Next Article in Journal
Triclustering Implementation Using Hybrid δ-Trimax Particle Swarm Optimization and Gene Ontology Analysis on Three-Dimensional Gene Expression Data
Next Article in Special Issue
On the Fractional Derivative Duality in Some Transforms
Previous Article in Journal
Young Duality for Variational Inequalities and Nonparametric Method of Demand Analysis in Input–Output Models with Inputs Substitution: Application for Kazakhstan Economy
Previous Article in Special Issue
New Stability Results for Abstract Fractional Differential Equations with Delay and Non-Instantaneous Impulses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pontryagin Maximum Principle for Incommensurate Fractional-Orders Optimal Control Problems

by
Faïçal Ndaïrou
and
Delfim F. M. Torres
*
Center for Research and Development in Mathematics and Applications (CIDMA), Department of Mathematics, University of Aveiro, 3810-193 Aveiro, Portugal
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(19), 4218; https://doi.org/10.3390/math11194218
Submission received: 12 September 2023 / Revised: 7 October 2023 / Accepted: 8 October 2023 / Published: 9 October 2023
(This article belongs to the Special Issue Recent Research on Fractional Calculus: Theory and Applications)

Abstract

:
We introduce a new optimal control problem where the controlled dynamical system depends on multi-order (incommensurate) fractional differential equations. The cost functional to be maximized is of Bolza type and depends on incommensurate Caputo fractional-orders derivatives. We establish continuity and differentiability of the state solutions with respect to perturbed trajectories. Then, we state and prove a Pontryagin maximum principle for incommensurate Caputo fractional optimal control problems. Finally, we give an example, illustrating the applicability of our Pontryagin maximum principle.

1. Introduction

The celebrated Pontryagin maximum principle was first formulated in 1956, being widely regarded as the central result of optimal control theory [1]. The significance of the maximum principle lies in the fact that rather than maximizing over a function space, the problem is converted to a pointwise optimization. Various generalizations of the Pontryagin maximum principle are still being investigated nowadays [2,3,4].
Fractional optimal control theory, as a branch of applied mathematics, deals with optimization issues for controlled fractional dynamics coupled with a performance index functional [5,6]. The development of the theory started shortly at the beginning of 21th century with some pioneering works [7,8], where necessary optimality conditions are derived using techniques from variational analysis. A major contribution from those works is weak versions of the Pontryagin maximum principle that consists in reducing the fractional optimal control problem (FOCP) to a fractional boundary value problem together with an optimality condition. Later, the theory emerged with the work of [9], by establishing a strong maximum principle of Pontryagin type, which introduces a Hamiltonian maximality condition in the set of necessary optimality conditions instead of the optimality condition from weak versions. Indeed, since this seminal work of 2014, the subject of maximum principle for FOCPs gained more interest. For instance, a new approach to the Pontryagin maximum principle for a problem involving a Lagrangian cost of Riemann–Liouville fractional integral subject to Caputo fractional dynamics is considered in [10]. The authors of [11], derived a Pontryagin maximum principle for FOCP defined by a general Bolza cost of fractional integral type with terminal constraints on the Caputo fractional dynamics. More recently, in 2022, Kamocki investigated an optimal control problem of multi-order fractional systems under a Lagrange-type functional, proving two existence results [12]. In contrast with Kamocki, here we are interested in proving the necessary optimality conditions of Pontryagin type.
Another important development in control theory is the so-called Bellman’s dynamic programming principle, which gives a necessary and sufficient condition of optimality [13]. In this direction, the approach for solving an optimal control problem can be achieved by determining a certain value function that might happen to be a viscosity solution to a Hamilton–Jacobi–Bellman (HJB) equation. The dynamic programming principle has been extended for fractional discrete-time systems by the authors of references [14,15]. Furthermore, an attempt to derive a fractional version of the Hamilton–Jacobi–Bellman (HJB) equation has been investigated in [16]. In [17], Gomoyunov studies an extension of the dynamic programming principle to the case of fractional-order dynamical systems. A new approach is required for such problems so that for every intermediate time t ( 0 , T ) , it is necessary to introduce an auxiliary optimal control problem (sub-problem) with this time t considered as the initial one. Also, a derivation of the fractional version of the HJB equation is studied deeply [17]. To sum up, it is important to mention that the Pontryagin maximum principle, along with Bellman’s dynamic programming principle, is one of the most effective and fundamental tools for investigating solutions to various optimal control problems.
Nowadays, there are many different definitions of fractional-order integrals or derivatives [18] and, in some sense, it is possible to consider a single broad class of fractional-order operators that include existing ones as particular cases [19]. This is important in analyzing a single operator rather than focusing on each individual one separately. Also, it enriches the subject of Pontryagin’s maximum principle to handle a more general and wide class of fractional-order operators. For instance, a maximum principle is obtained for a combined fractional operator with a general analytic kernel in [20]. Also, some recent results for the Pontryagin maximum principle are investigated for fractional stochastic delayed systems with non-instantaneous impulses [21], for a degenerate fractional differential equation [22], and for distributed-order fractional derivatives [23]. In contrast, the main aim of our current study is to utilize the idea of multi-order or incommensurate orders of derivatives in the definition of optimal control problems and then analyze their solutions. In doing so, we start with the most popular fractional model, which is the Caputo.
The structure of the article is as follows. In Section 2, we present some basic definitions and properties from fractional calculus. In Section 3, our contribution is given: we start by introducing the incommensurate non-local fractional optimal control problem, then we prove the continuity of solutions (Lemma 3), differentiability of perturbed trajectories (Lemma 4), and, finally, the proof of the Pontryagin maximum principle (Theorem 1). In Section 4, we give an example illustrating the application of the new Pontryagin maximum principle. We end with Section 5, summarizing our work and giving some perspectives for future work.

