}^{T}E_{\alpha,\alpha}[-\gamma t^{\alpha}]vdt. (10)

Using the boundary conditions (5), we immediately get the following Corollary.

Corollary 1.

Let y,vC(QT¯)y,v\in{C^{\infty}}(\overline{Q_{T}}). Then,

Ω0T(Dtα0abcyλy)v𝑑x𝑑t=B(α)1αΩ0Ty(x,0)Eα,α[γtα]+0TΩy(TabcDtαvλv)𝑑x𝑑t+B(α)1αΩv(x,T)0TyEα,α[γ(Tt)α]𝑑t𝑑x.\begin{split}\int_{\Omega}&\int_{0}^{T}\left({}^{abc}_{0}D_{t}^{\alpha}y-\lambda\triangle y\right)vdxdt\\ &=-\frac{B(\alpha)}{1-\alpha}\int_{\Omega}\int_{0}^{T}y(x,0)E_{\alpha,\alpha}[-\gamma t^{\alpha}]+\int_{0}^{T}\int_{\Omega}y\left(-^{abc}_{T}D_{t}^{\alpha}v-\lambda\triangle v\right)dxdt\\ &\quad+\frac{B(\alpha)}{1-\alpha}\int_{\Omega}v(x,T)\int_{0}^{T}yE_{\alpha,\alpha}[-\gamma(T-t)^{\alpha}]dtdx.\end{split}

We proceed similarly as in [15]. Let VmV_{m} define a subspace of H1(Ω)H^{1}(\Omega) generated by the w1w_{1}, w2w_{2}, \ldots, wmw_{m} space vectors of orthogonal eigenfunctions of the operator Δ\Delta. We seek um:t(0,T]um(t)Vmu_{m}:t\in(0,T]\rightarrow u_{m}(t)\in V_{m}, solution of the fractional differential equation

{ΩabcDtα0ymv𝑑x+Ωymvdx=(f(ym),v) for all vVm,ym(x,0)=y0m for xΩ.\begin{cases}\displaystyle\int_{\Omega}\,^{abc}{}_{0}D_{t}^{\alpha}y_{m}vdx+\int_{\Omega}\nabla y_{m}\nabla vdx=(f(y_{m}),v)&\mbox{ for all }v\in V_{m},\\ y_{m}(x,0)=y_{0m}&\mbox{ for }x\in\Omega.\end{cases}

To continue the proof of existence, we recall the following auxiliary result.

Theorem 4.2 (See [15]).

Let α(0,1)\alpha\in(0,1). Assume that fL2(QT)f\in L^{2}(Q_{T}), y0L2(Ω)y_{0}\in L^{2}(\Omega). Let (,)(\cdot,\cdot) be the scalar product in L2(Ω)L^{2}(\Omega) and a(,)a(\cdot,\cdot) be the bilinear form in H01(Ω)H^{1}_{0}(\Omega) defined by

a(ϕ,ψ)=Ωϕ(x)ψ(x)𝑑xϕ,ψH1(Ω).a(\phi,\psi)=\int_{\Omega}\nabla\phi(x)\nabla\psi(x)dx\quad\forall\phi,\psi\in H^{1}(\Omega).

Then the problem

{(Dtα0abcy,v)+a(y(t),v)=(f(t),v), for all t(0,T),y(x,0)=y0, for xΩ,\begin{cases}\left({}^{abc}_{0}D_{t}^{\alpha}y,v\right)+a(y(t),v)=(f(t),v),&\mbox{ for all }t\in(0,T),\\ y(x,0)=y_{0},&\mbox{ for }x\in\Omega,\end{cases}

has a unique solution yL2(0,T,H01(Ω))C(0,T,L2(Ω))y\in L^{2}(0,T,H_{0}^{1}(\Omega))\bigcap{C}(0,T,L^{2}(\Omega)) given by

y(x,t)=j=1+(ζjEα[γjtα]yj0+(1α)ζjB(α)fj(t)+Kj0t(ts)α1Eα,α[γj(ts)α]fj(s)ds)wj,y(x,t)=\sum_{j=1}^{+\infty}\bigg{(}\zeta_{j}E_{\alpha}[-\gamma_{j}t^{\alpha}]y^{0}_{j}+\frac{(1-\alpha)\zeta_{j}}{B(\alpha)}f_{j}(t)\\ +K_{j}\int_{0}^{t}(t-s)^{\alpha-1}E_{\alpha,\alpha}\left[-\gamma_{j}(t-s)^{\alpha}\right]f_{j}(s)ds\bigg{)}w_{j}, (11)

where γj\gamma_{j} and ζj\zeta_{j} are constants. Moreover, provided y0L2(Ω)y_{0}\in L^{2}(\Omega), yy satisfies the inequalities

yL2(0,T,H01(Ω))μ1(y0H01(Ω)+fL2(QT))\|y\|_{L^{2}(0,T,H_{0}^{1}(\Omega))}\leq\mu_{1}(\|y_{0}\|_{H_{0}^{1}(\Omega)}+\|f\|_{L^{2}(Q_{T})}) (12)

and

yL2(Ω))μ2(y0L2(Ω)+fL2(QT)),\|y\|_{L^{2}(\Omega))}\leq\mu_{2}(\|y_{0}\|_{L^{2}(\Omega)}+\|f\|_{L^{2}(Q_{T})}), (13)

where μ1\mu_{1} and μ2\mu_{2} are positive constants.

Since f(ym)L2(QT)f(y_{m})\in L^{2}(Q_{T}), Theorem 4.2 implies that ymy_{m} is given in a explicit form. The existence of a solution is obtained by using the a priori estimate of Theorem 4.2 and the same arguments used to pass to the limit as those used by us below in the proof of Theorem 5.1.

5. Existence of an optimal control

We prove existence of an optimal control by using minimizing sequences.

Theorem 5.1.

