Home Optimization of Lagrange problem with higher-order differential inclusion and special boundary-value conditions
Article Open Access

Optimization of Lagrange problem with higher-order differential inclusion and special boundary-value conditions

  • Uğur Yıldırım , Dilara I. Mastaliyeva and Elimhan N. Mahmudov EMAIL logo
Published/Copyright: June 27, 2025
Become an author with De Gruyter Brill

Abstract

This article concerns about optimality conditions for boundary-value problems related to differential inclusions (DFIs) of higher orders. We intend to attain optimality conditions when a general Lagrange functional takes place in the cost function. Moreover, it is intended that these conditions are applicable to the non-convex case as well. The notion of locally adjoint mapping for both convex and non-convex functions is used via Hamiltonian functions and arg-max sets of set-valued functions to obtain results. The presented main problem turns into a problem in the calculus of variations with some simplifications. It is noteworthy to see that the famous Euler-Poisson equation arises in this case. Furthermore, a higher-order semilinear optimal control problem is considered as an application, and its sufficient conditions, including Weierstrass-Pontryagin maximum principle, are derived. Then, the dual problems for the presented primal problems are established and their duality theorems are proved. Finally, the third-order polyhedral DFI with duality relations is considered.

MSC 2010: 34A60; 49K21; 49N15; 90C46

1 Introduction

Many extremal problems, including classical optimal control problems, differential games, theory of economic dynamics, and macroeconomic problems, are commonly expressed using multi-valued mappings or variational inclusions, and they constitute an essential component of the mathematical theory of advanced optimization [112]. Several papers [1323] and monographs [4,24,25] have formulated optimal conditions, which are both necessary or sufficient for various optimal control problems involving discrete and continuous processes. The study by Cernea [26] examined a boundary-value problem involving a third-order nonconvex differential inclusion (DFI). It presents several existence conditions achieved through the utilization of the set-valued contraction principle. Andres [27] examined the existence of solutions for a nontraditional target problem involving set-valued flows produced by a vector field represented as a Marchaud map. The proof relies on the generalized Lefschetz trace formula that could be simplified to calculate the sum of local indices under some assumptions on a constraint. In addition to these works, duality, which is a robust and widely utilized technique in convex optimization, plays a crucial role in demonstrating the relationships between the primal problem and its dual counterpart for a variety of reasons. The convexity properties of the primal problem are particularly crucial to this context [5,21,2830] and related references. In the study by Rockafellar [29], an insightful framework is presented for analyzing optimality conditions for mathematical programming problems through dual programming. The fundamental importance of notions such as the Lagrange function, saddle point, and saddle value is emphasized, with generic cases derived from optimal control, approximation theory, nonlinear optimization, stochastic optimization, and variational calculus. Mordukhovich et al. [5] presented a geometric methodology for variational analysis using Fenchel conjugates as a whole, for convex objects examined within locally convex topological spaces and Banach spaces.

Several scholars have taken into consideration some significant qualitative problems with DFIs such as the existence of optimality [3139]. In the study by Abbas and Benchohra [39], Ulam’s type stability is studied for the Darboux problem of partial fractional DFIs when the right-hand side of inclusion is nonconvex-valued. Their outcomes depend on the utilization of the Covitz-Nadler fixed point theorem and a fractional variant of Gronwall’s inequality. Furthermore, substantial qualitative research is currently going forward on the subject of higher-order differential inclusions (HODIs) [4044]. In the study by Bartuzel and Fryszkowski [43], a suggested version of the Filippov lemma is proposed for third-order DFIs that are associated with constraints on boundary conditions. The existence of solutions of kth-order DFIs is studied under various boundary conditions in the study by Frigon and Kaczynski [44].

Boundary-value problems have significant implications in fields ranging from physics and engineering to economics and biology, making them a vital area of research in applied mathematics. Azzam et al. [45] examined the issue of three boundary conditions for second-order DIs. In the work by Benchohra et al. [46], existence problems for DFIs with boundary conditions are studied. The study by Mahmudov [11] is about to formulating optimality conditions for Darboux-type DFI with constraints on the boundary by utilizing adjoint partial DFIs. Matzakos and Papageorgiou [47] investigated a periodic evolution equation governed by a time-dependent subdifferential and featuring a multivalued forcing term. By employing a fixed-point theorem for pseudo-acyclic multifunctions, the existence of periodic trajectories is established. This methodology entails an examination of the solution set’s structure for the Cauchy problem addressed in the work.

In addition to the Lagrangian and Fenchel dualities, the notion of infimal convolution of convex functions has been effectively utilized in some papers [5,21,29]. Mahmudov and Mardanov [21] explored the optimality conditions for the k th-order DFIs and their corresponding duality approach. As far as we know, there have been such works [5,29] dedicated to the dual problems of the first-order DFIs. Some duality results by employing the dual operations of addition and infimal convolution of convex functions are obtained. It is worth noting that the challenges encountered in this approach stem from constructing duality framework for both discrete and discrete approximate problems.

In the current work, we discuss one of the most challenging and captivating optimization problems with higher-order DFIs and special boundary-value constraints. Indeed, the challenge in higher-order DFIs lies more in formulating the optimality conditions, which are, in general, Euler-Lagrange-type higher-order adjoint inclusions and transversality conditions. We use the so-called Mahmudov’s adjoint inclusions to obtain sufficient conditions for the k th-order DFIs instead of Euler-Langrange inclusions, which are used to obtain sufficient conditions for the first-order DFIs. It is evident that when k = 1 , Mahmudov’s and Euler-Lagrange adjoint inclusions coincide. Among the published works, no research has been exclusively dedicated to sufficient conditions for the k th-order DFIs. Therefore, we aim to address this issue, which constitutes a novelty of our specific boundary-value problem with higher-order DFIs. The presented problem is convex; however, without loss of generality, we can extend our results to the non-convex case. Moreover, this article is devoted to the duality of k th-order ordinary DFIs with special BVPs. Then, we applied our duality results to linear and polyhedral optimal control problems. The results for constructed dual problems are also a novelty to optimization literature. The structure of the work is outlined in the following order.

In Section 2, we summarize the basic definitions, concepts, and fundamental outcomes presented in Mahmudov’s book [4]. In particular, the Hamiltonian function and the sets of argmaximum of a multi-valued mapping together with LAM are introduced. Our main problem (PHC) formulated by higher-order DFIs and special boundary-value constraints is defined. Moreover, discrete-approximate problem related to our problem (PHC) is expressed.

In Section 3, we examine the optimization problem (PHC) for an arbitrary kth-order DFI with special boundary-value constraints. The sufficient conditions for BVPs with kth-order DFIs are obtained using convex and nonsmooth analysis in the form of general Mahmudov’s inclusion and with its corresponding transversality conditions. The higher-order adjoint inclusions for a closed multi-valued mapping F could be reformulated suitably with regard to the Hamiltonian function. The relation of LAMs with the subdifferential of Hamiltonian function and arg-maximum sets is essential in this simplification. Note that the main difficulty in obtaining sufficient conditions for optimality in our optimization problem is closely associated with the generalization of the Lagrange functional in cost function and DFI and the existence of special boundary constraints. We have not conducted an analysis of the discrete-approximate problem involving higher-order difference operators in this study.

In Section 4, we apply the results from Section 3 to variational problem and find out the Euler-Poisson equation, which is well known in the calculus of variations. It should be noted that the Euler-Poisson equation is obtained using the adjoint Mahmudov inclusion. Then, the higher-order “linear” optimal control problem is considered as a second application because of its applicability to control problems in engineering and science. Finally, the higher-order Euler-Lagrange-type adjoint differential equation, Weierstrass-Pontryagin maximum principle, and transversality conditions are proved for the arbitrary-order semi-linear problem (PHL).

In Section 5, we construct the dual problem (PHC*) and give an example from linear optimal control theory. The dual problem is generated without computing discrete and discrete-approximation problems. It is proved that there is no gap between the optimal values of the primal convex (PHC) and dual concave (PHC*) problems. Then, the dual problem (PHL*) by utilizing our dual theorem is formulated for the primal problem (PHL).

In Section 6, sufficient conditions and duality theorem are calculated for third-order polyhedral DFIs with BVP.

2 Preliminaries and problem statements

We summarize the essential ideas, definitions, and notions from Mahmudov’s monograph [4]. Let us assume that R n is an n -dimensional Euclidean space, ( x , v ) is a pair of elements x , v R n , and x , v is an inner product of x and v . Suppose F : R k n R n is a multi-valued (set-valued) mapping from R k n to P ( R n ) . Then, F is said to be convex if the graph of F , which is gph F = { ( x , v 1 , , v k 1 , v k ) : v k F ( x , v 1 , , v k 1 ) } is a convex subset in R ( k + 1 ) n . When graph of a multi-valued function F is a closed set in R ( k + 1 ) n , the mapping F can be said to be closed. When F ( x , v 1 , , v k 1 ) is a convex set for each ( x , v 1 , , v k 1 ) dom F = { ( x , v 1 , , v k 1 ) : F ( x , v 1 , , v k 1 ) } , F can be said to be convex-valued. The Hamiltonian function and argmaximum set for a multi-valued mapping F are described by the following relations:

H F ( x , v 1 , , v k 1 , v k * ) = sup v k { v k , v k * : v k F ( x , v 1 , , v k 1 ) } , v k * R n , F A ( x , v 1 , , v k 1 ; v k * ) = { v k F ( x , v 1 , , v k 1 ) : v k , v k * = H F ( x , v 1 , , v k 1 , v k * ) } ,

respectively. When F is convex multi-valued mapping, we place H F ( x , v 1 , , v k 1 , v k * ) = if F ( x , v 1 , , v k 1 ) = . Assume that int A is the interior of the set A R ( k + 1 ) n and ri A denotes the relative interior of the set A , which is defined as the set of interior points of A in terms of its affine hull Aff A .

A convex cone K A ( x , v 1 , , v k 1 ) is a cone of tangent directions at a point ( x , v 1 , , v k 1 ) A R ( k + 1 ) n if from ( x ¯ , v ¯ 1 , , v ¯ k 1 ) K A ( x , v 1 , , v k 1 ) , it follows that ( x ¯ , v ¯ 1 , , v ¯ k 1 ) is a tangent vector at point ( x , v 1 , , v k 1 ) A . Moreover, let us denote K A * ( ) as the dual cone to cone K A ( ) .

A function g = g ( x , v 1 , , v k 1 ) is said to be a proper function when it does not take the value of and is not identical to + . That g is proper function is necessary and sufficient for that dom g and g ( x , v 1 , , v k 1 ) is finite for ( x , v 1 , , v k 1 ) dom g = { ( x , v 1 , , v k 1 ) : g ( x , v 1 , , v k 1 ) < + } .

A function g = g ( x , v 1 , , v k 1 ) is referred to as a closed function when the epigraph of g ( ) , which is epi g = { ( η , x , v 1 , , v k 1 ) : η g ( x , v 1 , , v k 1 ) } , is a closed set.

The function g * ( x * , v 1 * , , v k 1 * ) described as follows is referred to the conjugate of g ( ) :

g * ( x * , v 1 * , , v k 1 * ) = sup x , v 1 , , v k 1 { x , x * + v 1 , v 1 * + + v k 1 , v k 1 * g ( x , v 1 , , v k 1 ) } .