2. Preliminaries

In this section, we briefly recall the necessary notions and results from fractional calculus. For more on the subject, we refer the interested readers to the books [24,25,26].
Let ( α i ) i = 1 , , n ( 0 , 1 ) be a multi-order of real numbers. In the sequel, we use the following notation:
L ( α ) [ a , b ] , R n : = x L 1 [ a , b ] , R n : I a + α i x i , I b α i x i A C [ a , b ] , R ,
where I a + α i and I b α i represent, respectively, the left and right Riemann–Liouville integrals of order α i . We also use the notation A C ( α ) [ a , b ] , R n to represent the set of absolutely continuous functions that can be represented as
x i ( t ) = x i ( a ) + I a + α i f ( t ) and x i ( t ) = x i ( b ) + I b α i f ( t ) ,
for some functions f L α i : = x L 1 [ a , b ] , R : I a + α i x i , I b α i x i A C [ a , b ] , R .
Definition 1.
The left- and right-sided Riemann–Liouville incommensurate fractional derivative of orders ( α i ) i = 1 , , n ( 0 , 1 ) of a function x L ( α ) are defined, respectively, by
D a + ( α ) x ( t ) = D a + α 1 x 1 ( t ) , D a + α 2 x 2 ( t ) , D a + α n x n ( t ) , ; D b ( α ) x ( t ) = D b α 1 x 1 ( t ) , D b α 2 x 2 ( t ) , D b α n x n ( t ) ,
with
D a + α i x i ( t ) = d d t I a + 1 α i x i ( t ) , D b α i x i ( t ) = d d t I b 1 α i x i ( t ) ,
where I a + 1 α i and I b 1 α i represent, respectively, the left- and right-sided Riemann–Liouville fractional integrals of order 1 α i .
Definition 2.
The left- and right-sided Caputo incommensurate fractional derivatives of order ( α i ) i = 1 , , n ( 0 , 1 ) of a function x A C ( α ) are defined, respectively, by
c D a + ( α ) x ( t ) = c D a + α 1 x 1 ( t ) , c D a + α 2 x 2 ( t ) , c D a + α n x n ( t ) , ; c D b ( α ) x ( t ) = c D b α 1 x 1 ( t ) , c D b α 2 x 2 ( t ) , c D b α n x n ( t ) ,
where
D a + α i x i ( t ) = I a + 1 α i d d t x i ( t ) , D b α i x i ( t ) = I b 1 α i d d t x i ( t ) .
Note that integration by parts is a powerful tool when two functions are multiplied together, being useful for our purposes in the proof of the Pontryagin maximum principle. In the sequel, the dot · is used for indicating scalar products.
Lemma 1
(Integration by parts formula [25]). Let x L ( α ) and y A C ( α ) . Then,
a b x ( t ) · c D a + ( α ) y ( t ) d t = y ( t ) · I b 1 ( α ) x ( t ) a b + a b y ( t ) · D b ( α ) x ( t ) d t ,
where
I b 1 ( α ) x ( t ) = I b 1 α 1 x 1 ( t ) , I b 1 α 2 x 2 ( t ) , I b 1 α n x n ( t ) .
In what follows, we recall a generalized Gronwall inequality that is useful to prove the continuity and differentiability of perturbed trajectories.
Lemma 2
(Generalized Gronwall inequality [27]). Let α be a positive real number and let p ( · ) , q ( · ) , and u ( · ) be non-negative continuous functions on [ a , b ] with q ( · ) monotonic increasing on [ a , b ) . If
u ( t ) p ( t ) + q ( t ) a t ( t s ) α 1 u ( s ) d s ,
then
u ( t ) p ( t ) + a t n = 1 q ( t ) Γ ( α ) n Γ ( n α ) ( t s ) n α 1 p ( s ) d s
for all t [ a , b ) .