There exists at least an optimal solution y(u)L(QT)y^{*}(u^{*})\in L^{\infty}(Q_{T}) satisfying (4)–(6) and minimizing (7).

Proof.

Let (yn,un)(y^{n},u^{n}) be a minimizing sequence of J(y,u)J(y,u) such that

limn+J(yn,un)=J(y,u)=infJ(y,u)\lim_{n\rightarrow+\infty}J(y^{n},u^{n})=J(y^{*},u^{*})=\inf J(y,u)

with un,uUadu^{n},u\in U_{ad} and yn=(y1n,y2n,y3n)y^{n}=(y_{1}^{n},y_{2}^{n},y_{3}^{n}) satisfying the corresponding system to (4)

Dtα0abcynλyn=f(yn), in QT=Ω×(0,T),ynν=0, on ΣT=Ω×(0,T),yn(0,x)=y0, in Ω.\begin{split}{}^{abc}_{0}D_{t}^{\alpha}y^{n}-\lambda\triangle y^{n}&=f(y^{n})\,,\mbox{ in }Q_{T}=\Omega\times(0,T),\\ \frac{\partial y^{n}}{\partial\nu}&=0\,,\mbox{ on }\Sigma_{T}=\partial\Omega\times(0,T),\\ y^{n}(0,x)&=y^{0}\,,\mbox{ in }\Omega.\end{split}

By Theorem 4.2, we know that (yn)(y_{n}) is bounded, independently of nn, in L2(0,T,H1(Ω))L^{2}(0,T,H^{1}(\Omega)) and satisfying the inequalities

ynL2(0,T,H01(Ω))μ1(y0H01(Ω)+fL2(QT))\|y^{n}\|_{L^{2}(0,T,H_{0}^{1}(\Omega))}\leq\mu_{1}(\|y^{0}\|_{H_{0}^{1}(\Omega)}+\|f\|_{L^{2}(Q_{T})}) (14)

and

ynL2(Ω))μ2(y0L2(Ω)+fL2(QT)),\|y^{n}\|_{L^{2}(\Omega))}\leq\mu_{2}(\|y^{0}\|_{L^{2}(\Omega)}+\|f\|_{L^{2}(Q_{T})}), (15)

where μ1\mu_{1} and μ2\mu_{2} are positive constants. Then (yn)(y^{n}) is bounded in L(0,T,L2(Ω))L^{\infty}(0,T,L^{2}(\Omega)) and L2(0,T,H01(Ω))L^{2}(0,T,H_{0}^{1}(\Omega)). By using the boundedness of yiny_{i}^{n} (|yi|N|y_{i}|\leq N, for i=1,2,3i=1,2,3), the second member ff is in L(QT)L^{\infty}(Q_{T}). Then, we have, for a positive constant independent of nn, that

0abcDtαynλynL2(QT)c.\|^{abc}_{0}D_{t}^{\alpha}y^{n}-\lambda\triangle y^{n}\|_{L^{2}(Q_{T})}\leq c.

Therefore, there exists a subsequence of yny^{n}, still denoted by (yn)(y^{n}), and unUadu^{n}\in U_{ad} such that

Dtα0abcynλynδ weakly in L2(QT),yny weakly in L2(0,T,H01(Ω)).\begin{gathered}{}^{abc}_{0}D_{t}^{\alpha}y^{n}-\lambda\triangle y^{n}\rightharpoonup\delta\mbox{ weakly in }L^{2}(Q_{T}),\\ y^{n}\rightharpoonup y^{*}\mbox{ weakly in }L^{2}(0,T,H_{0}^{1}(\Omega)).\end{gathered}

We now show that ynt\frac{\partial y^{n}}{\partial t} is bounded in L1(0,T,H1(Ω))L^{1}(0,T,H^{-1}(\Omega)). We shall use the following lemma.

Lemma 5.2.

If uL(0,T,L2(Ω))H1(0,T,L1(Ω))u\in L^{\infty}(0,T,L^{2}(\Omega))\cap H^{1}(0,T,L^{1}(\Omega)), then there exists a positive constant cc such that

tuL1(0,T,L1(Ω))cEα(γTα)uL(0,T,L2(Ω)).\|\partial_{t}u\|_{L^{1}(0,T,L^{1}(\Omega))}\leq\frac{c}{E_{\alpha}(-\gamma T^{\alpha})}\|u\|_{L^{\infty}(0,T,L^{2}(\Omega))}.
Proof.

Since for 0stT0\leq s\leq t\leq T, tEα(t)t\rightarrow E_{\alpha}(-t) is completely monotonic, we have

Eα(γTα)Eα(γ(ts)α).E_{\alpha}(-\gamma T^{\alpha})\leq E_{\alpha}(-\gamma(t-s)^{\alpha}).

It yields that

Eα(γTα)0t|su|𝑑s0t|su|Eα(γ(ts)α)𝑑s.E_{\alpha}(-\gamma T^{\alpha})\int_{0}^{t}|\partial_{s}u|ds\leq\int_{0}^{t}|\partial_{s}u|E_{\alpha}(-\gamma(t-s)^{\alpha})ds.

Using the well-known inequality 0abcDutαL(0,T)B(α)1αuL(0,T)\|^{abc}_{0}D{{}_{t}^{\alpha}}u\|_{L^{\infty}(0,T)}\leq\frac{B(\alpha)}{1-\alpha}\|u\|_{L^{\infty}(0,T)}, we get

Eα(γTα)B(α)1α0t|su|𝑑s\displaystyle E_{\alpha}(-\gamma T^{\alpha})\frac{B(\alpha)}{1-\alpha}\int_{0}^{t}|\partial_{s}u|ds B(α)1α0t|su|Eα(γ(ts)α)𝑑s\displaystyle\leq\frac{B(\alpha)}{1-\alpha}\int_{0}^{t}|\partial_{s}u|E_{\alpha}(-\gamma(t-s)^{\alpha})ds (16)
0abcDutαL(0,T)B(α)1αuL(0,T).\displaystyle\leq\|^{abc}_{0}D{{}_{t}^{\alpha}}u\|_{L^{\infty}(0,T)}\leq\frac{B(\alpha)}{1-\alpha}\|u\|_{L^{\infty}(0,T)}.