Support function for the set A is denoted as W A ( x * ) and defined as follows

W A ( x * ) = sup x { x , x * : x A } .

It is well known that conjugate functions are always convex and closed regardless of the original function. Let us indicate

M F ( x * , v 1 * , , v k 1 * , v * ) = inf { x , x * + v 1 , v 1 * + + v k 1 , v k 1 * v , v * : ( x , v 1 , , v k 1 , v ) gph F } , ( x , v 1 , , v k 1 ) R k n .

We can conclude easily the relation between M F and H F as follows:

M F ( x * , v 1 * , , v k 1 * , v * ) x , x * + v 1 , v 1 * + + v k 1 , v k 1 * H F ( x , v 1 , , v k 1 , v * ) .

It is straightforward to observe that the function has the following property:

M F ( x * , v 1 * , , v k 1 * , v * ) = inf x , v 1 , , v k 1 { x , x * + v 1 , v 1 * + + v k 1 , v k 1 * H F ( x , v 1 , , v k 1 , v * ) }

represents a support function with a negative sign. Moreover, this implies for a fixed v * :

M F ( x * , v 1 * , , v k 1 * , v * ) = [ H F ( , v * ) ] * ( x * , v 1 * , , v k 1 * ) .

In other words, M F is the conjugate function of H F ( , v * ) with a negative sign.

The multi-valued mapping F * ( , x , v 1 , , v k 1 , v k ) : R n R k n can be characterized as follows when F is a convex multi-valued mapping:

F * ( y k * ; ( x 0 , v 1 0 , , v k 1 0 , v k 0 ) ) { ( x * , v 1 * , , v k 1 * ) : ( x * , v 1 * , , v k 1 * , y k * ) K gph F * ( x 0 , v 1 0 , , v k 1 0 , v k 0 ) }

is referred to locally adjoint mapping (LAM) of the multi-valued F at a point ( x 0 , v 1 0 , , v k 1 0 , v k 0 ) gph F . By Lemma 2.6 [4, p. 64], it is noteworthy to see that ( x * , v 1 * , , v k 1 * ) is an element of the LAM F * in the case that

M F ( x * , v 1 * , , v k 1 * , v * ) = x , x * + v 1 , v 1 * + + v k 1 , v k 1 * H F ( x , v 1 , , v k 1 , v * ) .

A multi-valued mapping for “nonconvex” multi-valued mapping F at a point ( x 0 , v 1 0 , , v k 1 0 , v k 0 ) gph F defined by

F * ( y k * ; ( x 0 , v 1 0 , , v k 1 0 , v k 0 ) ) ( x * , v 1 * , , v k 1 * ) : H F ( x , v 1 , , v k 1 , v k * ) H F ( x 0 , v 1 0 , , v k 1 0 , v k * ) x * , x x 0 + j = 1 k 1 v j * , v j y j 0 , ( x , v 1 , , v k 1 ) R n k , y k F A ( x , v 1 , , v k 1 ; v k * )

is labeled as the LAM. If F is convex mapping, the Hamiltonian H F ( , v k * ) is clearly concave; therefore, this description of LAM coincides with the earlier description of LAM in that case. Indeed, the closely related concept of coderivative (Frechet) generated by normal cones of multi-valued mappings with their graphs has been introduced by Mordukhovich [24]. This notion of derivative is essentially different from LAM in the case the multi-valued mapping F is not convex. Nevertheless, if the multi-valued mapping F is smooth and convex, then the coderivative notion and LAM are equivalent to each other.

In Section 3, we address the Lagrange problem of HODIs with special boundary conditions:

(1) minimize J [ x ( ) ] = 0 T g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) d t ,

(2) ( PHC ) d k x ( t ) d t k F ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) , a.e. t [ 0 , T ] ,

(3) x ( s ) ( 0 ) x ( s ) ( T ) = μ s , s = 0 , 1 , , k 1 ,

where F ( , t ) : R n k R n is a convex multi-valued mapping, g ( , t ) : R n k R 1 is a proper convex function, and μ s R n ( s = 0 , , k 1 ) are constant vectors, k is a random natural number, and T is a random positive real number. It can be seen easily that the problem (PHC) is convex since F ( , t ) is a convex multi-valued mapping and g ( , t ) is a convex proper function. In this study, we have to find an admissible trajectory arc x ˜ ( t ) of the problem (PHC) that satisfies condition (2) almost everywhere (a.e.) on a time interval [ 0 , T ] , and the boundary conditions (3) minimizing the so-called Lagrange cost functional J [ x ( ) ] . It should be mentioned when μ s = 0 , s = 0 , 1 , , k 1 , conditions (3) are periodic boundary conditions. The notion of a solution of the problem for k th-order DFIs (2) should be clarified. Assume that A C j ( [ 0 , T ] , R n ) is the space of j times differentiable functions x ( ) : [ 0 , T ] R n , where j th-order derivative x ( j ) ( ) d ( j ) x ( ) d t j ( j = 1 , , k ) is absolutely continuous. Moreover, we can suppose that L 1 ( [ 0 , T ] , R n ) is the Banach space of integrable functions u ( ) : [ 0 , T ] R n endowed with the norm u ( ) 1 = 0 T u ( t ) d t in the Lebesgue sense. A function x ( ) A C k 1 ( [ 0 , T ] , R n ) is a feasible solution of a problems (2) and (3) if there exists an integrable function v ( t ) F ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) , a.e. t [ 0 , T ] such that x ( k ) ( t ) = v ( t ) a.e. t [ 0 , T ] and x ( ) satisfies the boundary conditions (3). It should be noted that having the boundary conditions as a constraint x ( s ) ( 0 ) x ( s ) ( T ) = μ s , s = 0 , 1 , , k 1 are especially substantial for applied scientists. For example, in the book of Aubin and Cellina [1], first order DFIs and their various applications are introduced in detail. These applications frequently emerge in the field of optimal control, mathematical finance, classical mechanics, designing optimal or stabilizing feedback, as well as aerospace engineering, anti-vibration, burning in rocket motors, and biophysics.

Note that to obtain sufficient conditions for optimality of the higher-order Langrange problem with special boundary conditions, we can use the discrete-approximate method. Hence, our problem (PHC) is replaced by the subsequent k th-order discrete approximate problem (PHDA):

(4) min J h [ x ( ) ] = t = k h T h g ( x ( t ( k 1 ) h ) , Δ x ( t ( k 1 ) h ) , , Δ k 1 x ( t ( k 1 ) h ) , t ) ,

(5) ( PHDA ) Δ k x ( t ) F ( x ( t ) , Δ x ( t ) , , Δ k 1 x ( t ) , t ) , t = 0 , h , , T k h ,

(6) x ( 0 ) x ( T ) = μ 0 , Δ x ( 0 ) Δ x ( T ) = μ 1 , , Δ k 1 x ( 0 ) Δ k 1 x ( T ) = μ k 1 .

Here, the k th-order finite difference operator can be described in this fashion:

Δ k x ( t ) = 1 h k s = 0 k ( 1 ) s C k s x ( t + ( k s ) h ) , C k s = k ! s ! ( k s ) ! .

The main idea for this method is to transform the continuous problem (PHC) to discrete problem (PHDA) by the finite difference approximation of differential operators. Then, optimality conditions for discrete inclusion need to be derived and applied to discrete-approximation problem. Subsequently, taking limit as h 0 , we can construct the optimality conditions for continuous problem (PHC). It is obvious that these procedures are rather complicated due to high-order derivatives. Although the method of discrete-approximation is a brilliant technique for the investigation of optimality conditions of our problem, we omit it to avoid long and tedious calculations. Therefore, we will approach to the problem (PHC) in a different way so as to prove theorems in this article.

3 Sufficient conditions for problem (PHC)

We initiate our consideration with sufficient conditions for the addressed problem (PHC). To begin, we relate the problem (PHC) to the succeeding kth-order Euler-Lagrange-type DFI or the so-called Mahmudov’s inclusion:

  1. ( 1 ) k d k x * ( t ) d t k + d φ k 1 * ( t ) d t , φ k 1 * ( t ) + d φ k 2 * ( t ) d t , , φ 2 * ( t ) + d φ 1 * ( t ) d t , φ 1 * ( t ) F * ( x * ( t ) ; ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k ) ( t ) ) , t ) z g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) , a.e. t [ 0 , T ] , where x ˜ ( t ) is a feasible solution of the problem ( PHC ) , z = ( x , v 1 , , v k 1 ) , x * ( t ) and φ * ( t ) are the so-called adjoint variables and the transversality conditions at the points t = 0 and t = T :

  2. x * ( s ) ( 0 ) = x * ( s ) ( T ) ; s = 0 , 1 , , k 1 .

  3. φ j * ( 0 ) = φ j * ( T ) ; j = 1 , 2 , , k 1 .

Subsequently, we assume x * ( t ) , t [ 0 , T ] is absolutely continuous function with the derivatives up to k 1 th-order and d k x * ( ) d t k L 1 n ( [ 0 , T ] ) . Moreover, we suppose φ j * ( t ) , j = 1 , , k 1 , t [ 0 , T ] is absolutely continuous and d φ j * ( ) d t L 1 n ( [ 0 , T ] ) , j = 1 , , k 1 .

Additionally, concerning the argmaximum set, we will need an additional condition stipulating LAM F * ( , t ) at a specified point:

  1. d k x ˜ ( t ) d t k F A ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) ; x * ( t ) , t ) , a.e. t [ 0 , T ] .

We can conclude that the subsequent theorem is valid.

Theorem 3.1

Suppose that F ( , t ) : R k n R n is a convex multi-valued mapping and g ( , t ) : R k n R 1 { + } is a continuous convex function. For the trajectory x ˜ ( ) to be optimal in the convex problem ( P H C ) , it suffices that there exists a set of absolutely continuous functions { x * ( t ) , φ j * ( t ) ; j = 1 , , k 1 } , t [ 0 , T ] that fulfill, almost everywhere, the kth-order Mahmudov’s inclusion (1), the non-emptiness condition of LAM (4), and the transversality conditions (2), (3) at the points t = 0 and t = T .

Proof

Clearly, by Theorem 2.1 [4], the LAM is expressed as follows: F * ( x * ( t ) ; ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k ) ( t ) ) , t ) = z H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x * ( t ) ) . Here, to take subdifferential of Hamiltonian function, we should use this convention: z H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x * ( t ) ) = z [ H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x * ( t ) ) ] . Then, by taking into account the Moreau-Rockafellar theorem [4,24] for condition (1), we obtain the k th-order adjoint DFI as follows:

(7) ( 1 ) k d k x * ( t ) d t k + d φ k 1 * ( t ) d t , φ k 1 * ( t ) + d φ k 2 * ( t ) d t , , φ 2 * ( t ) + d φ 1 * ( t ) d t , φ 1 * ( t ) z [ H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k ) ( t ) , x * ( t ) ) g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) ] .

By the definition of subdifferential, we rewrite the last relation (7) in the form:

(8) H F ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , x * ( t ) ) H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x * ( t ) ) g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) + g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) ( 1 ) k x * ( k ) ( t ) + d φ k 1 * ( t ) d t , x ( t ) x ˜ ( t ) + φ k 1 * ( t ) + d φ k 2 * ( t ) d t , x ( t ) x ˜ ( t ) + φ k 2 * ( t ) + d φ k 3 * ( t ) d t , x ( t ) x ˜ ( t ) + + φ 3 * ( t ) + d φ 2 * ( t ) d t , x ( k 3 ) ( t ) x ˜ ( k 3 ) ( t ) + φ 2 * ( t ) + d φ 1 * ( t ) d t , x ( k 2 ) ( t ) x ˜ ( k 2 ) ( t ) + φ 1 * ( t ) , x ( k 1 ) ( t ) x ˜ ( k 1 ) ( t ) .