3. Main Results

In this section, our main concern is to find a control function u L [ a , b ] , R m and its corresponding state trajectory x A C ( α ) [ a , b ] , R n , solution to the following incommensurate non-local fractional optimal control problem:
J [ x ( · ) , u ( · ) ] = φ ( b , x ( b ) ) + a b L t , x ( t ) , u ( t ) d t max , c D a + ( α ) x ( t ) = f t , x ( t ) , u ( t ) , t [ a , b ] a . e . , x ( · ) A C ( α ) , u ( · ) L , x ( a ) = x a R n , u ( t ) Ω ,
where Ω is a closed subset of R m . The data functions L : [ a , b ] × R n × R m R , φ : R n R and f : [ a , b ] × R n × R m R n are subject to the following assumptions:
  • The function φ is of class C 1 .
  • Functions L and f are continuous in all its three arguments and of class C 1 with respect to x and, in particular, locally Lipschitz-continuous, that is, for every compact B R n and for all x , y B there is K > 0 such that L ( t , x , u ) L ( t , y , u ) K x y and f ( t , x , u ) f ( t , y , u ) K x y .
  • There exists also N > 0 , such that L ( t , x , u ) x N and f ( t , x , u ) x N .
  • With respect to the control u, there exists M > 0 such that
    L ( t , x , u ) M , f ( t , x , u ) M , ( t , x ) [ a , b ] × R n .