It follows that

0T|su|𝑑s1Eα(γTα)uL(0,T).\int_{0}^{T}|\partial_{s}u|ds\leq\frac{1}{E_{\alpha}(-\gamma T^{\alpha})}\|u\|_{L^{\infty}(0,T)}.

Integrating over Ω\Omega, we have

0TΩ|su|𝑑s𝑑x1Eα(γTα)ΩuL(0,T)𝑑x.\int_{0}^{T}\int_{\Omega}|\partial_{s}u|dsdx\leq\frac{1}{E_{\alpha}(-\gamma T^{\alpha})}\int_{\Omega}\|u\|_{L^{\infty}(0,T)}dx.

Then, for a positive constant cc, one has

tuL1(0,T,L1(Ω))\displaystyle\|\partial_{t}u\|_{L^{1}(0,T,L^{1}(\Omega))} 1Eα(γTα)uL(0,T,L1(Ω))\displaystyle\leq\frac{1}{E_{\alpha}(-\gamma T^{\alpha})}\|u\|_{L^{\infty}(0,T,L^{1}(\Omega))} (17)
cEα(γTα)uL(0,T,L2(Ω)).\displaystyle\leq\frac{c}{E_{\alpha}(-\gamma T^{\alpha})}\|u\|_{L^{\infty}(0,T,L^{2}(\Omega))}.

The proof is complete. ∎

By the estimate (15) of yny^{n} and Lemma 5.2, we have that tyn\partial_{t}y^{n} is bounded in L1(0,T,L1(Ω))L^{1}(0,T,L^{1}(\Omega)). Due to (14), we have that yny^{n} is bounded in L2(0,T,H01(Ω))L^{2}(0,T,H_{0}^{1}(\Omega)). Set

W={vL2(0,T,H01(Ω)),tvL1(0,T,L1(Ω))}.W=\{v\in L^{2}(0,T,H_{0}^{1}(\Omega)),\partial_{t}v\in L^{1}(0,T,L^{1}(\Omega))\}.

Using the classical argument of Aubin, the space WW is compactly embedded in L2(0,T,L2(Ω))=L2(QT)L^{2}(0,T,L^{2}(\Omega))=L^{2}(Q_{T}). We can then extract a subsequence from yny^{n}, not relabeled, such that

yny weakly in L(0,T,L2(Ω)) and in L2(QT),yny strongly in L2(QT),yny a. e. in L2(QT),yn(T)y(T) in L2(Ω).\begin{gathered}y^{n}\rightharpoonup y^{*}\mbox{ weakly in }L^{\infty}(0,T,L^{2}(\Omega))\mbox{ and in }L^{2}(Q_{T}),\\ y^{n}\rightarrow y^{*}\mbox{ strongly in }L^{2}(Q_{T}),\\ y^{n}\rightarrow y^{*}\mbox{ a. e. in }L^{2}(Q_{T}),\\ y^{n}(T)\rightarrow y^{*}(T)\mbox{ in }L^{2}(\Omega).\end{gathered}

Denote 𝔻(QT)\mathbb{D}^{\prime}(Q_{T}) the dual of 𝔻(QT)\mathbb{D}(Q_{T}), the set of C{C}^{\infty} functions on QTQ_{T} with compact support. We claim that

Dtα0abcynλynDtα0abcyλy weakly in 𝔻(QT).{}^{abc}_{0}D_{t}^{\alpha}y^{n}-\lambda\triangle y^{n}\rightharpoonup{{}^{abc}_{0}D_{t}^{\alpha}y^{*}}-\lambda\triangle y^{*}\mbox{ weakly in }\mathbb{D}^{\prime}(Q_{T}).

Indeed, we have

0TΩyn(0abcDtαvλv)dxdt0TΩy(TabcDtαvλv)dxdt,v𝔻(QT)\int_{0}^{T}\int_{\Omega}y^{n}(^{abc}_{0}D_{t}^{\alpha}v-\lambda\triangle v)dxdt\rightarrow\int_{0}^{T}\int_{\Omega}y^{*}(-^{abc}_{T}D_{t}^{\alpha}v-\lambda\triangle v)dxdt,\forall v\in\mathbb{D}(Q_{T})

and

Ωv(x,T)0TynEα,α[γ(Tt)α]𝑑t𝑑xΩv(x,T)0TyEα,α[γ(Tt)α]𝑑t𝑑x.\int_{\Omega}v(x,T)\int_{0}^{T}y^{n}E_{\alpha,\alpha}[-\gamma(T-t)^{\alpha}]dtdx\rightarrow\int_{\Omega}v(x,T)\int_{0}^{T}y^{*}E_{\alpha,\alpha}[-\gamma(T-t)^{\alpha}]dtdx.

On the other hand, the convergence yinyiy_{i}^{n}\rightarrow y_{i}^{*} in L2(QT)L^{2}(Q_{T}) and the essential boundedness of y1ny_{1}^{n} and y2ny_{2}^{n} imply y1ny2ny1y2y_{1}^{n}y_{2}^{n}\rightarrow y_{1}^{*}y_{2}^{*} in L2(QT)L^{2}(Q_{T}). Modulo a subsequence denoted unu^{n}, we have

unu weakly in L2(QT).u^{n}\rightharpoonup u^{*}\text{ weakly in }L^{2}(Q_{T}).

We deduce that uUadu^{*}\in U_{ad} as a consequence of the closure and the boundedness of this set in L2(QT)L^{2}(Q_{T}) and thus it is weakly closed. Similarly, we can prove that

uny1nuy1 in L2(QT).u^{n}y_{1}^{n}\rightarrow u^{*}y_{1}^{*}\text{ in }L^{2}(Q_{T}).