Furthermore, by condition (4) and description of the Hamiltonian function, we can deduce

(9) H F ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , x * ( t ) ) H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x * ( t ) ) x ( k ) ( t ) x ˜ ( k ) ( t ) , x * ( t ) .

Therefore, it can be noted that equation (8) turns into an important equation via equation (9):

(10) ( 1 ) k x * ( k ) ( t ) , x ( t ) x ˜ ( t ) x ( k ) ( t ) x ˜ ( k ) ( t ) , x * ( t ) + j = 1 k 1 d d t φ j * ( t ) , x ( k j 1 ) ( t ) x ˜ ( k j 1 ) ( t )

g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) .

Now, if we integrate both sides of the inequality (10) from 0 to T , we obtain

(11) 0 T [ ( 1 ) k x * ( k ) ( t ) , x ( t ) x ˜ ( t ) x ( k ) ( t ) x ˜ ( k ) ( t ) , x * ( t ) ] d t + j = 1 k 1 0 T d φ j * ( t ) , x ( k j 1 ) ( t ) x ˜ ( k j 1 ) ( t ) 0 T ( g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) ) d t .

Afterward, it is not hard to see that the integrand of the first integral in (11) is converted into the helpful equation, which is as follows:

(12) ( 1 ) k x * ( k ) ( t ) , x ( t ) x ˜ ( t ) x ( k ) ( t ) x ˜ ( k ) ( t ) , x * ( t ) = d d t ( 1 ) k x * ( k 1 ) ( t ) , x ( t ) x ˜ ( t ) + d d t ( 1 ) k 1 x * ( k 2 ) ( t ) , x ( t ) x ˜ ( t ) + d d t ( 1 ) k 2 x * ( k 3 ) ( t ) , x ( t ) x ˜ ( t ) + d d t ( 1 ) k 3 x * ( k 4 ) ( t ) , x ( t ) x ˜ ( t ) + d d t x * ( t ) , x ( k 1 ) ( t ) x ˜ ( k 1 ) ( t ) .

Let us denote

Ω = 0 T [ ( 1 ) k x * ( k ) ( t ) , x ( t ) x ˜ ( t ) x ( k ) ( t ) x ˜ ( k ) ( t ) , x * ( t ) ] d t + j = 1 k 1 0 T d φ j * ( t ) , x ( k j 1 ) ( t ) x ˜ ( k j 1 ) ( t ) .

Now if we consider (12) and use higher-order differential calculus from [48], we can compute Ω easily as follows:

Ω = 0 T [ ( 1 ) k x * ( k ) ( t ) , x ( t ) x ˜ ( t ) x ( k ) ( t ) x ˜ ( k ) ( t ) , x * ( t ) ] d t + j = 1 k 1 0 T d φ j * ( t ) , x ( k j 1 ) ( t ) x ˜ ( k j 1 ) ( t ) = x ( T ) x ˜ ( T ) , ( 1 ) k x * ( k 1 ) ( T ) + x ( T ) x ˜ ( T ) , ( 1 ) k 1 x * ( k 2 ) ( T ) + x ( T ) x ˜ ( T ) , ( 1 ) k 2 x * ( k 3 ) ( T ) + x ( k 1 ) ( T ) x ˜ ( k 1 ) ( T ) , x * ( T ) x ( 0 ) x ˜ ( 0 ) , ( 1 ) k x * ( k 1 ) ( 0 ) x ( 0 ) x ˜ ( 0 ) , ( 1 ) k 1 x * ( k 2 ) ( 0 ) x ( 0 ) x ˜ ( 0 ) , ( 1 ) k 2 x * ( k 3 ) ( 0 ) + x ( k 1 ) ( 0 ) x ˜ ( k 1 ) ( 0 ) , x * ( 0 ) + j = 1 k 1 φ j * ( T ) , x ( k j 1 ) ( T ) x ˜ ( k j 1 ) ( T ) j = 1 k 1 φ j * ( 0 ) , x ( k j 1 ) ( 0 ) x ˜ ( k j 1 ) ( 0 ) = x * ( 0 ) , x ( k 1 ) ( 0 ) x ˜ ( k 1 ) ( 0 ) j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( 0 ) + φ j * ( 0 ) , x ( k j 1 ) ( 0 ) x ˜ ( k j 1 ) ( 0 ) + j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( T ) + φ j * ( T ) , x ( k j 1 ) ( T ) x ˜ ( k j 1 ) ( T ) x * ( T ) , x ( k 1 ) ( T ) x ˜ ( k 1 ) ( T ) .

Now if we employ boundary conditions of the problem (PHC) and the feasibility of x ( ) and x ˜ ( ) as follows:

x ( s ) ( 0 ) = x ( s ) ( T ) + μ s , s = 0 , 1 , , k 1 , x ˜ ( s ) ( 0 ) = x ˜ ( s ) ( T ) + μ s , s = 0 , 1 , , k 1 .

Next, we can obtain the following:

Ω = x * ( 0 ) , x ( k 1 ) ( T ) + μ k 1 x ˜ ( k 1 ) ( T ) μ k 1 j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( 0 ) + φ j * ( 0 ) , x ( k j 1 ) ( T ) + μ k j 1 x ˜ ( k j 1 ) ( T ) μ k j 1 + j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( T ) + φ j * ( T ) , x ( k j 1 ) ( T ) x ˜ ( k j 1 ) ( T ) x * ( T ) , x ( k 1 ) ( T ) x ˜ ( k 1 ) ( T ) = x * ( 0 ) , x ( k 1 ) ( T ) x ˜ ( k 1 ) ( T ) j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( 0 ) + φ j * ( 0 ) , x ( k j 1 ) ( T ) x ˜ ( k j 1 ) ( T ) + j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( T ) + φ j * ( T ) , x ( k j 1 ) ( T ) x ˜ ( k j 1 ) ( T ) x * ( T ) , x ( k 1 ) ( T ) x ˜ ( k 1 ) ( T ) .

Hence, using transversality conditions (2), and (3) of Theorem 3.1, we arrive at the subsequent expression:

(13) Ω = x * ( T ) , x ( k 1 ) ( T ) x ˜ ( k 1 ) ( T ) j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( T ) + φ j * ( T ) , x ( k j 1 ) ( T ) x ˜ ( k j 1 ) ( T ) + j = 1 k 1 ( 1 ) j + 1 x * ( j ) ( T ) + φ j * ( T ) , x ( k j 1 ) ( T ) x ˜ ( k j 1 ) ( T ) x * ( T ) , x ( k 1 ) ( T ) x ˜ ( k 1 ) ( T ) = 0 .

Then, if we substitute Ω = 0 into (11), we find the following inequality:

(14) 0 0 T ( g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) ) d t .

Hence, we attain the wished outcome for every feasible trajectory x ( ) ,

(15) 0 T g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) d t 0 T g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) d t .

It means that J [ x ( t ) ] J [ x ˜ ( t ) ] , x ( t ) , t [ 0 , T ] , and then, x ˜ ( t ) is optimal.□

It is worth noting that for a convex and closed multi-valued mapping F , conditions (1) and (4) of Theorem 3.1 can be substituted with the subdifferentials of Hamiltonian functions. Then, we can conclude the following corollary.

Corollary 3.1

When F ( , t ) : R k n R n is a closed multi-valued mapping, conditions (1) and (4) of Theorem 3.1 can be reformulated with regard to Hamiltonian function in a substantially more useful manner:

( 1 ) k d k x * ( t ) d t k + d φ k 1 * ( t ) d t , φ k 1 * ( t ) + d φ k 2 * ( t ) d t , , φ 2 * ( t ) + d φ 1 * ( t ) d t , φ 1 * ( t ) z H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k ) ( t ) , x * ( t ) ) z g ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) . d k x ˜ ( t ) d t k z H F ( x ˜ ( t ) , x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x * ( t ) ) , a.e. t [ 0 , T ] .

Proof

According to Theorem 2.1 [4], we are able to express

F * ( v k * ; ( z , v k ) , t ) = z H F ( z , v k * ) , F A ( z ; v k * , t ) = v k * H F ( z , v k * ) .

Therefore, the statements of the corollary are identical to conditions (1) and (4) of Theorem 3.1.□

4 Some applications of higher-order continuous problem (PHC)

Let us show by applying Theorem 3.1, we can obtain the Euler-Poisson equation for the calculus of variations. We address the Cauchy problem by employing μ s = 0 , s = 0 , 1 , , k 1 and without DFI (2), which is easier than our main problem. The variational problem involving only one function and its higher-order derivatives up to k 1 is presented here:

minimize J [ x ( ) ] = 0 T g ( x ( t ) , x ( t ) , x ( t ) , x ( k 1 ) ( t ) , t ) d t , ( PV ) x ( j ) ( 0 ) = x j 0 ; x ( j ) ( T ) = x j T , j = 0 , , k 1 ,

where g is a real-valued function that has continuous first-order partial derivatives, x ( ) C k 1 ( [ 0 , T ] ) and x j 0 , x j T , j = 0 , 1 , , k 1 are constant vectors. Within the instance that is being displayed, g ( x , x , , x ( k 1 ) , t ) = g x , g x , , g x ( k 1 ) , dom F ( , t ) R k n and F ( , t ) R n . It follows that K F ( z , v k ) = R ( k + 1 ) n and K F * ( z , v k ) { 0 } R ( k + 1 ) n . Hence, F * ( v k * ; ( z , v k ) ) = { 0 } R k n , and then, v k * = 0 , which implies x * ( j ) ( t ) 0 , j = 0 , , k , t [ 0 , T ] . Thus, it can be obtained from condition (1) of Theorem 3.1 that

(16) d φ k 1 * ( t ) d t + g x = 0 , φ k 1 * ( t ) + d φ k 2 * ( t ) d t + g x = 0 , φ 2 * ( t ) + d φ 1 * ( t ) d t + g x ( k 2 ) = 0 , φ 1 * ( t ) + g x ( k 1 ) = 0 .

Subsequently, we obtain the Euler-Poisson equation for the investigated problem by applying the last relation and putting g ˜ g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) by using sequential substitution:

(17) g ˜ x d d t g ˜ x + d 2 d t 2 g ˜ x + ( 1 ) k 1 d k 1 d t k 1 g ˜ x ( k 1 ) = 0 .

Corollary 4.1

Assume that the Lagrangian g is a real-valued function that has continuous first-order partial derivatives and x ( ) C k 1 ( [ 0 , T ] ) . Then, for x ˜ ( t ) , t [ 0 , T ] to be optimal solution of the problem ( P V ) in calculus of variations, it is sufficient that the Euler-Poisson equation is satisfied:

m = 1 k ( 1 ) m 1 d m 1 d t m 1 g x ( m 1 ) = 0 .