3.1. Needle-like Perturbation of the Optimal Control

We will prove a first-order necessary optimality condition of the Pontryagin type, which is only sufficient in very particular cases [28]. Any proof of a necessary optimality condition, e.g., Fermat theorem about stationary points or the classical Pontryagin maximum principle [1], begins by assuming the existence of a solution. We do the same here: we assume u * ( t ) Ω to be an optimal control to problem (1) for t [ a , b ] . The reader who is interested in the question of the existence of u * is referred to the recent paper [12]. Our aim, in this section, is to derive continuity and differentiability properties of perturbed trajectories, which are crucial to proving the necessary optimality condition for the optimal control problem (1). One way of achieving this is to perturb the optimal control by a needle-like variation and study the behavior of the corresponding state with respect to the optimal curve.
Denote by L F ( · ) the set of all Lebesgue points in [ a , b ) of the essentially bounded functions t f ( t , x ( t ) , u ( t ) ) and t L ( t , x ( t ) , u ( t ) ) . Then, for ( τ , v ) L F ( · ) × Ω , and for every θ [ 0 , b τ ) , let us consider the needle-like variation u θ L [ a , b ] , R n associated to the optimal control u * , which is given by
u θ ( t ) = u * ( t ) if t [ τ , τ + θ ) , v if t [ τ , τ + θ ) ,
for almost every t [ a , b ] .
Lemma 3
(Continuity of solutions). For any ( τ , v ) L F ( · ) × Ω , denote by x θ the corresponding state trajectory to the needle-like variation u θ , that is, the state solution of
c D a + ( α ) x θ ( t ) = f t , x θ ( t ) , u θ ( t ) , x θ ( a ) = x a .
Then, the state x θ converges uniformly to the optimal state trajectory x * whenever θ tends to zero.
Proof. 
By definition of incommensurate Caputo derivative, we have
c D a + ( α ) x θ ( t ) x * ( t ) = f ( t , x θ ( t ) , u θ ( t ) ) f t , x * ( t ) , u * ( t ) .
Then, the integral representation is obtained as
x θ ( t ) x * ( t ) = 0 t diag ( t s ) α i 1 Γ ( α i ) · f ( t , x θ ( t ) , u θ ( t ) ) f t , x * ( t ) , u * ( t ) d s .
Moreover, the function α i c α i 1 Γ ( α i ) is continuous on [ min 1 i n α i , 1 ] , where c is a non-zero constant. Thus, by the extreme value theorem due to Weierstrass, it attains a maximum. Hence, there exists α ¯ [ min 1 i n α i , 1 ] such that c α i 1 Γ ( α i ) c α ¯ 1 Γ ( α ¯ ) . This leads to
x θ ( t ) x * ( t ) 0 t ( t s ) α ¯ 1 Γ ( α ¯ ) · f ( t , x θ ( t ) , u θ ( t ) ) f t , x * ( t ) , u * ( t ) d s = I a + α ¯ f ( t , x θ ( t ) , u θ ( t ) ) f t , x * ( t ) , u * ( t ) .
Further, the following relation holds for function f,
f ( t , x θ ( t ) , u θ ( t ) ) f t , x * ( t ) , u * ( t ) = { f ( t , x θ ( t ) , u θ ( t ) ) f ( t , x * ( t ) , u θ ( t ) ) } + { f ( t , x * ( t ) , u θ ( t ) ) f ( t , x * ( t ) , u * ( t ) ) } .
Using the triangular inequality, and noticing that u θ and u * are different only on [ τ , τ + θ ) , we obtain
x θ ( t ) x * ( t ) I a + α ¯ f ( t , x θ ( t ) , u θ ( t ) ) f ( t , x * ( t ) , u θ ( t ) ) + I τ + α ¯ f ( t , x * ( t ) , u θ ( t ) ) f t , x * ( t ) , u * ( t ) .
Next, from data assumptions, K is a Lipschitz constant of f and M is an upper bound of f with respect to the control function. Thus, it follows that
x θ ( t ) x * ( t ) K I a + α ¯ x θ ( t ) x * ( t ) + M θ α ¯ Γ ( α ¯ + 1 ) .
Now, by applying the generalized Gronwall inequality of Lemma 2, it follows that
x θ ( t ) x * ( t ) M θ α ¯ Γ ( α + 1 ) 1 + a t n = 1 K n Γ ( n α ¯ ) ( t s ) n α ¯ 1 d s M θ α ¯ Γ ( α + 1 ) E α , 1 ( K ( b a ) α ) ,
where E α , 1 is the Mittag–Leffler function of parameter α ¯ [29]. Hence, we obtain that x θ x * converges to zero on [ a , b ] , which ends the proof. □
Lemma 4
(Differentiability of the perturbed trajectory). Suppose that ( x * , u * ) is an optimal pair to problem (1). Then, for ( τ , v ) L F ( · ) × Ω , the quotient variational trajectory x θ ( · ) x * ( · ) θ converges uniformly on [ τ + θ , b ] to η ( · ) when θ tends to zero, where η ( · ) is the unique solution to the incommensurate left Caputo fractional Cauchy problem
c D τ + ( α ) η ( t ) = f ( t , x * ( t ) , u * ( t ) ) x · η ( t ) , t ] τ , b ] , I τ + 1 α ¯ η ( τ ) = f ( τ , x * ( τ ) , v ) f ( τ , x * ( τ ) , u * ( τ ) ) ,
with α ¯ [ min 1 i n α i , 1 ] , such that c α i 1 Γ ( α i ) c α ¯ 1 Γ ( α ¯ ) ,   for   some   non-zero   constant   c and i = 1 , , n .
Proof. 
Set z θ ( t ) = x θ ( t ) x * ( t ) θ η ( t ) for all t [ τ + θ , b ] . Our aim is to prove that z θ converges uniformly to zero whenever θ 0 .
The integral representation of z θ is obtained as follows:
z θ ( t ) = τ τ + θ diag ( t s ) α i 1 Γ ( α i ) · f ( s , x θ ( s ) , v ) f ( s , x * ( s ) , v ) θ d s + τ τ + θ diag ( t s ) α i 1 Γ ( α i ) · f ( s , x * ( s ) , v ) f ( s , x * ( s ) , u * ( s ) ) θ d s 1 Γ ( α ¯ ) ( t τ ) α ¯ 1 f ( τ , x * ( τ ) , v ) f ( τ , x * ( τ ) , u * ( τ ) ) + τ + θ t diag ( t s ) α i 1 Γ ( α i ) · f ( s , x θ ( s ) , u * ( s ) ) f ( s , x * ( s ) , u * ( s ) ) θ f ( s , x * ( s ) , u * ( s ) ) x × x θ ( s ) x * ( s ) θ d s τ τ + θ diag ( t s ) α i 1 Γ ( α i ) · f ( s , x * ( s ) , u * ( s ) ) x × η ( s ) d s + τ + θ t diag ( t s ) α i 1 Γ ( α i ) · f ( s , x * ( s ) , u * ( s ) ) x × z θ ( s ) d s .
Note that by the existence property of α ¯ , we have that ( t s ) α i 1 Γ ( α i ) ( t s ) α ¯ 1 Γ ( α ¯ ) for all i = 1 , , n . Thus, we can deduce that
z θ ( t ) 1 Γ ( α ¯ ) τ τ + θ ( t s ) α ¯ 1 f ( s , x θ ( s ) , v ) f ( s , x * ( s ) , v ) θ d s + 1 Γ ( α ¯ ) τ τ + θ ( t s ) α ¯ 1 f ( s , x * ( s ) , v ) f ( s , x * ( s ) , u * ( s ) ) θ d s 1 Γ ( α ¯ ) ( t τ ) α ¯ 1 f ( τ , x * ( τ ) , v ) f ( τ , x * ( τ ) , u * ( τ ) ) + 1 Γ ( α ¯ ) τ + θ t ( t s ) α ¯ 1 f ( s , x θ ( s ) , u * ( s ) ) f ( s , x * ( s ) , u * ( s ) ) θ f ( s , x * ( s ) , u * ( s ) ) x × x θ ( s ) x * ( s ) θ d s 1 Γ ( α ¯ ) τ τ + θ ( t s ) α ¯ 1 f ( s , x * ( s ) , u * ( s ) ) x × η ( s ) d s + 1 Γ ( α ¯ ) τ + θ t ( t s ) α ¯ 1 f ( s , x * ( s ) , u * ( s ) ) x × z θ ( s ) d s .
In line with the work of [11], precisely the proof of Proposition 3.3 of this reference, we obtain that each term appearing in the right hand side of (5) is bounded, which yields the following estimate:
z θ ( t ) Θ 1 θ ( t ( τ + θ ) ) α ¯ 1 E α ¯ , α ¯ N ( b a ) α ¯ + Θ 2 θ ( t ( τ + θ ) ) α ¯ 1 E α ¯ , α ¯ N ( b a ) α ¯ ,
where functions Θ 1 θ and Θ 2 θ both converge uniformly to zero whenever θ tends to zero. Hence, we conclude that z θ ( t ) converges uniformly to zero as θ goes to zero, which is the desired result. □