Therefore,

Dtα0abcynλyn0abcDtαyy weakly in 𝔻(QT).{}^{abc}_{0}D_{t}^{\alpha}y^{n}-\lambda\triangle y^{n}\rightarrow^{abc}_{0}D_{t}^{\alpha}y^{*}-\triangle y^{*}\mbox{ weakly in }\mathbb{D}^{\prime}(Q_{T}).

From the uniqueness of the limit, we have

Dtα0abcyy=δ.{}^{abc}_{0}D_{t}^{\alpha}y^{*}-\triangle y^{*}=\delta.

By passing to the limit as nn\rightarrow\infty in the equation satisfied by yny^{n}, we deduce that yy^{*} is a solution of (4). Finally, the lower semi-continuity of JJ leads to J(y,u)=infJ(y,u)J(y^{*},u^{*})=\inf J(y,u). Therefore, y(u)y^{*}(u^{*}) is an optimal solution. ∎

6. Necessary optimality conditions

In this section, our aim is to obtain optimality conditions. As we shall see, our necessary optimality conditions involve an adjoint system defined by means of the backward Atangana–Baleanu fractional-time derivative.

Let y(u)y^{*}(u^{*}) be an optimal solution and uε=u+εuUadu^{\varepsilon}=u^{*}+\varepsilon u\in U_{ad} be a control function such that uUadu\in U_{ad} and ε>0\varepsilon>0. Denote yε=(y1ε,y2ε,y3ε)=(y1,y2,y3)(uε)y^{\varepsilon}=(y_{1}^{\varepsilon},y_{2}^{\varepsilon},y_{3}^{\varepsilon})=(y_{1},y_{2},y_{3})(u^{\varepsilon}) and y=(y1,y2,y3)=(y1,y2,y3)(u)y^{*}=(y_{1}^{*},y_{2}^{*},y_{3}^{*})=(y_{1},y_{2},y_{3})(u^{*}) the solutions of (4)–(6) corresponding to uεu^{\varepsilon} and uu^{*}, respectively. Setting yε=y+εzεy^{\varepsilon}=y^{*}+\varepsilon z^{\varepsilon} and subtracting the system corresponding to yy^{*} from the one corresponding to yεy^{\varepsilon}, we have

abc0Dtα(yεyε)λ(yεyε)=f(yε)f(y)ε.^{abc}_{0}D_{t}^{\alpha}\left(\frac{y^{\varepsilon}-y^{*}}{\varepsilon}\right)-\lambda\triangle\left(\frac{y^{\varepsilon}-y^{*}}{\varepsilon}\right)=\dfrac{f(y^{\varepsilon})-f(y^{*})}{\varepsilon}. (18)

System (18) can be rewritten as

Dtα0abczελzε=f(yε)f(y)ε,{}^{abc}_{0}D_{t}^{\alpha}z^{\varepsilon}-\lambda\triangle z^{\varepsilon}=\dfrac{f(y^{\varepsilon})-f(y^{*})}{\varepsilon},

associated to Neumann boundary conditions

z1εν=z2εν=z3εν=0 on ΣT\frac{\partial z_{1}^{\varepsilon}}{\partial\nu}=\frac{\partial z_{2}^{\varepsilon}}{\partial\nu}=\frac{\partial z_{3}^{\varepsilon}}{\partial\nu}=0\text{ on }\Sigma_{T}

and initial condition

zε=0 in Ω,z^{\varepsilon}=0\text{ in }\Omega\,,

where zε=(z1ε,z2ε,z3ε)z^{\varepsilon}=(z_{1}^{\varepsilon},z_{2}^{\varepsilon},z_{3}^{\varepsilon}) and

f(yε)f(y)ε={f1(yε)f1(y)ε=(μβy2εduε)z1ε(βy1+μ)z2ε+μz3εuy1,f2(yε)f2(y)ε=βy2εz1ε+(βy1dr)z2ε,f3(yε)f3(y)ε=uεz1ε+rz2εdz3ε+uy1.\dfrac{f(y^{\varepsilon})-f(y^{*})}{\varepsilon}=\begin{cases}\dfrac{f_{1}(y^{\varepsilon})-f_{1}(y^{*})}{{\varepsilon}}=&(\mu-\beta y_{2}^{\varepsilon}-d-u^{\varepsilon})z_{1}^{\varepsilon}-(\beta y_{1}^{*}+\mu)z_{2}^{\varepsilon}\\ &\quad+\mu z_{3}^{\varepsilon}-uy_{1}^{*},\\ \dfrac{f_{2}(y^{\varepsilon})-f_{2}(y^{*})}{{\varepsilon}}=&\beta y_{2}^{\varepsilon}z_{1}^{\varepsilon}+(\beta y_{1}^{*}-d-r)z_{2}^{\varepsilon},\\ \dfrac{f_{3}(y^{\varepsilon})-f_{3}(y^{*})}{{\varepsilon}}=&u^{\varepsilon}z_{1}^{\varepsilon}+rz_{2}^{\varepsilon}-dz_{3}^{\varepsilon}+uy_{1}^{*}.\end{cases}

Set

Fε=(μβy2εduεβy1+μμβy2εβy1dr0uεrd)F^{\varepsilon}=\left(\begin{array}[]{ccc}\mu-\beta y_{2}^{\varepsilon}-d-u^{\varepsilon}&-\beta y_{1}^{*}+\mu&\mu\\ \beta y_{2}^{\varepsilon}&\beta y_{1}^{*}-d-r&0\\ u^{\varepsilon}&r&-d\\ \end{array}\right)

and

G=(y10y1).G=\left(\begin{array}[]{c}-y_{1}^{*}\\ 0\\ y_{1}^{*}\\ \end{array}\right).