Now, let us study k th-order “linear” DFIs with special boundary-value problem:

minimize J [ x ( ) ] = 0 T g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) d t , ( PHL ) d k x ( t ) d t k F ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) , a.e. t [ 0 , T ] , x ( s ) ( 0 ) x ( s ) ( T ) = μ s , s = 0 , 1 , , k 1 , F ( x , v 1 , , v k 1 ) A 0 x + j = 1 k 1 A j v j + B U ,

where g ( , t ) is a real-valued function that has continuous first-order partial derivatives, A j for j = 0 , , k 1 and B are n × n and n × r matrices, respectively, U is a convex compact set of R r , μ s , s = 0 , 1 , , k 1 are constant vectors. The aim is to determine the controlling parameter w ˜ ( t ) U when J [ x ( ) ] is minimized by arc x ˜ ( t ) :

H F ( x , v 1 , , v k 1 , v k * ) = sup v k { v k , v k * : v k F ( x , v 1 , , v k 1 ) }

= sup w A 0 x + j = 1 k 1 A j v j + B U , v k * : w U = x , A 0 * v k * + j = 1 k 1 v j , A j * v k * + max w { B w , v k * : w U } ,

where A j * is the adjoint matrix of A j for j = 0 , , k 1 . In this case, A j * is equal to transpose of A j .

Then, if v ˜ k = A 0 x ˜ + j = 1 k 1 A j v ˜ j + B w ˜ , w ˜ U , one has

(18) F * ( v k * ; ( x ˜ , v 1 ˜ , , v ˜ k ) ) = ( A 0 * v k * , A 1 * v k * , , A k 1 * v k * ) , if v k F A ( x , v 1 , , v k 1 ; v k * ) , if v k F A ( x , v 1 , , v k 1 ; v k * ) .

Now, if we apply (1) in Theorem 3.1 for the present problem (PHL), we can conclude

( 1 ) k d k x * ( t ) d t k + d φ k 1 * ( t ) d t , φ k 1 * ( t ) + d φ k 2 * ( t ) d t , , φ 2 * ( t ) + d φ 1 * ( t ) d t , φ 1 * ( t ) , ( A 0 * x * ( t ) , A 1 * x * ( t ) , , A k 1 * x * ( t ) ) g ˜ x , g ˜ x , , g ˜ x ( k 1 ) .

If we write more clearly, we can see

(19) ( 1 ) k d k x * ( t ) d t k + d φ k 1 * ( t ) d t = A 0 * x * ( t ) g ˜ x , φ k 1 * ( t ) + d φ k 2 * ( t ) d t = A 1 * x * ( t ) g ˜ x φ 2 * ( t ) + d φ 1 * ( t ) d t = A k 2 * x * ( t ) g ˜ x ( k 2 ) , φ 1 * ( t ) = A k 1 * x * ( t ) g ˜ x ( k 1 ) .

We can attain easily the following differential equation by sequential substitution into (19):

(20) ( 1 ) k d k x * ( t ) d t k = j = 0 k 1 ( 1 ) j A j * x * ( j ) ( t ) g ˜ x + d d t g ˜ x ( 1 ) k 1 d k 1 d t k 1 g ˜ x ( k 1 ) .

Then, if we apply transversality conditions (2) and (3) in Theorem 3.1 to equation (19), the following result can be obtained:

(21) g ˜ x ( k 1 ) t = 0 = g ˜ x ( k 1 ) t = T g ˜ x ( k 2 ) + d d t g ˜ x ( k 1 ) t = 0 = g ˜ x ( k 2 ) + d d t g ˜ x ( k 1 ) t = T g ˜ x + + ( 1 ) k d k 2 d t k 2 g ˜ x ( k 1 ) t = 0 = g ˜ x + + ( 1 ) k d k 2 d t k 2 g ˜ x ( k 1 ) t = T .

Theorem 4.1

The trajectory x ˜ ( t ) , governed by the control function w ˜ ( t ) , has been a solution of the problem (PHL) when there exists an absolutely continuous function x * ( t ) along with its ( k 1 ) th-order derivatives, which fulfills the kth-order linear adjoint differential equation (20), condition (2) of Theorem 3.1, equation (21), and the following Weierstrass-Pontryagin maximum principle:

B w ˜ ( t ) , x * ( t ) = max w { B w , x * ( t ) : w U } .

Proof

We showed (20) and (21) for Theorem 4.1 and condition (2) is the same as in Theorem 3.1. Then, the maximality condition is left to complete the proof. Let us denote F 0 ( x , v 1 , , v k 1 ) = B U and define the semilinear multi-valued mapping F = A 0 x + j = 1 k 1 A j v j + F 0 . Then, we obtain F * ( v k * ; ( x ˜ , v 1 ˜ , , v ˜ k ) ) = ( A 0 * v k * , A 1 * v k * , , A k 1 * v k * ) + F 0 * ( v k * ; ( x ˜ , v 1 ˜ A 0 x ˜ , , v ˜ k A 0 x ˜ j = 1 k 1 A j v ˜ j ) ) , where v k belongs to the argmaximum set F A ( x , v 1 , , v k 1 ; v k * ) , which means that v A 0 x j = 1 k 1 A j v j F 0 A ( x , v 1 , , v k 1 ; v k * ) . Furthermore, since H F 0 ( x , v 1 , , v k 1 , v k * ) = max w U B w , v k * , it follows that F 0 * ( v k * ; ( x ˜ , v 1 ˜ A 0 x ˜ , , v ˜ k A 0 x ˜ j = 1 k 1 A j v ˜ j ) ) = ( x , v 1 , , v k 1 ) H F 0 ( x , v 1 , , v k 1 , v k * ) = { ( 0 , , 0 ) k } . That implies F * ( v k * ; ( x ˜ , v 1 ˜ , , v ˜ k ) ) , and v ˜ k F A ( x , v 1 , , v k 1 ; v k * ) means B w ˜ F 0 A ( x , v 1 , , v k 1 ; v k * ) . Then, B w ˜ , v k * = H F 0 ( x , v 1 , , v k 1 , v k * ) = max w { B w , v k * : w U } . Hence, we obtained the Weierstrass-Pontryagin maximum principle for the problem (PHL).□

5 Duality of problem (PHC)

For the previous problem (PHC), we formulate the corresponding dual problem as (PHC*):

sup 0 T g * ( z 1 * ( t ) , z 2 * ( t ) , , z k * ( t ) , t ) d t + 0 T M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , ( PHC * ) φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) d t j = 1 k 1 μ k j 1 , ( 1 ) j + 1 x * ( j ) ( 0 ) φ j * ( 0 ) μ k 1 , x * ( 0 ) .

The supremum in the problem (PHC*) is taken over the set of functions:

k , j , i ( T , 0 ) { x * ( ) , z 1 * ( ) , , z k * ( ) , φ 1 * ( ) , , φ k 1 * ( ) : x * ( j ) ( 0 ) = x * ( j ) ( T ) , φ i * ( 0 ) = φ i * ( T ) , j = 0 , , k 1 , i = 1 , , k 1 } .

We can define similarly ˜ k , j , i ( T , 0 ) as follows:

˜ k , j , i ( T , 0 ) { x ˜ * ( ) , z ˜ 1 * ( ) , , z ˜ k * ( ) , φ ˜ 1 * ( ) , , φ ˜ k 1 * ( ) : x ˜ * ( j ) ( 0 ) = x ˜ * ( j ) ( T ) , φ ˜ i * ( 0 ) = φ ˜ i * ( T ) , j = 0 , , k 1 , i = 1 , , k 1 } .

Moreover, we suppose x * ( t ) for t [ 0 , T ] is absolutely continuous function comprising the higher-order derivatives up to k 1 , i.e., x * ( ) A C k 1 ( [ 0 , T ] , R n ) and x * ( k ) ( ) L 1 n ( [ 0 , T ] , R n ) . Furthermore, z m * ( ) , φ j * ( ) A C ( [ 0 , T ] , R n ) and z m * ( ) , φ j * ( ) L 1 n ( [ 0 , T ] , R n ) for m = 1 , , k , j = 1 , , k 1 . Now, it is time to express our duality theorem.

Theorem 5.1

Assume that g ( , t ) is a continuous proper convex function and F ( , t ) is a convex closed multi-valued mapping. Suppose that x ˜ ( ) is an optimal solution of the primal convex problem (PHC). Whence, a family of functions ˜ k , j , i ( T , 0 ) : ( z ˜ 1 * ( t ) , , z ˜ k * ( t ) ) g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) is considered as optimal solution of the dual problem (PHC*) if and only if conditions (1)–(4) of Theorem 3.1 are fulfilled. Furthermore, the primal (PHC) and dual (PHC*) problems have the same optimal values.

Proof

Initially, we must establish weak duality for all feasible solutions between the problem (PHC) and the dual problem (PHC*). It means that the optimal value of the primal problem is greater than or equal to the optimal value of the dual problem. Namely, we need to show the following inequality:

(22) 0 T g ( x ( t ) , x ( t ) , , x ( k 1 ) ( t ) , t ) d t 0 T g * ( z 1 * ( t ) , z 2 * ( t ) , , z k * ( t ) , t ) d t + 0 T M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) d t j = 1 k 1 μ k j 1 , ( 1 ) j + 1 x * ( j ) ( 0 ) φ j * ( 0 ) μ k 1 , x * ( 0 ) .

If we apply the Young’s inequality [4] for g * ( ) , it will be concluded that

(23) g * ( z 1 * ( t ) , , z k * ( t ) , t ) g ( x ( t ) , , x ( k 1 ) ( t ) , t ) x ( t ) , z 1 * ( t ) x ( t ) , z 2 * ( t ) x ( k 1 ) ( t ) , z k * ( t ) .

When we integrate both sides of inequality (23) over the interval [ 0 , T ] , it can be obtained that

(24) 0 T g * ( z 1 * ( t ) , , z k * ( t ) , t ) d t 0 T g ( x ( t ) , , x ( k 1 ) ( t ) , t ) d t 0 T j = 0 k 1 x ( j ) ( t ) , z j + 1 * ( t ) d t .

Now, if we use the meanings of M F and Hamiltonian functions, it can be written that

(25) M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) x ( t ) , ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + x ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + + x ( k 1 ) ( t ) , φ 1 * ( t ) x ( k ) ( t ) , x * ( t ) + x ( t ) , z 1 * ( t ) + x ( t ) , z 2 * ( t ) + + x ( k 1 ) ( t ) , z k * ( t ) = d d t x ( t ) , φ k 1 * ( t ) + d d t x ( t ) , φ k 2 * ( t ) + + d d t x ( k 2 ) ( t ) , φ 1 * ( t ) + j = 0 k 1 x ( j ) ( t ) , z j + 1 * ( t ) + x ( t ) , ( 1 ) k x * ( k ) ( t ) x ( k ) ( t ) , x * ( t ) .

Therefore, integrating both sides over the interval [ 0 , T ] yields that

(26) 0 T M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) d t 0 T d x ( t ) , φ k 1 * ( t ) + 0 T d x ( t ) , φ k 2 * ( t ) + + 0 T d x ( k 2 ) ( t ) , φ 1 * ( t ) + 0 T [ x ( t ) , ( 1 ) k x * ( k ) ( t ) x ( k ) ( t ) , x * ( t ) ] d t + 0 T j = 0 k 1 x ( j ) ( t ) , z j + 1 * ( t ) d t = x ( T ) , φ k 1 * ( T ) + x ( T ) , φ k 2 * ( T ) + + x ( k 2 ) ( T ) , φ 1 * ( T ) x ( 0 ) , φ k 1 * ( 0 ) x ( 0 ) , φ k 2 * ( 0 ) x ( k 2 ) ( 0 ) , φ 1 * ( 0 ) + 0 T [ x ( t ) , ( 1 ) k x * ( k ) ( t ) x ( k ) ( t ) , x * ( t ) ] d t + 0 T j = 0 k 1 x ( j ) ( t ) , z j + 1 * ( t ) d t .