3.2. Pontryagin’s Maximum Principle for Problem (1)

The fractional Pontryagin maximum principle has many applications, for example in Engineering, Economics and Health [30,31]. Here, we state and prove the main result of our work: a Pontryagin maximum principle for the incommensurate fractional-order optimal control problem (1).
Theorem 1
(Pontryagin Maximum Principle for (1)). If ( x * ( · ) , u * ( · ) ) is an optimal pair for (1) , then there exists λ L ( α ) , called the adjoint function variable, such that the following conditions hold for all t in the interval [ a , b ] :
  • the maximality condition
    H ( t , x * ( t ) , u * ( t ) , λ ( t ) ) = max ω Ω H t , x * ( t ) , ω , λ ( t ) ;
  • the adjoint system
    D b ( α ) λ ( t ) = H x ( t , x * ( t ) , u * ( t ) , λ ( t ) ) ;
  • the transversality condition
    I b 1 ( α ) λ ( b ) = φ x ( b , x * ( b ) ) ,
where the Hamiltonian function H is defined by
H ( t , x , u , λ ) = L ( t , x , u ) + λ · f ( t , x , u ) .
Proof. 
Let x θ ( t ) be the corresponding state trajectory to the needle-like variation u θ defined by (2). Observe that, by integration by parts, we have for λ ( · ) L ( α ) ,
a b λ ( t ) · c D a + ( α ) x θ ( t ) d t = x θ ( t ) · I b 1 ( α ) λ ( t ) a b + a b x θ ( t ) · D b ( α ) λ ( t ) d t .
This relation can be added to the objective functional at ( x θ , u θ ) defined by
J ( x θ , u θ ) = φ ( b , x θ ( b ) ) + a b L t , x θ ( t ) , u θ ( t ) d s ,
meaning that
J ( x θ , u θ ) = φ ( b , x θ ( b ) ) + a b L t , x θ ( t ) , u θ ( t ) + λ ( t ) · c D a + ( α ) x θ ( t ) x θ ( t ) · D b ( α ) λ ( t ) d t x θ ( b ) · I b 1 ( α ) λ ( b ) + x θ ( a ) · I a 1 ( α ) λ ( a ) ,
which, by substituting (3) to this latter expression, leads to
J ( x θ , u θ ) = φ ( b , x θ ( b ) ) + a b H t , x θ ( t ) , u θ ( t ) , λ ( t ) x θ ( t ) · D b ( α ) λ ( t ) d t x θ ( b ) · I b 1 ( α ) λ ( b ) + x a · I a 1 ( α ) λ ( a ) ,
where H ( t , x , u , λ ) = L ( t , x , u ) + λ · f ( t , x , u ) . Next, we write down the Taylor expansion
φ ( b , x θ ( b ) ) = φ ( b , x * ( b ) ) + x θ ( b ) x * ( b ) · φ x ( b , x * ( b ) ) + x θ x * ; H t , x θ ( t ) , u θ ( t ) , λ ( t ) = H t , x * ( t ) , u θ ( t ) , λ ( t ) + x θ ( t ) x * ( t ) · H x t , x * ( t ) , u θ ( t ) , λ ( t ) + o x θ x * .
Note that, by the continuity Lemma 3, we have the uniform convergence of x θ x * 0 whenever θ 0 . Thus, the residue term in the Taylor expansion can be expressed as a function of θ . Therefore, we can evaluate the quotient J ( x θ , u θ ) J ( x * , u * ) θ : = δ J as follows:
δ J = x θ ( b ) x * ( b ) θ · φ x ( b , x * ( b ) ) + a b H ( t , x * ( t ) , u θ ( t ) , λ ( t ) ) H t , x * ( t ) , u * ( t ) , λ ( t ) θ d t + a b H x t , x * ( t ) , u θ ( t ) , λ ( t ) D b ( α ) λ ( t ) · x θ ( t ) x * ( t ) θ d t x θ ( b ) x * ( b ) θ · I b 1 ( α ) λ ( b ) + o ( θ ) ( 1 + b a ) .
Now, by the differentiability Lemma 4, we have that x θ ( t ) x * ( t ) θ converges uniformly to η ( t ) when θ tends to zero. Therefore, the limit of δ J when θ tends to zero can be expressed as
lim θ 0 δ J = η ( b ) · φ x ( b , x * ( b ) ) + lim θ 0 a b H ( t , x * ( t ) , u θ ( t ) , λ ( t ) ) H t , x * ( t ) , u * ( t ) , λ ( t ) θ d t + a b H x t , x * ( t ) , u * ( t ) , λ ( t ) D b ( α ) λ ( t ) · η ( t ) d t η ( b ) · I b 1 ( α ) λ ( b ) .
Next, we fix
D b ( α ) λ ( t ) = H x t , x * ( t ) , u * ( t ) , λ ( t ) with I b 1 ( α ) λ ( b ) = φ x ( b , x * ( b ) ) ,
that is, the adjoint Equation (7) and the transversality condition (8). Thus, we are left with
lim θ 0 δ J = lim θ 0 a b H ( t , x * ( t ) , u θ ( t ) , λ ( t ) ) H t , x * ( t ) , u * ( t ) , λ ( t ) θ d t .
Moreover, recalling that u θ ( t ) = u * ( t ) if t [ τ , τ + θ ) ; v if t [ τ , τ + θ ) , for almost every t [ a , b ] , it follows that
lim θ 0 δ J = lim θ 0 + 1 θ τ τ + θ H ( s , x * ( s ) , v , λ ( s ) ) H ( s , x * ( s ) , u * ( s ) , λ ( s ) ) d s .
However, notice that τ is a Lebesgue point of
H ( s , x * ( s ) , v , λ ( s ) ) H ( s , x * ( s ) , u * ( s ) , λ ( s ) ) : = ψ ( s ) .
Thus, from the Lebesgue differentiation property,
1 θ τ τ + θ ψ ( s ) d s ψ ( τ ) = 1 θ τ τ + θ ψ ( s ) ψ ( τ ) d s 1 θ τ τ + θ ψ ( s ) ψ ( τ ) d s ,
and we have that the right-hand side tends to zero for almost every point τ . As a consequence,
lim θ 0 + 1 θ τ τ + θ H ( s , x * ( s ) , v , λ ( s ) ) H ( s , x * ( s ) , u * ( s ) , λ ( s ) ) d s = H ( τ , x * ( τ ) , v , λ ( τ ) ) H ( τ , x * ( τ ) , u * ( τ ) , λ ( τ ) ) .
Further, by the optimality assumption of the pair
( x * , u * ) , one has also lim θ 0 δ J 0 ,
altogether, we obtain that
H ( τ , x * ( τ ) , v , λ ( τ ) ) H ( τ , x * ( τ ) , u * ( τ ) , λ ( τ ) ) 0 .
Finally, because τ is an arbitrary Lebesgue point of the control u * and v is an arbitrary element of the set Ω , it follows that the relation
H ( t , x * ( t ) , u * ( t ) , λ ( t ) ) = max ω Ω H ( t , x * ( t ) , ω , λ ( t ) )
holds at all Lebesgue points, which ends the proof. □
Theorem 1 gives a necessary optimality condition that provides an algorithm that can be used to solve general Bolza-type fractional optimal control problems that depend on multi-order Caputo fractional derivatives: given a problem (1),
(i)
one writes the associated Hamiltonian (9);
(ii)
we use the maximality condition (6) to obtain an expression of the optimal controls in terms of the state and adjoint variables;
(iii)
we substitute the expressions obtained in step (ii) in the adjoint system (7);
(iv)
finally, we solve the system obtained in step (iii) together with the initial conditions x ( a ) = x a and the transversality condition (8).
A simple example of the usefulness of our result is given in Section 4.