Then, (18) can be reformulated in the following form:

{Dtα0abczελzε=Fεzε+Gu for t[0,T],zε(0)=0.\begin{cases}{}^{abc}_{0}D_{t}^{\alpha}z^{\varepsilon}-\lambda\triangle z^{\varepsilon}&=F^{\varepsilon}z^{\varepsilon}+Gu\text{ for }t\in[0,T],\\ z^{\varepsilon}(0)&=0.\end{cases}

Since the elements of the matrix FεF^{\varepsilon} are uniformly bounded with respect to ε\varepsilon and (uy1,0,uy1)(-uy_{1}^{*},0,uy_{1}^{*}) is bounded in L2(QT)L^{2}(Q_{T}), it follows by Theorem 4.2 that zε=yεyεz^{\varepsilon}=\frac{y^{\varepsilon}-y^{*}}{\varepsilon} is bounded in L(0,T,L2(Ω))L2(0,T,H1(Ω))L^{\infty}(0,T,L^{2}(\Omega))\bigcap L^{2}(0,T,H^{1}(\Omega)). Therefore, up to a subsequence of zεz^{\varepsilon}, there exists zz such that as ε\varepsilon tends to zero we have

zεz weakly in L(0,T,L2(Ω)) and in L2(0,T,H1(Ω)),zεtzt weakly in L2(0,T,H1Ω)).\begin{gathered}z^{\varepsilon}\rightarrow z\mbox{ weakly in }L^{\infty}(0,T,L^{2}(\Omega))\mbox{ and in }L^{2}(0,T,H^{1}(\Omega)),\\ \frac{\partial z^{\varepsilon}}{\partial t}\rightarrow\frac{\partial z}{\partial t}\mbox{ weakly in }L^{2}(0,T,H^{-1}\Omega)).\end{gathered} (19)

Put

F=(μβy2duβy1+μμβy2βy1dr0urd).F=\left(\begin{array}[]{ccc}\mu-\beta y_{2}^{*}-d-u^{*}&-\beta y_{1}^{*}+\mu&\mu\\ \beta y_{2}^{*}&\beta y_{1}^{*}-d-r&0\\ u^{*}&r&-d\\ \end{array}\right).

Note that all the components of the matrix FεF^{\varepsilon} tend to the corresponding ones of the matrix FF in L2(QT)L^{2}(Q_{T}) as ε0\varepsilon\rightarrow 0. From equations satisfied by yεy_{\varepsilon} and yy, we have that

Ω0TDtα0abczεv𝑑x𝑑t+Ω0Tλzεvdxdt=Ω0T(f(yε)f(y))εv𝑑x𝑑t.\int_{\Omega}\int_{0}^{T}\,{}^{abc}_{0}D_{t}^{\alpha}z^{\varepsilon}vdxdt+\int_{\Omega}\int_{0}^{T}\lambda\nabla z^{\varepsilon}\nabla vdxdt=\int_{\Omega}\int_{0}^{T}\,\dfrac{(f(y^{\varepsilon})-f(y))}{\varepsilon}vdxdt.

Letting ε0\varepsilon\rightarrow 0, we get

Ω0TDtα0abczv𝑑x𝑑t+Ω0Tλzvdxdt=Ω0T(Fz+Gu)v𝑑x𝑑t\int_{\Omega}\int_{0}^{T}\,{}^{abc}_{0}D_{t}^{\alpha}zvdxdt+\int_{\Omega}\int_{0}^{T}\lambda\nabla z\nabla vdxdt=\int_{\Omega}\int_{0}^{T}\,(Fz+Gu)vdxdt

with z(0)=0z(0)=0. By Green’s formula, it follows that

Ω0TDtα0abczv𝑑x𝑑tΩ0TλΔzv𝑑x𝑑t+0TΩzηv𝑑s𝑑t=Ω0T(Fz+Gu)v𝑑x𝑑t.\int_{\Omega}\int_{0}^{T}\,{}^{abc}_{0}D_{t}^{\alpha}z\cdot vdxdt-\int_{\Omega}\int_{0}^{T}\lambda\Delta z\cdot vdxdt+\int_{0}^{T}\int_{\partial\Omega}\frac{\partial z}{\partial\eta}vdsdt\\ =\int_{\Omega}\int_{0}^{T}\,(Fz+Gu)vdxdt.

Then zz verifies

Dtα0abczλz=Fz+Gu, in Ω,zν=0 on Ω,z(0)=0.\begin{gathered}{}^{abc}_{0}D_{t}^{\alpha}z-\lambda\triangle z=Fz+Gu,\mbox{ in }\Omega,\\ \frac{\partial z}{\partial\nu}=0\quad\mbox{ on }\partial\Omega,\\ z(0)=0.\end{gathered}

To derive the adjoint operator associated with zz, we need to introduce an enough smooth adjoint variable pp defined in QTQ_{T}. We have

Ω0T(Dtα0abczλz)p𝑑x𝑑t=QT(Fz+Gu)p𝑑x𝑑t.\int_{\Omega}\int_{0}^{T}\left({}^{abc}_{0}D_{t}^{\alpha}z-\lambda\triangle z\right)pdxdt=\int_{Q_{T}}(Fz+Gu)pdxdt.

Integrating by parts, one has

Ω0TDtα0abcz.pdtdx=Ω0TDtαTabcp.zdtdx+B(α)1αΩp(x,T)0TzEα,α[γ(Tt)α]𝑑t.\begin{split}\int_{\Omega}\int_{0}^{T}\,{}^{abc}_{0}D_{t}^{\alpha}z.pdtdx&=-\int_{\Omega}\int_{0}^{T}\,{}^{abc}_{T}D_{t}^{\alpha}p.zdtdx\\ &\quad+\frac{B(\alpha)}{1-\alpha}\int_{\Omega}p(x,T)\int_{0}^{T}zE_{\alpha,\alpha}[-\gamma(T-t)^{\alpha}]dt.\end{split}

We conclude that the adjoint function pp satisfies the adjoint system given by

TabcDtαpλpFp=DDy,t[0,T],pν=0 on Ω×(0,T),p(T,x)=DDy(T,x),\begin{gathered}-^{abc}_{T}D_{t}^{\alpha}p-\lambda\triangle p-Fp=D^{*}Dy^{*},\,t\in[0,T],\\ \frac{\partial p}{\partial\nu}=0\mbox{ on }\partial\Omega\times(0,T),\\ p(T,x)=D^{*}Dy^{*}(T,x),\end{gathered} (20)

where DD is the matrix defined by

(000010000).\left(\begin{array}[]{ccc}0&0&0\\ 0&1&0\\ 0&0&0\\ \end{array}\right).