Then, we will calculate the integral of Λ ( t ) x ( t ) , ( 1 ) k x * ( k ) ( t ) x ( k ) ( t ) , x * ( t ) independently. It can be easily seen that Λ can be transformed as

(27) Λ ( t ) = d d t x ( t ) , ( 1 ) k x * ( k 1 ) ( t ) + d d t x ( t ) , ( 1 ) k 1 x * ( k 2 ) ( t ) + d d t x ( t ) , ( 1 ) k 2 x * ( k 3 ) ( t ) + + d d t x ( k 2 ) ( t ) , x * ( t ) d d t x ( k 1 ) ( t ) , x * ( t ) .

Now, by employing equation (27), if we calculate the integral of Λ ( t ) over the interval [ 0 , T ] , we obtain

(28) 0 T Λ ( t ) d t = x ( T ) , ( 1 ) k x * ( k 1 ) ( T ) + x ( T ) , ( 1 ) k 1 x * ( k 2 ) ( T ) + x ( T ) , ( 1 ) k 2 x * ( x 3 ) ( T ) + + x ( k 2 ) ( T ) , x * ( T ) x ( k 1 ) ( T ) , x * ( T ) x ( 0 ) , ( 1 ) k x * ( k 1 ) ( 0 ) x ( 0 ) , ( 1 ) k 1 x * ( k 2 ) ( 0 ) x ( 0 ) , ( 1 ) k 2 x * ( x 3 ) ( 0 ) x ( k 2 ) ( 0 ) , x * ( 0 ) + x ( k 1 ) ( 0 ) , x * ( 0 ) = j = 1 k 1 μ k j 1 , ( 1 ) j x * ( j ) ( 0 ) + μ k 1 , x * ( 0 ) .

Then, if we plug equation (28) into equation (26), the subsequent inequality holds

(29) 0 T M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) d t j = 1 k 1 μ k j 1 , ( 1 ) j x * ( j ) ( 0 ) + μ k 1 , x * ( 0 ) + 0 T j = 0 k 1 x ( j ) ( t ) , z j + 1 * ( t ) d t j = 1 k 1 μ k j 1 , φ j * ( 0 ) .

Now, if we add inequalities (24) and (29), it can be attained that

(30) 0 T g * ( z 1 * ( t ) , , z k * ( t ) , t ) d t + 0 T M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) d t 0 T g ( x ( t ) , , x ( k 1 ) ( t ) , t ) d t + j = 1 k 1 μ k j 1 , ( 1 ) j x * ( j ) ( 0 ) + μ k 1 , x * ( 0 ) j = 1 k 1 μ k j 1 , φ j * ( 0 ) .

Here, if we leave the integral of g ( ) alone in the right-hand side of inequality, then it yields that

(31) 0 T g * ( z 1 * ( t ) , , z k * ( t ) , t ) d t + 0 T M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) d t j = 1 k 1 μ k j 1 , ( 1 ) j + 1 x * ( j ) ( 0 ) μ k 1 , x * ( 0 ) j = 1 k 1 μ k j 1 , φ j * ( 0 ) 0 T g ( x ( t ) , , x ( k 1 ) ( t ) , t ) d t .

Finally, we have the following desired result:

(32) 0 T g * ( z 1 * ( t ) , , z k * ( t ) , t ) d t + 0 T M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) d t j = 1 k 1 μ k j 1 , ( 1 ) j + 1 x * ( j ) ( 0 ) φ j * ( 0 ) μ k 1 , x * ( 0 ) 0 T g ( x ( t ) , , x ( k 1 ) ( t ) , t ) d t .

Therefore, inequality (22) holds. Now, assume a family of dual functions { ˜ k , j , i ( T , 0 ) : ( z ˜ 1 * ( t ) , , z ˜ k * ( t ) ) g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) } fulfills conditions (1)–(4) of Theorem 3.1. Therefore, conditions (1) and (4) imply that

(33) H F ( x ( t ) , , x ( k 1 ) ( t ) , x ˜ * ( t ) ) H F ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x ˜ * ( t ) ) ( 1 ) x ˜ * ( k ) ( t ) + φ ˜ k 1 * ( t ) , x ( t ) x ˜ ( t ) + φ ˜ k 1 * ( t ) + φ ˜ k 2 * ( t ) , x ( t ) x ˜ ( t ) + + φ ˜ 2 * ( t ) + φ ˜ 1 * ( t ) , x ( k 2 ) ( t ) x ˜ ( k 2 ) ( t ) + φ ˜ 1 * ( t ) , x ( k 1 ) ( t ) x ˜ ( k 1 ) ( t ) + g ( x ( t ) , , x ( k 1 ) ( t ) , t ) g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) .

From the meaning of function M F , we can express the following equality:

(34) ( 1 ) k x ˜ * ( k ) ( t ) + φ ˜ k 1 * ( t ) + z ˜ 1 * ( t ) , x ˜ ( t ) + φ ˜ k 1 * ( t ) + φ ˜ k 2 * ( t ) + z ˜ 2 * ( t ) , x ˜ ( t ) + + φ ˜ 2 * ( t ) + φ ˜ 1 * ( t ) + z ˜ k 1 * ( t ) , x ˜ ( k 2 ) ( t ) + φ ˜ 1 * ( t ) + z ˜ k * ( t ) , x ˜ ( k 1 ) ( t ) H F ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , x ˜ * ( t ) ) = M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) .

And also, we have

(35) ( z ˜ 1 * ( t ) , , z ˜ k * ( t ) ) g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) .

By Theorem 1.27 [4], that is equivalent to

(36) g * ( z ˜ 1 * ( t ) , , z ˜ k * ( t ) , t ) + g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) = x ˜ ( t ) , z ˜ 1 * ( t ) + x ˜ ( t ) , z ˜ 2 * ( t ) + + x ˜ ( k 1 ) ( t ) , z ˜ k * ( t ) .

Therefore, it is guaranteed that by taking into account the relations (34)–(36) for x ˜ ( ) and { ˜ k , j , i ( T , 0 ) : ( z ˜ 1 * ( t ) , , z ˜ k * ( t ) ) g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) } , the primal and dual problems have the same optimal values. We conclude that if x ˜ ( ) is optimal solution of the primal problem and the family { ˜ k , j , i ( T , 0 ) : ( z ˜ 1 * ( t ) , , z ˜ k * ( t ) ) g ( x ˜ ( t ) , , x ˜ ( k 1 ) ( t ) , t ) } satisfying conditions (1)–(4) of Theorem 3.1 is optimal solution of dual problem, there is no gap between primal and dual problem. The proof of converse statement for the theorem can be shown similarly. By Lemma 2.6 [4, p. 64], equation (34) implies equation (25) for our problem; therefore, this leads to the Euler-Lagrange-type inclusion (1) of Theorem 3.1. Because the LAM F * is not empty, Theorem 3.1’s condition (4) is met. Transversality conditions (2) and (3) of Theorem 3.1 are fulfilled as they are supposed on the set k , j , i ( T , 0 ) . Then, we complete the proof of theorem.□

Next, we formulate the dual problem corresponding to our linear problem (PHL) with higher-order DFI and special boundary-values, where F ( x , v 1 , , v k 1 ) = A 0 x + j = 1 k 1 A j v j + B U . Then, we calculate the M F function:

(37) M F ( x * , v 1 * , , v k 1 * , v * ) = inf ( x , v 1 , , v k 1 , v ) g p h F { x , x * + v 1 , v 1 * + + v k 1 , v k 1 * v , v * } = inf x , v 1 , , v k 1 [ x , x * A 0 * v * + v 1 , v 1 * A 1 * v * + + v k 1 , v k 1 * A k 1 * v * ] sup w U u , B * v * = W U ( B * v * ) , if x * = A 0 * v * , , v k 1 * = A k 1 * v * , otherwise.

Then, by virtue of (37), we can express M F as follows:

M F ( ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) , , φ 1 * ( t ) + z k * ( t ) , x * ( t ) ) = W U ( B * x * ) , if ( 1 ) k x * ( k ) ( t ) + φ k 1 * ( t ) + z 1 * ( t ) = A 0 * x * ( t ) , φ k 1 * ( t ) + φ k 2 * ( t ) + z 2 * ( t ) = A 1 * x * ( t ) , φ 2 * ( t ) + φ 1 * ( t ) + z k 1 * ( t ) = A k 2 * x * ( t ) , φ 1 * ( t ) + z k * ( t ) = A k 1 * x * ( t ) , otherwise.

We assumed that g has continuous first partial derivatives in problem (PHL), we can immediately conclude the following:

(38) ( z 1 * ( t ) , z 2 * ( t ) , , z k * ( t ) ) = g x , g x , , g x ( k 1 ) .

By sequentially differentiating φ j * ( t ) , j = 1 , , k 1 and inserting into the previous equation, it can be deduced that

(39) φ j * ( t ) = A k j * x * ( t ) A k j + 1 * x * ( t ) + + ( 1 ) j 1 A k 1 * x * ( j 1 ) ( t ) g x ( k j ) + d d t g x ( k j + 1 ) + ( 1 ) j d j 1 d t j 1 g x ( k 1 ) , j = 1 , , k 1 ,

and consequently, it follows that

(40) ( 1 ) k d k x * ( t ) d t k = j = 0 k 1 ( 1 ) j A j * x * ( j ) ( t ) g x + d d t g x + ( 1 ) k 1 d k 1 d t k 1 g x ( k 1 ) .

Note that the latter equation (40) is the same as equation (20). It shows that x * ( t ) in both primal and dual problems satisfies the same equation for optimality. Then, using (39) the dual problem (PHL*) can be readily obtained as

sup x * ( ) 0 T g * ( z 1 * ( t ) , z 2 * ( t ) , , z k * ( t ) , t ) d t 0 T W U ( B * x * ) d t μ k 1 , x * ( 0 ) ( PHL * ) j = 1 k 1 μ k j 1 , ( 1 ) j + 1 x * ( j ) ( 0 ) A k j * x * ( 0 ) + A k j + 1 * x * ( 0 ) + ( 1 ) j A k 1 * x * ( j 1 ) ( 0 ) + g x ( k j ) t = 0 d d t g x ( k j + 1 ) t = 0 + + ( 1 ) j 1 d j 1 d t j 1 g x ( k 1 ) t = 0 ,

where x * ( ) is a solution of the adjoint Euler-Lagrange-type equation (40) and z m * ( ) satisfies equation (38) for m = 1 , , k .

Theorem 5.2

Assume that g ( , t ) is a continuously differentiable with respect to each variable and proper convex function and F be a convex closed multi-valued mapping. Suppose that x ˜ ( t ) is an optimal solution of the primal problem (PHL). The function x ˜ ( t ) for t [ 0 , T ] serves as an optimal solution to the dual problem (PHL) if and only if the conditions of Theorem 4.1 are fulfilled. Besides, the dual (PHL*) and primal (PHL) problems have the same optimal values.