4. An Illustrative Example

Let us study the following optimal control multi-order fractional Caputo problem:
x 2 ( 5 ) + 1 5 1 + exp ( 2 u ( t ) ) d t max , x = ( x 1 , x 2 ) A C ( 1 3 , 1 2 ) [ 1 , 5 ] , R 2 , u L [ 1 , 5 ] , R , c D 1 + 1 3 x 1 ( t ) = 1 exp ( 2 u ( t ) ) , c D 1 + 1 2 x 2 ( t ) = x 1 ( t ) , x ( 1 ) = ( 1 , 1 ) , u ( t ) [ 2 , 7 ] .
The Hamiltonian function and the running cost of this problem (13) are given, respectively, by
H ( t , x , u , λ ) = 1 + exp ( 2 u ) + λ 1 ( 1 exp ( 2 u ) ) + λ 2 x 1 , and φ ( b , x ( b ) ) = x 2 ( 5 ) .
We do not know if the problem has a solution or not but, if the problem has a solution, then it must satisfy the necessary optimality condition is given by our Theorem 1. Precisely, if ( x * , u * ) is an optimal pair solution to problem (13), then by application of our Theorem 1, there exists an adjoint function λ L ( 1 3 , 1 2 ) , satisfying
D 5 1 3 λ 1 ( t ) = H x 1 = λ 2 , I 5 1 1 3 = φ x 1 = 0 ; D 5 1 2 λ 2 ( t ) = H x 2 = 0 , I 5 1 1 2 = φ x 1 = 1 .
We integrate, to obtain that
λ 1 ( t ) ( 5 t ) 2 3 1 Γ ( 2 3 ) , and λ 2 ( t ) = ( 5 t ) 1 2 1 Γ ( 1 2 ) .
Moreover, from the maximality condition (6), it yields that
u * ( t ) arg v [ 2 , 7 ] 1 exp ( 2 v ( t ) ) 1 + exp ( 2 v ( t ) ) · λ 1 1 ,
for almost t [ 1 , 5 ] . Therefore, using the classical Cauchy–Schwartz inequality, we obtain
1 exp ( 2 u * ( t ) ) 1 + exp ( 2 u * ( t ) ) = 1 | ( λ 1 ( t ) , 1 ) | R 2 λ 1 1 .
This leads to
tanh u * ( t ) = 1 λ 1 ( t ) = Γ 2 3 ( 5 t ) 1 3 .
Finally, we obtain that u * is given by
u * ( t ) = arctan h Γ 2 3 ( 5 t ) 1 3 , for almost t [ 1 , 5 ] .
Note that (14) is just a Pontryagin extremal (candidate).
For practical applications, the problems are difficult and one needs to use numerical methods. Many software packages for computing fractional-order derivatives and solving fractional differential systems are now available. We refer the interested reader to [32] and references therein.