Similarly to the existence result of Theorem 4.2, one can show that the solution of the adjoint system (20) exists.

Theorem 6.1.

Given an optimal control uu^{*} and corresponding state yy^{*}, there exists a solution pp to the adjoint system. Furthermore, uu^{*} can be characterized, in explicit form, as

u=min(1,max(0,1θGp))=min(1,max(0,y1θ(p1p3))).\begin{split}u^{*}=&\min\left(1,\max\left(0,-\frac{1}{\theta}G^{*}p\right)\right)\\ =&\min\left(1,\max\left(0,-\frac{y_{1}^{*}}{\theta}(p_{1}-p_{3})\right)\right).\end{split} (21)

The proof of Theorem 6.1 is classical and follows exactly the same arguments as in [17], using the fact that the minimum of the objective function JJ is achieved at uu^{*}. For small ε\varepsilon such that uε=u+εhUadu^{\varepsilon}=u^{*}+\varepsilon h\in U_{ad}, one can prove that

J(u+εh)J(u)ε0,\begin{split}\frac{J(u^{*}+\varepsilon h)-J(u)}{\varepsilon}&\geq 0,\end{split}

equivalent to

0T(Gp+θu,h)𝑑t0hUad.\begin{split}\int_{0}^{T}(G^{*}p+\theta u^{*},h)dt&\geq 0\quad\forall h\in U_{ad}.\end{split}

The characterization result is obtained by standard arguments of variations of hh.

7. Numerical results

In this section, we study the effect of the order of differentiation α\alpha to the dynamic of infection in space during a given time interval. We can mention two cases: absence and presence of vaccination. In the following, we consider a domain of 10km210km^{2} square grid, which represents a city for the population under consideration. We assume that the infection originates in the subdomain Ω1=cell(1,1)\Omega_{1}=cell(1,1) when the disease starts at the lower left corner of Ω\Omega. At t=0t=0, we assume that the susceptible people are homogeneously distributed with 5050 in each 1km21km^{2} cell except at the subdomain Ω1\Omega_{1} of Ω\Omega, where we introduce 77 infected individuals and keep 4343 susceptible there. The parameters and initial values are given in Table 1.

Table 1. Values of initial conditions and parameters.
Symbol Description (Unit) Value
S0(x,y)S_{0}(x,y) Initial susceptible population (people/km2)(people/km^{2}) 4343 for (x,y)Ω1(x,y)\in\Omega_{1}               5050 for (x,y)Ω1(x,y)\notin\Omega_{1}
I0(x,y)I_{0}(x,y) Initial infected population (people/km2)(people/km^{2}) 77 for (x,y)Ω1(x,y)\in\Omega_{1}               0 for (x,y)Ω1(x,y)\notin\Omega_{1}
R0(x,y)R_{0}(x,y) Initial recovered population (people/km2)(people/km^{2}) 0 for (x,y)Ω1(x,y)\in\Omega_{1}               0 for (x,y)Ω1(x,y)\notin\Omega_{1}
λ1=λ2=λ3\lambda_{1}=\lambda_{2}=\lambda_{3} Diffusion coefficient (km2/daykm^{2}/day) 0.6
μ\mu Birth rate (day1)(day^{-1}) 0.02
dd Natural death rate (day1)(day^{-1}) 0.03
β\beta Transmission rate ((people/km2)1.day1)((people/km^{2})^{-1}.day^{-1}) 0.9
rr Recovery rate (day1)(day^{-1}) 0.04
TT Final time (day)(day) 20

We have used MATLAB to implement the so-called forward-backward sweep method [18] to solve our fractional optimal control problem (4)–(8). The state system and the adjoint equations are numerically integrated using an approximation of the (left/right) ABC fractional derivative, based on a explicit finite difference method [29, 4]. The algorithm can be summarized as follows:

Algorithm 1 Forward-backward sweep method
1:  Set nn the number of subdivisions, hh the step size, mm the number of time steps, τ\tau the step time, δ=0.001\delta=0.001 the tolerance, and test=1test=-1.
2:  Initiate the control uoldu_{old}, the state (Sold,Iold,Rold)(S_{old},I_{old},R_{old}) and adjoint ((pold)1,(pold)2,(pold)3)((p_{old})_{1},(p_{old})_{2},(p_{old})_{3}).
3:  while test<0test<0 do
4:     Solve the state equation (4) for (S,I,R)(S,I,R) with initial guess (S0,I0,R0)(S_{0},I_{0},R_{0}), using an explicit finite difference method forward in time.
5:     Solve the adjoint equation (20) for (p1,p2,p3)(p_{1},p_{2},p_{3}) using the transversality conditions pi(T)p_{i}(T) and (S,I,R)(S,I,R) backward in time.
6:     Update the control using the gradient equation (21) to reach uu.
7:     Compute the tolerance criteria ψ1=δSSSold\psi_{1}=\delta\|S\|-\|S-S_{old}\|, ψ2=δIIIold\psi_{2}=\delta\|I\|-\|I-I_{old}\|, ψ3=δRRRold\psi_{3}=\delta\|R\|-\|R-R_{old}\|, ψ4=δp1p1(pold)1\psi_{4}=\delta\|p_{1}\|-\|p_{1}-(p_{old})_{1}\|, ψ5=δp2p2(pold)2\psi_{5}=\delta\|p_{2}\|-\|p_{2}-(p_{old})_{2}\|, ψ6=δp3p3(pold)3\psi_{6}=\delta\|p_{3}\|-\|p_{3}-(p_{old})_{3}\|, ψ7=δuuuold\psi_{7}=\delta\|u\|-\|u-u_{old}\|, and calculate test=min{ψ1,ψ2,ψ3,ψ4,ψ5,ψ6,ψ7}test=\min\{\psi_{1},\psi_{2},\psi_{3},\psi_{4},\psi_{5},\psi_{6},\psi_{7}\}.
8:  end while