6 Duality for the third-order polyhedral DFIs

Here, we aim to construct the dual problem ( PLC * ) of the primal problem (PLC), which is defined with the third-order polyhedral DFI:

infimum J [ x ( t ) ] = 0 T g ( x ( t ) , x ( t ) , x ( t ) , t ) d t , (PLC) x ( t ) F ( x ( t ) , x ( t ) , x ( t ) ) , a.e. t [ 0 , T ] , F ( x , v 1 , v 2 ) = { v : A x + B v 1 + C v 2 D v d } , x ( s ) ( 0 ) x ( s ) ( T ) = μ s , s = 0 , 1 , 2 ,

where A , B , C , D are m × n dimensional matrices, d is a m -dimensional column-vector, and g ( , t ) is a real-valued function with continuous first partial derivatives. The objective is to determine the trajectory x ˜ ( ) of problem (PLC) minimizing the Lagrange functional J [ x ( t ) ] . First, let us present the theorem on optimality conditions.

Theorem 6.1

Assume g ( , t ) : R 3 n R 1 is continuously differentiable with respect to each variable and proper convex function and F is a third-order polyhedral multi-valued mapping for the problem (PLC). For the optimality of the trajectory x ˜ ( ) in problem (PLC), it suffices to have a nonnegative function λ ( t ) 0 for t [ 0 , T ] , which satisfies both the following third-order Euler-Lagrange-type polyhedral differential equation almost everywhere and transversality conditions at the endpoints t = 0 and t = T :

( 1 ) D * λ ( t ) + C * λ ( t ) B * λ ( t ) + A * λ ( t ) = d 2 d t 2 g ˜ x + d d t g ˜ x g ˜ x , λ ( t ) 0 , A x ˜ ( t ) + B x ˜ ( t ) + C x ˜ ( t ) D x ˜ ( t ) d , λ ( t ) = 0 , a.e. t [ 0 , T ] . ( 2 ) D * ( λ ( s ) ( 0 ) λ ( s ) ( T ) ) = 0 , s = 0 , 1 , 2 , C * ( λ ( 0 ) λ ( T ) ) = g ( x ˜ ( T ) , x ˜ ( T ) , x ˜ ( T ) , T ) x g ( x ˜ ( 0 ) , x ˜ ( 0 ) , x ˜ ( 0 ) , 0 ) x , C * ( λ ( 0 ) λ ( T ) ) B * ( λ ( 0 ) λ ( T ) ) = g ( x ˜ ( T ) , x ˜ ( T ) , x ˜ ( T ) , T ) x g ( x ˜ ( 0 ) , x ˜ ( 0 ) , x ˜ ( 0 ) , 0 ) x + g ( x ˜ ( 0 ) , x ˜ ( 0 ) , x ˜ ( 0 ) , 0 ) x g ( x ˜ ( T ) , x ˜ ( T ) , x ˜ ( T ) , T ) x .

Proof

By applying the Euler-Lagrange-type inclusion (1) of Theorem 3.1, we find:

( x * ( t ) + φ 2 * ( t ) , φ 2 * ( t ) + φ 1 * ( t ) , φ 1 * ( t ) ) F * ( x * ( t ) ; ( x ˜ ( t ) , x ˜ ( t ) , x ˜ ( t ) , x ˜ ( t ) ) , t ) ( x , x , x ) g ( x ˜ ( t ) , x ˜ ( t ) , x ˜ ( t ) , t ) .

Therefore, it is necessary to compute the LAM F * ( v * ; ( x , v 1 , v 2 , v ) ) . Clearly, in that particular instance:

gph F = { ( x , v 1 , v 2 , v ) : A x + B v 1 + C v 2 D v d } .

Consider A i , B i , C i , and D i as the i th row of matrices A , B , C , and D respectively, and d i as the i th component of vector d . Suppose that w ˜ = ( x ˜ , v ˜ 1 , v ˜ 2 , v ˜ ) gph F and the following:

I ( w ˜ ) = { i : A i x ˜ + B i v ˜ 1 + C i v ˜ 2 D i v ˜ = d i , i = 1 , , m } .

Using the meaning of cone of tangent directions K gph F ( w ˜ ) = { w ¯ : w ˜ + γ w ¯ gph F for sufficiently small γ > 0 } , we can easily obtain the inequality for i I ( w ˜ ) if A i x ¯ + B i v ¯ 1 + C i v ¯ 2 D i v ¯ 0 , i I ( w ˜ ) :

A i ( x ˜ + γ x ¯ ) + B i ( v ˜ 1 + γ v ¯ 1 ) + C i ( v ˜ 2 + γ v ¯ 2 ) D i ( v ˜ + γ v ¯ ) = d i + γ ( A i x ¯ + B i v ¯ 1 + C i v ¯ 2 D i v ¯ ) d i .

Nevertheless, when i I ( w ˜ ) , we obtain the following expression:

A i ( x ˜ + γ x ¯ ) + B i ( v ˜ 1 + γ v ¯ 1 ) + C i ( v ˜ 2 + γ v ¯ 2 ) D i ( v ˜ + γ v ¯ ) = ( A i x ˜ + B i v ˜ 1 + C i v ˜ 2 D i v ˜ ) + γ ( A i x ¯ + B i v ¯ 1 + C i v ¯ 2 D i v ¯ ) < d i .

Irrespective of the selection of w ¯ = ( x ¯ , v ¯ 1 , v ¯ 2 , v ¯ ) , it is accurate for sufficiently small γ > 0 . That implies K gphF ( w ˜ ) = { ( w ¯ : A i x ¯ + B i v ¯ 1 + C i v ¯ 2 D i v ¯ ) 0 , i I ( w ˜ ) } . It is easy to note that ( x * , v 1 * , v 2 * , v * ) K gph F * ( w ˜ ) by  applying Farkas theorem [4, p. 22] if and only if

(41) x * = i I ( w ˜ ) A i * λ i , v 1 * = i I ( w ˜ ) B i * λ i , v 2 * = i I ( w ˜ ) C i * λ i , v * = i I ( w ˜ ) D i * λ i , λ i 0 , i = 1 , , m ,

where A i * , B i * , C i * , and D i * are the transposed vector-columns. Then, if we set λ i = 0 , for i I ( w ˜ ) and denote the vector-column with components λ i by λ , then the following equality from (41) can be obtained:

K gph F * ( w ˜ ) = { ( x * , v 1 * , v 2 * , v * ) : x * = A * λ , v 1 * = B * λ , v 2 * = C * λ , v * = D * λ , λ 0 , A x ˜ + B v ˜ 1 + C v ˜ 2 D v ˜ d , λ = 0 } .

Therefore, the expression for the polyhedral LAM obtained from (41) is as follows:

(42) F * ( v * ; ( x ˜ , v ˜ 1 , v ˜ 2 , v ˜ ) ) = { ( A * λ , B * λ , C * λ ) : v * = D * λ , λ 0 , A x + B v 1 + C v 2 D v d , λ = 0 } .

Equations (41) and (42) actually show that F * ( v * ; ( x ˜ , v ˜ 1 , v ˜ 2 , v ˜ ) ) relies on the set I ( w ˜ ) rather than the point w ˜ = ( x ˜ , v ˜ 1 , v ˜ 2 , v ˜ ) . As there are finitely many such sets, there are also a finite number of distinct LAM F * ( v * ; ( x ˜ , v ˜ 1 , v ˜ 2 , v ˜ ) ) . If we use together (41) and (42), we derive the following relations:

(43) x * ( t ) + φ 2 * ( t ) = A * λ ( t ) g ˜ x , φ 2 * ( t ) + φ 1 * ( t ) = B * λ ( t ) g ˜ x , φ 1 * ( t ) = C * λ ( t ) g ˜ x , x * ( t ) = D * λ ( t ) , λ ( t ) 0 , A x ˜ ( t ) + B x ˜ ( t ) + C x ˜ ( t ) D x ˜ ( t ) d , λ ( t ) = 0 , a.e. t [ 0 , T ] .

The first condition of the theorem is derived by differentiating φ j * ( t ) , j = 1,2 and sequentially replacing each φ j * ( t ) in the preceding equation in (43):

(44) φ 2 * ( t ) = C * λ ( t ) + d d t g ˜ x B * λ ( t ) g ˜ x , D * λ ( t ) + C * λ ( t ) B * λ ( t ) + A * λ ( t ) = d 2 d t 2 g ˜ x + d d t g ˜ x g ˜ x .

Therefore, the following expression D * ( λ ( s ) ( 0 ) λ ( s ) ( T ) ) = 0 , s = 0 , 1 , 2 can be used to formulate the transversality condition (2) of Theorem 3.1. Then, if we apply the transversality condition (3) of Theorem 3.1, the theorem has been proven.□

Now the dual problem (PLC*) to the problem (PLC) can be generated. We must initially calculate M F ( x * , v 1 * , v 2 * , v * ) for (PLC*):

(45) M F ( x * , v 1 * , v 2 * , v * ) = inf { x , x * + v 1 , v 1 * + v 2 , v 2 * v , v * : ( x , v 1 , v 2 , v ) gph F } .

Let us denote w = ( x , v 1 , v 2 , v ) R 4 n and w * = ( x * , v 1 * , v 2 * , v * ) R 4 n . Then, a linear programming problem confronts us:

(46) inf { w , w * : L w d } ,

where L = [ A B C D ] is m × 4 n block matrix. Hence, when w ˜ = ( x ˜ , v ˜ 1 , v ˜ 2 , v ˜ ) is a solution of (46), there is m -dimensional vector λ 0 having the following relation:

w * = L * λ , A x ˜ + B v ˜ 1 + C v ˜ 2 D v ˜ d , λ = 0 .

Therefore, w * = L * λ implies x * = A * λ , v 1 * = B * λ , v 2 * = C * λ , v * = D * λ , λ 0 . Therefore, we find the following equality:

(47) M F ( x * , v 1 * , v 2 * , v * ) = x ˜ , A * λ + v ˜ 1 , B * λ + v ˜ 2 , C * λ v ˜ , D * λ = A x ˜ , λ B v ˜ 1 , λ C v ˜ 2 , λ + D v ˜ , λ = d , λ .

Applying the subsequent formula M F ( x * ( t ) + φ 2 * ( t ) + z 1 * ( t ) , φ 2 * ( t ) + φ 1 * ( t ) + z 2 * ( t ) , φ 1 * ( t ) + z 3 * ( t ) , x * ( t ) ) , we derive that

(48) x * ( t ) + φ 2 * ( t ) + z 1 * ( t ) = A * λ ( t ) , φ 2 * ( t ) + φ 1 * ( t ) + z 2 * ( t ) = B * λ ( t ) φ 1 * ( t ) + z 3 * ( t ) = C * λ ( t ) , x * ( t ) = D * λ ( t ) , λ ( t ) 0 , ( z 1 * ( t ) , z 2 * ( t ) , z 3 * ( t ) ) = g ˜ x , g ˜ x , g ˜ x ,

or

(49) φ 1 * ( t ) = C * λ ( t ) g ˜ x , φ 2 * ( t ) = C * λ ( t ) B * λ ( t ) + d d t g ˜ x g ˜ x D * λ ( t ) + C * λ ( t ) B * λ ( t ) + A * λ ( t ) = d 2 d t 2 g ˜ x + d d t g ˜ x g ˜ x .