5. Conclusions

In this paper, we have studied a general Bolza-type fractional optimal control problem that depends on multi-order Caputo fractional derivatives. We have established a Pontryagin maximum principle for this incommensurate fractional-order problem. Our approach starts with a sensitivity analysis from which we prove the continuity and differentiability of perturbed trajectories to the optimal state. An illustrative example shows the applicability of our main result (Theorem 1), which provides a Pontryagin maximum principle for incommensurate fractional-order optimal control problems.
Recent applications of fractional mathematical modeling have shown the importance of considering incommensurate orders in infectious disease dynamics [33]. This might permit greater flexibility in capturing the heterogeneous nature of disease dynamics, accounting for factors such as population demographics or social behaviors of individuals [34]. Further, it has an advantage when introducing a control function in such models, by the fact that the more accurate the model, the better the control. Therefore, for future work, it would be interesting to emphasize numerical methods for incommensurate-order problems in order to handle applications on existing models of multi-order derivatives.

Author Contributions

Conceptualization, F.N. and D.F.M.T.; methodology, F.N. and D.F.M.T.; validation, F.N. and D.F.M.T.; formal analysis, F.N. and D.F.M.T.; investigation, F.N. and D.F.M.T.; writing—original draft preparation, F.N. and D.F.M.T.; writing—review and editing, F.N. and D.F.M.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundação para a Ciência e a Tecnologia, I.P. (FCT, Funder ID = 50110000187) under Grants UIDB/04106/2020 and UIDP/04106/2020.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