7.1. Fractional α\alpha-dynamics without control

Figures 1, 2 and 3 present the numerical results with different values of α\alpha in the case of absence of control. We observe that the susceptible individuals are transferred to the infected class while the disease spreads from the lower left corner to the upper right corner. In Figure 1, for α=1\alpha=1, we can see that the epidemic takes 2020 days to cover the entire area (50 infected per cell in all Ω\Omega), but in Figures 2 and 3 this is not the case for α=0.95\alpha=0.95 and α=0.9\alpha=0.9. It is clear that the number of individuals infected is almost 4444 per cell in the upper right corner.

Refer to caption
Figure 1. Dynamic of the system without control for α=1\alpha=1.
Refer to caption
Figure 2. Dynamic of the system without control for α=0.95\alpha=0.95.
Refer to caption
Figure 3. Dynamic of the system without control for α=0.9\alpha=0.9.

7.2. Fractional α\alpha-dynamics with optimal vaccination strategy

We compare the infection prevalence over a period of 2020 days in the presence of the vaccination strategy. We note that the susceptible individuals are transferred to the recovered class (see Figures 4, 5 and 6). In Figure 4, we see that the number of infected people is 4040 per cell and 1010 per cell for recovered individuals. In Figures 5 and 6, we have almost 55 susceptible people per cell, 3535 infected people per cell, and 1010 recovered individuals per cell.

Refer to caption
Figure 4. Dynamic of the system with control for α=1\alpha=1.
Refer to caption
Figure 5. Dynamic of the system with control for α=0.95\alpha=0.95.
Refer to caption
Figure 6. Dynamic of the system with control for α=0.9\alpha=0.9.

Next, we investigate the effect of the order α\alpha to the value of the cost functional JJ in absence and presence of vaccination. Before that, we present the results in Table  2 and Table  3, respectively.

Table 2. Values of the cost functional JJ without control for different α\alpha.
α\alpha 0.9 0.95 1
J 7.4350e+047.4350e^{+04} 7.1586e+047.1586e^{+04} 7.7019e+047.7019e^{+04}
Table 3. Values of the cost functional JJ with control for different α\alpha.
α\alpha 0.9 0.95 1
J 4.9157e+044.9157e^{+04} 4.7489e+044.7489e^{+04} 5.2503e+045.2503e^{+04}

We note that the functional JJ decreases under the effect of vaccination for different values of α\alpha, and the value of JJ is optimal as α=0.95\alpha=0.95. Furthermore, the results obtained in fractional order cases show that the spread of the disease takes more than 20 days to cover the entire space with the same cost of the vaccination strategy in the case of integer derivatives.

8. Conclusion

In this study, we investigated the optimal vaccination strategy for a fractional SIR model. Interactions between susceptible, infected, and recovered are modeled by a system of partial differential equations with Atangana–Baleanu–Caputo fractional time derivative. We proved existence of solutions to our fractional parabolic state system as well as the existence of an optimal control. For a given objective functional JJ, an optimal control is characterized in terms of the corresponding state and adjoint variables. In order to control the infection, we have compared the dynamics of our system with different values of α\alpha. We noticed that the values of JJ decreases under the effect of vaccination for different values of α\alpha. Moreover, with the presence of an optimal vaccination strategy, we found that the smallest value of the cost-functional JJ is obtained when α=0.95\alpha=0.95. Then, an analysis of the proposed fractional order strategy with a well chosen fractional order α\alpha shows that it is more cost-effective than the classical strategy. Finally, the results obtained when α\alpha takes a fractional value show that the spread of the disease takes more than 2020 days to cover the entire space with the same cost of the vaccination strategy in the case of integer-order derivatives.

Acknowledgments

Torres is supported by Fundação para a Ciência e a Tecnologia (FCT), within project UIDB/04106/2020 (CIDMA).

References

  • [1] (MR3646671) [10.22436/jnsa.010.03.20] T. Abdeljawad and D. Baleanu, Integration by parts and its applications of a new nonlocal fractional derivative with Mittag–Leffler nonsingular kernel, J. Nonlinear Sci. Appl., 10 (2017), 1098–1107.
  • [2] (MR1930721) [10.1016/S0022-247X(02)00180-4] O. P. Agrawal, Formulation of Euler–Lagrange equations for fractional variational problems, J. Math. Anal. Appl., 272 (2002), 368–379.
  • [3] (MR3719831) [10.2298/AADM170428002A] R. Almeida, What is the best fractional derivative to fit data?, Appl. Anal. Discrete Math., 11 (2017), 358–368.
  • [4] (MR3511679) [10.22436/jnsa.009.06.17] R. T. Alqahtani, Atangana–Baleanu derivative with fractional order applied to the model of groundwater within an unconfined aquifer, J. Nonlinear Sci. Appl., 9, (2016), 3647–3654.
  • [5] [10.2298/TSCI160111018A] A. Atangana and D. Baleanu, New fractional derivatives with nonlocal and non–singular kernel: theory and application to heat transfer model, Thermal Science, 20 (2016), 763–769.
  • [6] [10.1140/epjp/i2018-12021-3] A. Atangana and J. F. Gómez–Aguilar, Decolonisation of fractional calculus rules: Breaking commutativity and associativity to capture more natural phenomena, Eur. Phys. J. Plus, 133 (2018), Art. 166, 22 pp.
  • [7] (MR3764302) [10.1515/fca-2017-0076] G. M. Bahaa, Fractional optimal control problem for variable–order differential systems, Fract. Calc. Appl. Anal., 20 (2017), 1447–1470.
  • [8] [10.1115/DETC2009-87008] R. K. Biswas and S. Sen, Numerical method for solving fractional optimal control problems, In Proceedings of the ASME 2009 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference, San Diego, CA, USA, (2010), 1205–1208.
  • [9] (MR2849617) [10.1177/1077546310373618] R. K. Biswas and S. Sen, Fractional optimal control problems: A pseudo-state-space approach, J. Vib. Control, 17 (2010), 1034–1041.
  • [10] (MR2849617) [10.1177/1077546310373618] R. K. Biswas and S. Sen, Fractional optimal control problems with specified final time, J. Comput. Nonlinear Dyn., 6 (2011), 021009.
  • [11] [10.1115/DETC2011-48045] R. K. Biswas and S. Sen, Fractional optimal control within Caputo’s derivative, In Proceedings of the ASME 2011 International Design Engineering Technical Conferences and Computers and Information in Engineering Conference, Washington, DC, USA, (2012), 353–360.
  • [12] (MR3151796) [10.1016/j.jfranklin.2013.09.024] R. K. Biswas and S. Sen, Free final time fractional optimal control problems, J. Frankl. Inst., 351 (2014), 941–951.
  • [13] (MR3030124) [10.1007/s11071-012-0475-2] K. Diethelm, A fractional calculus based model for the simulation of an outbreak of dengue fever, Nonlinear Dyn., 71 (2013), 613–619.
  • [14] [10.1109/TCST.2011.2153203] Y. Ding, Z. Wang and H. Ye, Optimal control of a fractional-order HIV-immune system with memory, IEEE Trans. Control Syst. Technol., 20 (2011), 763–769.
  • [15] (MR3968302) [10.1007/s10957-018-1305-6] J. D. Djida, G. M. Mophou and I. Area, Optimal control of diffusion equation with fractional time derivative with nonlocal and nonsingular Mittag-Leffler kernel, J. Optim. Theory Appl. 182 (2019), 540–557.
  • [16] (MR3019305) [10.1007/s10957-012-0233-0] T. L. Guo, The necessary conditions of fractional optimal control in the sense of Caputo, J. Optim. Theory Appl., 156 (2013), 115–126.
  • [17] (MR3985793) [10.1007/s40435-019-00525-w] A. A. Laaroussi, R. Ghazzali, M. Rachik and S. Benrhila, Modeling the spatiotemporal transmission of Ebola disease and optimal control: a regional approach, Int. J. Dyn. Control, 7 (2019), no. 3, 1110–1124.
  • [18] (MR2316829) S. Lenhart and J. T. Workman, Optimal Control Applied to Biological Models, Chapman and Hall, CRC Press, 2007.
  • [19] (MR0259693) J-L. Lions Quelques méthodes de résolution des problèmes aux limites non linéaires, Dunod, Gauthier-Villars, Paris 1969.
  • [20] (MR2739436) [10.1016/j.camwa.2010.10.030] G. M. Mophou, Optimal control of fractional diffusion equation, Comput. Math. Appl. 61 (2011), 68–78.
  • [21] (MR2824729) [10.1016/j.camwa.2011.04.044] G. M. Mophou and  G. M. N’Guérékata, Optimal control of fractional diffusion equation with state constraints, Comput. Math. Appl. 62 (2011), 1413–1426.
  • [22] E. Okyere, F. T. Oduro, S. K. Amponsah and I. K. Dontwi, Fractional order optimal control model for malaria infection, arXiv preprint https://arxiv.org/abs/1607.01612, (2016).
  • [23] (MR1658022) I. Podlubny, Fractional Differential Equations, Mathematics in Science and Engineering, 198, Academic Press, Inc., San Diego, CA, 1999.
  • [24] (MR3872489) [10.1016/j.chaos.2018.10.021] S. Rosa and  D. F. M. Torres, Optimal control of a fractional order epidemic model with application to human respiratory syncytial virus infectio, Chaos Solitons Fractals 117 (2018), 142–149. \arXiv1810.06900
  • [25] (MR4232864) [10.1007/s11786-020-00467-z] M. R. Sidi Ammi, M. Tahiri and D. F. M. Torres, Global stability of a Caputo fractional SIRS model with general incidence rate, Math. Comput. Sci. 15 (2021), no. 1, 91–105. \arXiv2002.02560
  • [26] (MR3988048) [10.1016/j.camwa.2019.03.043] M. R. Sidi Ammi and  D. F. M. Torres, Optimal control of a nonlocal thermistor problem with ABC fractional time derivatives, Comput. Math. Appl. 78 (2019), no. 5, 1507–1516. \arXiv1903.07961
  • [27] (MR3101449) [10.1016/j.mbs.2013.05.005] C. J. Silva and D. F. M. Torres, Optimal control for a tuberculosis model with reinfection and post-exposure interventions, Math. Biosci. 244, (2013), no. 2, 154–164. \arXiv1305.2145
  • [28] (MR3395619) [10.1186/s13662-015-0593-5] Q. Tang and QX. Ma, Variational formulation and optimal control of fractional diffusion equations with Caputo derivatives, Adv. Diff. Equa. 2015 (2015), Art. 283, 14 pp.
  • [29] (MR3876254) [10.1016/j.chaos.2018.11.009] S. Yadav, R.K. Pandey and A.K. Shukla, Numerical approximations of Atangana–Baleanu Caputo derivative and its application, Chaos, Solitons and Fractals, 118, (2019), 58–64.
  • [30] [10.1109/CCDC.2015.7162031] J. Yuan, B. Shi, D. Zhang and S. Cui, A formulation for fractional optimal control problems via Left and Right Caputo derivatives, The 27th Chinese Control and Decision Conference (2015 CCDC), (2015), 816–821.