Hence, by taking into account (47)–(49), the dual problem we encounter is as follows:

sup λ ( t ) 0 0 T g * ( z 1 * ( t ) , z 2 * ( t ) , z 3 * ( t ) , t ) d t 0 T d , λ ( t ) d t μ 0 , x * ( 0 ) C * λ ( 0 ) + B * λ ( 0 ) ( P L C * ) d d t g ˜ x t = 0 + g ˜ x t = 0 μ 1 , x * ( 0 ) + C * λ ( 0 ) + g ˜ x t = 0 μ 2 , x * ( 0 ) ,

where ( z 1 * ( t ) , z 2 * ( t ) , z 3 * ( t ) ) = ( g ˜ x , g ˜ x , g ˜ x ) . Finally, we state the duality theorem for the third-order polyhedral problem.

Theorem 6.2

Assume that the conditions of Theorem 3.1hold and x ˜ ( ) is an optimal solution to the third-order polyhedral problem (PLC). Therefore, x ˜ ( t ) for t [ 0 , T ] represents an optimal solution of the dual problem (PLC) if and only if the conditions outlined in Theorem 6.1 are fulfilled. Besides, primal (PLC) and dual (PLC*) problems have the same optimal values.

7 Conclusion

In this article, an arbitrary-order DFI is examined with special boundary conditions, which are particularly important in applied fields. The cost function, which is the Lagrange functional in the addressed problem, is studied for a general case in optimization and its optimality conditions are derived. Mahmudov’s inclusion, which is a higher-order adjoint inclusion, is introduced to our problem and LAM and Hamiltonian functions are utilized. It should be noted that for k = 1 , Mahmudov’s and Euler-Lagrange inclusions are equivalent. Another issue for the presented problem can be necessary conditions for optimality in future research. Therefore, it seems that discrete and discrete approximation problems will play a crucial role in finding necessary conditions and numerical optimal control problems as well. In addition, it is significant to formulate the dual problem from our primal problem and explore optimality conditions. Hence, there is no doubt that obtained theorems for presented problems will have remarkable effects on high-order optimal control problems, and the approaches, which can be used to solve HODIs will contribute to improving mathematical theory of optimization.

Acknowledgments

The authors would like to express sincere thanks to the Editor-in-Chief of the Journal Demonstratio Mathematica, and anonymous reviewers for their valuable and constructive suggestions, which improved the final manuscript.

  1. Funding information: Authors state no funding involved.

  2. Author contributions: UY: supervision, software, reviewing, and investigation. DM: reviewing and investigation. EM: writing–original draft, conceptualization, and methodology.

  3. Conflict of interest: The authors declare that they have no conflict of interest regarding the publication of the research article.

  4. Ethical approval: This article does not contain any studies with human participants or animals performed by any of the authors.

  5. Data availability statement: No data were used to support this study.

References

[1] J. P. Aubin and A. Cellina, Differential Inclusions, Springer, Berlin, Heidelberg, 1984. 10.1007/978-3-642-69512-4Search in Google Scholar

[2] I. Ekeland and R. Temam, Convex Analysis and Variational Problems, 1st edition (Classics in Applied Mathematics), Society for Industrial and Applied Mathematics, Philadelphia, 1987. Search in Google Scholar

[3] P. D. Loewen and R. T. Rockafellar, Optimal control of unbounded differential inclusions, SIAM J. Control Optim. 32 (1994), 442–470, DOI: https://doi.org/10.1137/S0363012991217494. 10.1137/S0363012991217494Search in Google Scholar

[4] E. N. Mahmudov, Approximation and Optimization of Discrete and Differential Inclusions, Boston, USA, Elsevier, 2011. Search in Google Scholar

[5] B. S. Mordukhovich, N. M. Nam, R. B Rector, and T. Tran, Variational geometric approach to generalized differential and conjugate calculi in convex analysis, Set-Valued Var. Anal. 25 (2017), 731–755, DOI: https://doi.org/10.1007/s11228-017-0426-7. 10.1007/s11228-017-0426-7Search in Google Scholar

[6] A. A. Khan, C. Tammer, and C. Zalinescu, Set-Valued Optimization, Springer, Berlin, 2016. 10.1007/978-3-642-54265-7Search in Google Scholar

[7] G. Dinca, P. Jebelean, and D. Motreanu, Existence and approximation for a general class of differential inclusions, Houston J. Math 28 (2002), no. 1, 193–215. Search in Google Scholar

[8] C. Izuchukwu, S. Reich, and Y. Shehu, Relaxed inertial methods for solving the split monotone variational inclusion problem beyond co-coerciveness, Optimization 72 (2023), no. 2, 607–646, DOI: https://doi.org/10.1080/02331934.2021.1981895. 10.1080/02331934.2021.1981895Search in Google Scholar

[9] T. Kamran, C. Vetro, M. U. Ali, and M. Waheed, A fixed point theorem for G-monotone multivalued mapping with application to nonlinear integral equations, Filomat 31 (2017), no. 7, 2045–2052, DOI: https://doi.org/10.2298/FIL1707045K. 10.2298/FIL1707045KSearch in Google Scholar

[10] S. Zeng, Y. Bai, V. D. Rădulescu, and P. Winkert, An inverse problem for a double phase implicit obstacle problem with multivalued terms, ESAIM Control Optim. Calc. Var. 29 (2023), 30, DOI: https://doi.org/10.1051/cocv/2023022. 10.1051/cocv/2023022Search in Google Scholar

[11] E. N. Mahmudov, Approximation and optimization of Darboux type differential inclusions with set-valued boundary conditions, Optim. Lett. 7 (2013), 871–891, DOI: https://doi.org/10.1007/s11590-012-0460-1. 10.1007/s11590-012-0460-1Search in Google Scholar

[12] F. Guan, J. Cen, B. Chen, and J. Yao, Existence of projected solutions for quasi-variational hemivariational inequality, Demonstr. Math. 57 (2024), no. 1, 20230139, DOI: https://doi.org/10.1515/dema-2023-0139. 10.1515/dema-2023-0139Search in Google Scholar

[13] E. N. Mahmudov, The optimality principle for discrete and first order partial differential inclusions, J. Math. Anal. Appl. 308 (2005), 605–619, DOI: https://doi.org/10.1016/j.jmaa.2004.11.047. 10.1016/j.jmaa.2004.11.047Search in Google Scholar

[14] E. N. Mahmudov, Convex optimization of second order discrete and differential inclusions with inequality constraints, J. Convex Anal. 25 (2018), 293–318. Search in Google Scholar

[15] E. N. Mahmudov, Optimal control of higher order differential inclusions with functional constraints, ESAIM Control Optim. Calc. Var. 26 (2020), 1–23, DOI: https://doi.org/10.1051/cocv/2019018. 10.1051/cocv/2019018Search in Google Scholar

[16] E. N. Mahmudov, Locally adjoint mappings and optimization of the first boundary-value problem for hyperbolic type discrete and differential inclusions, Nonlinear Anal. 67 (2007), 2966–2981, DOI: https://doi.org/10.1016/j.na.2006.09.054. 10.1016/j.na.2006.09.054Search in Google Scholar

[17] E. N. Mahmudov, Approximation and optimization of discrete and differential inclusions described by inequality constraints, Optimization 63 (2014), 1117–1133, DOI: https://doi.org/10.1080/02331934.2012.713360. 10.1080/02331934.2012.713360Search in Google Scholar

[18] E. N Mahmudov, Mathematical programming and polyhedral optimization of second order discrete and differential inclusions, Pac. J. Optim. 11 (2015), 511–525. Search in Google Scholar

[19] E. N. Mahmudov, Free time optimization of higher order differential inclusions with endpoint constraints, Appl. Anal. 97 (2018), 2071–2084, DOI: https://doi.org/10.1080/00036811.2017.1350851. 10.1080/00036811.2017.1350851Search in Google Scholar

[20] E. N. Mahmudov, Optimization of Mayer problem with Sturm-Liouville-type differential inclusions, J. Optim. Theory Appl. 177 (2018), 345–375, DOI: https://doi.org/10.1007/s10957-018-1260-2. 10.1007/s10957-018-1260-2Search in Google Scholar

[21] E. N. Mahmudov and M. Mardanov, Infimal convolution and duality in convex mathematical programming, Proc. Inst. Math. Mech. Natl. Acad. Sci. Azerb. 48 (2022), 50–62, DOI: https://doi.org/10.30546/2409-4994.48.1.2022.50. 10.30546/2409-4994.48.1.2022.50Search in Google Scholar

[22] E. N. Mahmudov and D. I. Mastaliyeva, Optimal control of higher-order retarded differential inclusions, J. Ind. Manag. Optim. 19 (2023), 6544–6557, DOI: https://doi.org/10.3934/jimo.2022226. 10.3934/jimo.2022226Search in Google Scholar

[23] B. S. Mordukhovich, Optimal control of semilinear unbounded evolution inclusions with functional constraints, J. Optim. Theory Appl. 167 (2015), 821–841, DOI: https://doi.org/10.1007/s10957-013-0301-0. 10.1007/s10957-013-0301-0Search in Google Scholar

[24] B. S. Mordukhovich, Variational Analysis and Generalized Differentiation, I: Basic Theory; II: Applications, Grundlehren Series (Fundamental Principles of Mathematical Sciences), vol. 330 and and 331, Springer, Berlin, Heidelberg, 2006. Search in Google Scholar

[25] L. S. Pontryagin, V. G. Boltyanskii, R. V. Gamkrelidze, and E. F. Mishchenko, The Mathematical Theory of Optimal Processes, John Wiley Sons, Inc., New York, London, Sydney, 1965. Search in Google Scholar

[26] A. Cernea, On a boundary-value problem for a third order differential inclusion, Demonstr. Math. 42 (2009), no. 4, 723–730, DOI: https://doi.org/10.1515/dema-2009-0407. 10.1515/dema-2009-0407Search in Google Scholar

[27] J. Andres, A target problem for differential inclusions with state-space constraints, Demonstr. Math. 30 (1997), no. 4, 783–790, DOI: https://doi.org/10.1515/dema-1997-0409. 10.1515/dema-1997-0409Search in Google Scholar

[28] R. I. Bot, Conjugate Duality in Convex Optimization, Lecture Notes in Economics and Mathematical Systems, Springer-Verlag, Berlin, Heidelberg, 2010. 10.1007/978-3-642-04900-2Search in Google Scholar

[29] R. T. Rockafellar, Conjugate Duality and Optimization, Society for Industrial and Applied Mathematics, SIAM, Philadelphia, 1974. 10.1137/1.9781611970524Search in Google Scholar

[30] E. Galewska, A general statement of dual first-order sufficient optimality conditions for the generalized problem of Bolza, Demonstr. Math. 36 (2003), no. 3, 711–720, DOI: https://doi.org/10.1515/dema-2003-0320. 10.1515/dema-2003-0320Search in Google Scholar

[31] A. Bressan, On differential relations with lower continuous right-hand side. An existence theorem, J. Differential Equations 37 (1980), 89–97, DOI: https://doi.org/10.1016/0022-0396(80)90090-X. 10.1016/0022-0396(80)90090-XSearch in Google Scholar

[32] N. S. Papageorgiou, C. Vetro, and F. Vetro, Extremal solutions and strong relaxation for nonlinear multivalued systems with maximalmonotone terms, J. Math. Anal. Appl. 461 (2018), 401–421, DOI: https://doi.org/10.1016/j.jmaa.2018.01.009. 10.1016/j.jmaa.2018.01.009Search in Google Scholar

[33] N. S. Papageorgiou, C. Vetro, and F. Vetro, Nonlinear vector Duffing inclusions with no growth restriction on the orientor field, Topol. Methods Nonlinear Anal. 54 (2019), no. 1, 257–274, DOI: https://doi.org/10.12775/TMNA.2019.041. 10.12775/TMNA.2019.041Search in Google Scholar

[34] K. Zhang and J. Xu, Solvability for a system of Hadamard-type hybrid fractional differential inclusions, Demonstr. Math. 56 (2023), no. 1, 20220226, DOI: https://doi.org/10.1515/dema-2022-0226. 10.1515/dema-2022-0226Search in Google Scholar

[35] B. Damjanovic, B. Samet, and C. Vetro, Common fixed point theorems for multi-valued maps, Acta Math. Sci. 32 (2012), no. 2, 818–824, DOI: https://doi.org/10.1016/S0252-9602(12)60063-0. 10.1016/S0252-9602(12)60063-0Search in Google Scholar

[36] S. A. Khan, K. R. Kazmi, W. Cholamjiak, and H. Dutta, Convergence analysis for combination of equilibrium problems and k-nonspreading multi-valued mappings, Proyecciones 39 (2021), no. 3, 599–619, DOI: https://doi.org/10.22199/issn.0717-6279-2020-03-0037. 10.22199/issn.0717-6279-2020-03-0037Search in Google Scholar

[37] S. Kumar and S. Luambano, On some fixed point theorems for multivalued F-contractions in partial metric spaces, Demonstr. Math. 54 (2021), no. 1, 151–161, DOI: https://doi.org/10.1515/dema-2021-0012. 10.1515/dema-2021-0012Search in Google Scholar

[38] S. Domachowski, Local bounds and existence of solutions to non-convex differential inclusions, Demonstr. Math. 42 (2009), no. 1, 65–74, DOI: https://doi.org/10.1515/dema-2009-0107. 10.1515/dema-2009-0107Search in Google Scholar

[39] S. Abbas and M. Benchohra, Ulam stabilities for the Darboux problem for partial fractional differential inclusions, Demonstr. Math. 47 (2014), no. 4, 826–838, DOI: https://doi.org/10.2478/dema-2014-0066. 10.2478/dema-2014-0066Search in Google Scholar

[40] R. P. Agarwal, S. R. Grace, and D. O’Regan, Oscillation theorems for second order differential inclusions, Int. J. Dyn. Syst. Differ. Equ. 1 (2007), 85–88, DOI: https://doi.org/10.1504/IJDSDE.2007.016509. 10.1504/IJDSDE.2007.016509Search in Google Scholar

[41] A. Auslender and J. Mechler, Second order viability problems for differential inclusions, J. Math. Anal. Appl. 181 (1994), 205–218, DOI: https://doi.org/10.1006/jmaa.1994.1015. 10.1006/jmaa.1994.1015Search in Google Scholar

[42] S. S. Yusubov, Nonlocal problem with integral conditions for a high-order hyperbolic equation, Ukrainian Math. J. 69 (2017), no. 1, 121–131, DOI: https://doi.org/10.1007/s11253-017-1353-3. 10.1007/s11253-017-1353-3Search in Google Scholar

[43] G. Bartuzel and A. Fryszkowski, Filippov lemma for certain differential inclusion of third order, Demonstr. Math. 41 (2008), no. 2, 337–352, DOI: https://doi.org/10.1515/dema-2008-0211. 10.1515/dema-2008-0211Search in Google Scholar

[44] M. Frigon and T. Kaczynski, Boundary value problems for systems of implicit differential equations, J. Math. Anal. Appl. 179 (1993), 317–326, DOI: https://doi.org/10.1006/jmaa.1993.1353. 10.1006/jmaa.1993.1353Search in Google Scholar

[45] D. L. Azzam, C. Castaing, and L. Thibault, Three boundary-value problems for second order differential inclusions in Banach spaces, Control Cybernet. 31 (2002), 659–693. Search in Google Scholar

[46] M. Benchohra, J. R. Graef, J. Henderson, and S. K. Ntouyas, Nonresonance impulsive higher order functional nonconvex-valued differential inclusions, Electron. J. Qual. Theory Differ. Equ. 13 (2002), 1–13. 10.14232/ejqtde.2002.1.13Search in Google Scholar

[47] N. Matzakos and N. S. Papageorgiou, Periodic solutions for evolution inclusions with time-dependent subdifferentials, Demonstr. Math. 35 (2002), no. 3, 521–538, DOI: https://doi.org/10.1515/dema-2002-0310. 10.1515/dema-2002-0310Search in Google Scholar

[48] E. N. Mahmudov, Single Variable Differential and Integral Calculus, Springer, Paris, 2013. 10.2991/978-94-91216-86-2Search in Google Scholar

Received: 2024-03-17
Revised: 2024-08-22
Accepted: 2025-02-11
Published Online: 2025-06-27

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. On approximation by Stancu variant of Bernstein-Durrmeyer-type operators in movable compact disks
  3. Circular n,m-rung orthopair fuzzy sets and their applications in multicriteria decision-making
  4. Grand Triebel-Lizorkin-Morrey spaces
  5. Coefficient estimates and Fekete-Szegö problem for some classes of univalent functions generalized to a complex order
  6. Proofs of two conjectures involving sums of normalized Narayana numbers
  7. On the Laguerre polynomial approximation errors and lower type of entire functions of irregular growth
  8. New convolutions for the Hartley integral transform imbedded in the Banach algebras and convolution-type integral equations
  9. Some inequalities for rational function with prescribed poles and restricted zeros
  10. Lucas difference sequence spaces defined by Orlicz function in 2-normed spaces
  11. Evaluating the efficacy of fuzzy Bayesian networks for financial risk assessment
  12. Fixed point results for contractions of polynomial type
  13. Estimation for spatial semi-functional partial linear regression model with missing response at random
  14. Investigating the controllability of differential systems with nonlinear fractional delays, characterized by the order 0 < η ≤ 1 < ζ ≤ 2
  15. New forms of bilateral inequalities for K-g-frames
  16. Rate of pole detection using Padé approximants to polynomial expansions
  17. Existence results for nonhomogeneous Choquard equation involving p-biharmonic operator and critical growth
  18. Note on the shape-preservation of a new class of Kantorovich-type operators via divided differences
  19. Geršhgorin-type theorems for Z1-eigenvalues of tensors with applications
  20. New topologies derived from the old one via operators
  21. Blow up solutions for two-dimensional semilinear elliptic problem of Liouville type with nonlinear gradient terms
  22. Infinitely many normalized solutions for Schrödinger equations with local sublinear nonlinearity
  23. Nonparametric expectile shortfall regression for functional data
  24. Advancing analytical solutions: Novel wave insights and methodologies for beta fractional Kuralay-II equations
  25. A generalized p-Laplacian problem with parameters
  26. A study of solutions for several classes of systems of complex nonlinear partial differential difference equations in ℂ2
  27. Towards finding equalities involving mixed products of the Moore-Penrose and group inverses by matrix rank methodology
  28. ω -biprojective and ω ¯ -contractible Banach algebras
  29. Coefficient functionals for Sakaguchi-type-Starlike functions subordinated to the three-leaf function
  30. Solutions of several general quadratic partial differential-difference equations in ℂ2
  31. Inequalities for the generalized trigonometric functions with respect to weighted power mean
  32. Optimization of Lagrange problem with higher-order differential inclusion and special boundary-value conditions
  33. Hankel determinants for q-starlike functions connected with q-sine function
  34. System of partial differential hemivariational inequalities involving nonlocal boundary conditions
  35. A new family of multivalent functions defined by certain forms of the quantum integral operator
  36. A matrix approach to compare BLUEs under a linear regression model and its two competing restricted models with applications
  37. Weighted composition operators on bicomplex Lorentz spaces with their characterization and properties
  38. Behavior of spatial curves under different transformations in Euclidean 4-space
  39. Commutators for the maximal and sharp functions with weighted Lipschitz functions on weighted Morrey spaces
  40. A new kind of Durrmeyer-Stancu-type operators
  41. A study of generalized Mittag-Leffler-type function of arbitrary order
  42. On the approximation of Kantorovich-type Szàsz-Charlier operators
  43. Split quaternion Fourier transforms for two-dimensional real invariant field
  44. Review Article
  45. Characterization generalized derivations of tensor products of nonassociative algebras
  46. Special Issue on Differential Equations and Numerical Analysis - Part II
  47. Existence and optimal control of Hilfer fractional evolution equations
  48. Persistence of a unique periodic wave train in convecting shallow water fluid
  49. Existence results for critical growth Kohn-Laplace equations with jumping nonlinearities
  50. Monotonicity and oscillation for fractional differential equations with Riemann-Liouville derivatives
  51. Nontrivial solutions for a generalized poly-Laplacian system on finite graphs
  52. Stability and bifurcation analysis of a modified chemostat model
  53. Special Issue on Nonlinear Evolution Equations and Their Applications - Part II
  54. Analytic solutions of a generalized complex multi-dimensional system with fractional order
  55. Extraction of soliton solutions and Painlevé test for fractional Peyrard-Bishop DNA model
  56. Special Issue on Recent Developments in Fixed-Point Theory and Applications - Part II
  57. Some fixed point results with the vector degree of nondensifiability in generalized Banach spaces and application on coupled Caputo fractional delay differential equations
  58. On the sum form functional equation related to diversity index
  59. Special Issue on International E-Conference on Mathematical and Statistical Sciences - Part II
  60. Simpson, midpoint, and trapezoid-type inequalities for multiplicatively s-convex functions
  61. Converses of nabla Pachpatte-type dynamic inequalities on arbitrary time scales
  62. Special Issue on Blow-up Phenomena in Nonlinear Equations of Mathematical Physics - Part II
  63. Energy decay of a coupled system involving a biharmonic Schrödinger equation with an internal fractional damping
  64. Special Issue on Some Integral Inequalities, Integral Equations, and Applications - Part II
  65. Nonlinear heat equation with viscoelastic term: Global existence and blowup in finite time
  66. New Jensen's bounds for HA-convex mappings with applications to Shannon entropy
  67. Special Issue on Approximation Theory and Special Functions 2024 conference
  68. Ulam-type stability for Caputo-type fractional delay differential equations
  69. Faster approximation to multivariate functions by combined Bernstein-Taylor operators
  70. (λ, ψ)-Bernstein-Kantorovich operators
  71. Some special functions and cylindrical diffusion equation on α-time scale
  72. (q, p)-Mixing Bloch maps
  73. Orthogonalizing q-Bernoulli polynomials
  74. On better approximation order for the max-product Meyer-König and Zeller operator
  75. Moment-based approximation for a renewal reward process with generalized gamma-distributed interference of chance
  76. Special Issue on Variational Methods and Nonlinear PDEs
  77. A note on mean field type equations
  78. Ground states for fractional Kirchhoff double-phase problem with logarithmic nonlinearity
  79. Solution of nonlinear Langevin equations involving Hilfer-Hadamard fractional order derivatives and variable coefficients
Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/dema-2025-0106/html?lang=en
Scroll to top button