The authors are very grateful to three Reviewers for several comments, questions, and suggestions, that helped them to improve the submitted paper.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Pontryagin, L.S.; Boltyanskii, V.G.; Gamkrelidze, R.V.; Mishchenko, E.F. The Mathematical Theory of Optimal Processes; Interscience Publishers John Wiley & Sons, Inc.: New York, NY, USA; London, UK, 1962. [Google Scholar]
  2. Arutyunov, A.V.; Aseev, S.M. State constraints in optimal control. The degeneracy phenomenon. Syst. Control Lett. 1995, 26, 267–273. [Google Scholar] [CrossRef]
  3. Aseev, S.M.; Kryazhimskiy, A.V. The Pontryagin maximum principle and transversality conditions for a class of optimal control problems with infinite time horizons. SIAM J. Control Optim. 2004, 43, 1094–1119. [Google Scholar] [CrossRef]
  4. Moon, J. A Pontryagin maximum principle for terminal state-constrained optimal control problems of Volterra integral equations with singular kernels. AIMS Math. 2023, 8, 22924–22943. [Google Scholar] [CrossRef]
  5. Butkovskii, A.G.; Postnov, S.S.; Postnova, E.A. Fractional integro-differential calculus and its control-theoretical applications. I. Mathematical fundamentals and the problem of interpretation. Autom. Remote Control 2013, 74, 543–574. [Google Scholar] [CrossRef]
  6. Butkovskii, A.G.; Postnov, S.S.; Postnova, E.A. Fractional integro-differential calculus and its control-theoretical applications. II. Fractional dynamic systems: Modeling and hardware implementation. Autom. Remote Control 2013, 74, 725–749. [Google Scholar] [CrossRef]
  7. Agrawal, O.P. A general formulation and solution scheme for fractional optimal control problems. Nonlinear Dyn. 2004, 38, 323–337. [Google Scholar] [CrossRef]
  8. Jelicic, Z.D.; Petrovacki, N. Optimality conditions and a solution scheme for fractional optimal control problems. Struct. Multidiscip. Optim. 2009, 38, 571–581. [Google Scholar] [CrossRef]
  9. Kamocki, R. Pontryagin maximum principle for fractional ordinary optimal control problems. Math. Methods Appl. Sci. 2014, 37, 1668–1686. [Google Scholar] [CrossRef]
  10. Ali, H.M.; Pereira, F.L.; Gama, S.M.A. A new approach to the Pontryagin maximum principle for nonlinear fractional optimal control problems. Math. Methods Appl. Sci. 2016, 39, 3640–3649. [Google Scholar] [CrossRef]
  11. Bergounioux, M.; Bourdin, L. Pontryagin maximum principle for general Caputo fractional optimal control problems with Bolza cost and terminal constraints. ESAIM Control Optim. Calc. Var. 2019, 26, 35. [Google Scholar] [CrossRef]
  12. Kamocki, R. Existence of optimal control for multi-order fractional optimal control problems. Arch. Control Sci. 2022, 32, 279–303. [Google Scholar]
  13. Bardi, M.; Dolcetta, I.C. Optimal Control and Viscosity Solutions of Hamilton-Jacobi-Bellman Equations; Springer: New York, NY, USA, 1997. [Google Scholar]
  14. Czyronis, P.M. Dynamic programming problem for fractional discrete-time dynamic systems. Quadratic index of performance case. Circuits Syst. Signal Process. 2014, 7, 2131–2149. [Google Scholar] [CrossRef]
  15. Dzielinski, A.; Czyronis, P.M. Dynamic Programming for Fractional Discrete-Time Systems. IFAC Proc. Vol. 2014, 47, 2003–2009. [Google Scholar] [CrossRef]
  16. Razminia, A.; Asadizadehshiraz, M.; Torres, D.F.M. Fractional order version of the Hamilton-Jacobi-Bellman equation. ASME J. Comput. Nonlinear Dyn. 2018, 1, 14. [Google Scholar] [CrossRef]
  17. Gomoyunov, M.I. Dynamic programming principle and Hamilton-Jacobi-Bellman equations for fractional-order systems. SIAM J. Control Optim. 2020, 58, 3185–3211. [Google Scholar] [CrossRef]
  18. Teodero, G.S.; Tenreiro Machado, J.A.; de Oliveira, E.C. A review of definitions of fractional derivatives and other operators. J. Comput. Phys. 2019, 388, 195–208. [Google Scholar] [CrossRef]
  19. Baleanu, D.; Fernandez, A. On fractional operators and their classifications. Mathematics 2019, 7, 830. [Google Scholar] [CrossRef]
  20. Ndaïrou, F.; Torres, D.F.M. Optimal control problems involving combined fractional operators with general analytic kernels. Mathematics 2021, 9, 2355. [Google Scholar] [CrossRef]
  21. Kumar, S.; Upadhyay, A. Optimal control problem for fractional stochastic delayed systems with noninstantaneous impulses. IMA J. Math. Control Inform. 2021, 38, 855–880. [Google Scholar] [CrossRef]
  22. Bandaliyev, R.A.; Mamedov, I.G.; Abdullayeva, A.B.; Safarova, K.H. Optimal control problem for a degenerate fractional differential equation. Lobachevskii J. Math. 2021, 42, 1239–1247. [Google Scholar] [CrossRef]
  23. Ndaïrou, F.; Torres, D.F.M. Pontryagin maximum principle for distributed-order fractional systems. Mathematics 2021, 9, 1883. [Google Scholar] [CrossRef]
  24. Baleanu, D.; Tenreiro Machado, J.A.; Luo, A.C.J. Fractional Dynamics and Control; Springer: New York, NY, USA, 2012. [Google Scholar]
  25. Samko, S.G.; Kilbas, A.A.; Marichev, O.I. Fractional Integrals and Derivatives; Gordon and Breach Science Publishers: Yverdon, Switzerland, 1993. [Google Scholar]
  26. Zhou, Y. Fractional Evolution Equations and Inclusions: Analysis and Control; Academic Press: London, UK, 2016. [Google Scholar]
  27. Ye, H.; Gao, J.; Ding, Y. A generalized Gronwall inequality and its application to a fractional differential equation. J. Math. Anal. Appl. 2007, 328, 1075–1081. [Google Scholar] [CrossRef]
  28. Nikol’skii, M.S. On the sufficiency of the Pontryagin maximum principle in some optimization problems. Moscow Univ. Comput. Math. Cybernet. 2005, 2005, 29–37. [Google Scholar]
  29. Baghani, H.; Nieto, J.J. Some new properties of the Mittag-Leffler functions and their applications to solvability and stability of a class of fractional Langevin differential equations. Qual. Theory Dyn. Syst. 2024, 23, 18. [Google Scholar] [CrossRef]
  30. Lu, Q.; Zhu, Y.; Li, B. Necessary optimality conditions of fractional-order discrete uncertain optimal control problems. Eur. J. Control 2023, 69, 100723. [Google Scholar] [CrossRef]
  31. El Ouissari, A.; El Moutaouakil, K. Genetic algorithm applied to fractional optimal control of a diabetic patient. Ufa Math. J. 2023, 15, 129–147. [Google Scholar]
  32. Garrappa, R. Numerical Solution of Fractional Differential Equations: A Survey and a Software Tutorial. Mathematics 2018, 6, 16. [Google Scholar] [CrossRef]
  33. Ndaïrou, F.; Khalighi, K.; Lahti, L. Ebola epidemic model with dynamic population and memory. Chaos Solitons Fractals 2023, 170, 113361. [Google Scholar] [CrossRef]
  34. Jahanshahi, H.; Munoz-Pacheco, J.M.; Bekiros, S.; Alotaibi, N.D. A fractional-order SIRD model with time-dependent memory indexes for encompassing the multi-fractional characteristics of the COVID-19. Chaos Solitons Fractals 2021, 143, 110632. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ndaïrou, F.; Torres, D.F.M. Pontryagin Maximum Principle for Incommensurate Fractional-Orders Optimal Control Problems. Mathematics 2023, 11, 4218. https://doi.org/10.3390/math11194218

AMA Style

Ndaïrou F, Torres DFM. Pontryagin Maximum Principle for Incommensurate Fractional-Orders Optimal Control Problems. Mathematics. 2023; 11(19):4218. https://doi.org/10.3390/math11194218

Chicago/Turabian Style

Ndaïrou, Faïçal, and Delfim F. M. Torres. 2023. "Pontryagin Maximum Principle for Incommensurate Fractional-Orders Optimal Control Problems" Mathematics 11, no. 19: 4218. https://doi.org/10.3390/math11194218

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop