1. Introduction
In the past decades, extensive research has been devoted to optimal control problems governed by integral equations. Foundational results for Fredholm-type systems can be found in [
1,
2] while Volterra-type formulations including nonlinear settings and state–control constraints are treated in [
3]. Further contributions appear in [
4,
5,
6]. Applications arise in population dynamics, viscoelastic systems, and epidemic models, often under linear dynamics or simplified kernels. However, many of these works rely on restrictive structural assumptions and do not fully exploit the general Dubovitskii–Milyutin framework.
Our work begins with a new lemma that characterizes the controllability of linear control systems governed by Volterra integral equations. This result, significant in its own right, provides a novel contribution to the theory of integral equations. More importantly, it allows us to eliminate the common assumption regarding the controllability of the variational linear equation around an optimal pair (see Equation (
19)).
Several recent contributions have addressed maximum principles for Volterra-type control systems. In [
7], systems with unilateral constraints are considered in linear–quadratic frameworks. In [
8], nonlinear Volterra equations are studied via discrete-time approximations and modified Hamiltonian methods. A maximum principle for singular kernels and terminal constraints with fractional Caputo dynamics appears in [
9]. A rigorous analysis under smooth nonlinearities and mixed control constraints, based on Dubovitskii–Milyutin theory, is presented in [
10]. A numerical collocation scheme using Genocchi polynomials for weakly singular kernels is developed in [
11].
While each of these studies provides valuable insights, most are limited by restrictive kernel assumptions, simplified dynamics, or narrow cost functionals. By contrast, the present work establishes a general Pontryagin-type maximum principle for nonlinear Volterra systems subject to both terminal and time-dependent state constraints. Building upon our controllability result, we derive necessary optimality conditions using conic approximations within the Dubovitskii–Milyutin framework. It is also worth noting that, beyond the derivation of necessary conditions, we establish
sufficient conditions for optimality. In particular, under convexity of the cost functional and linear Volterra dynamics, the maximum principle obtained in Theorem 1 becomes a sufficient criterion for global optimality. This result (
Section 5) shows that our framework extends Pontryagin’s maximum principle to integral equations and recovers its classical sufficiency in the differential equation case.
In the case where only terminal constraints are imposed, the adjoint equation simplifies to a modified Volterra integral equation. In the presence of time-dependent state constraints, the adjoint involves a Stieltjes integral with respect to a non-negative Borel measure concentrated on the active set. This structure captures the added complexity of pathwise constraints. Notably, when the Volterra kernel collapses to a differential operator, our results recover the classical Pontryagin maximum principle. Thus, our findings both unify and extend the existing theory for optimal control problems governed by integral and differential systems.
The Dubovitskii–Milyutin method has proved effective across a wide class of optimal control problems, see, for instance, refs. [
12,
13,
14]. In particular, ref. [
14] analyzes impulsive control problems without state constraints.
Beyond these theoretical contributions, our use of local cone approximations resonates with developments in numerical analysis and convex approximation methods. Moving asymptotes and local convex surrogates offer computational insights for constrained problems [
15], while enrichment techniques in finite element analysis (e.g., truncated Gegenbauer–Hermite weighted approaches) highlight application domains such as viscoelasticity [
16].
As a natural extension, the same methodology may be applied to fractional dynamics (see [
17,
18]), provided that the linear variational equation around an optimal pair is explicitly characterized. Analogous programs apply to integro-dynamic equations on time scales [
19] or to problems in spaces of functions of bounded variation [
20].
Finally, our results display a natural symmetry between the adjoint equation and the maximum condition. In the adjoint, derivatives act with respect to the state variables, while in the maximum condition they act with respect to the controls. This duality reflects the operator transposition inherent in the Dubovitskii–Milyutin framework, highlighting the structural balance between primal and dual formulations and situating our contribution within symmetry analysis.
2. Problem Statement and Principal Results
Having established the necessary background, we now formulate the optimal control problem studied in this paper. Our aim is to derive the necessary optimality conditions, specifically, a Pontryagin-type maximum principle for systems governed by nonlinear Volterra integral equations in which the state trajectory is subject to both control and state constraints.
We therefore consider the following optimal control problem, which constitutes the core of this work. The objective is to obtain Pontryagin’s necessary conditions for a system whose dynamics follows a nonlinear Volterra integral equation with general constraints on the control and the state, including fixed terminal constraints and time-dependent state bounds. The cost functional contains a terminal term and an integral term that may depend on the state. The general formulation is the following.
Problem 1. where we employ the notation “locmin” to denote a local minimum
of the cost functional. Here, denotes the state and the control, which is measurable and takes values in the admissible set . The pair lies in the product space , with and . In particular, the state is continuous, and the control is essentially bounded. The map represents the terminal cost, is the state constraint function, and is the running cost, depending on , , and time t. The function is given, and the nonlinear kernel specifies the Volterra-type dynamics.
We now state the standing assumptions of this paper.
Hypothesis 1. - (a)
The functions φ, K, h, are continuous on compact subsets of , and with .
- (b)
is closed and convex, with .
- (c)
g is continuous and satisfies and , where are fixed. Moreover, g has a continuous derivative with respect to its first variable, , such that whenever .
Now, the more involved version of the maximum principle can be stated. In this setting, the corresponding adjoint equation features a Stieltjes integral with respect to a non-negative Borel measure, thereby accounting for the effect of the state constraints at each time along the trajectory.
Theorem 1. Assume that conditions (a)–(c) hold for , which is a solution to Problem 1.
- (α)
There exists a non-negative Borel measure μ on with support contained in - (β)
There exist and a function such that and ψ do not vanish simultaneously, and ψ is a solution of the adjoint integral equationfor some . Moreover, for all and for almost every one has
Remark 1. If condition (6) is replaced by the following terminal state constraint:then a simpler version of the maximum principle can be established. In this case, the corresponding adjoint equation does not involve a Stieltjes integral with respect to a non-negative Borel measure, as the state constraint only acts at the terminal time. Theorem 2. Assume that conditions (a), (b), and (8) hold and that g has a continuous derivative with respect to its first variable, , with . Let be a solution of Problem 1. Then, there exist , and a function such that and ψ do not vanish simultaneously, and ψ solves the systemand, moreover, for all and for almost every one has Corollary 1. Assume that hypotheses (a)–(c) hold for , a solution to Problem 1, and that - (α)
There exists a non-negative Borel measure μ on whose support is contained in - (β)
There exist and a function such that and ψ do not vanish simultaneously, and ψ solves the integral equationfor some . Moreover, for all and for almost every one has
3. Preliminary Results
This section begins with a lemma that provides a characterization of the controllability of linear control systems governed by Volterra integral equations. The lemma is of independent interest and constitutes a novel contribution to the theory of integral equations. It also enables us to show that the commonly imposed controllability of the variational linear equation around an optimal pair is, in fact, redundant.
We then give a concise overview of the Dubovitskii–Milyutin (DM) framework. We formulate the constrained optimization problem and construct approximation cones for both the objective and the constraints. The optimality condition, written in an Euler–Lagrange (EL) form, is expressed via the corresponding dual cones. These statements rely on standard principles of DM theory; complete proofs can be found in [
13,
14,
21].
3.1. Controllability of a Linear Volterra Equation
In this subsection, we state a lemma that characterizes the controllability of the linear control system
where
is the state and
is the control. The matrix
is
and
is
, with all entries in
, where
. In particular,
A and
B are continuously differentiable in both variables on the triangular domain
.
Lemma 1. System (13) is not controllable if and only if there exists a nonzero function such thatandfor all . Proof. Assume first that (
13) is not controllable. Then, the set
is a proper linear subspace; that is,
. Hence, there exists
,
, such that
Let
be the solution of the equation
Fix any
. Let
be the corresponding solution of the Equation (
13). Then,
In addition,
Integrating by parts and using
, we obtain
Consequently,
for every
. Therefore,
Conversely, assume that there exists
with
satisfying (
14) and (
15). For any solution
x of (
13), repeating the steps above and integrating by parts yields
Thus,
D admits
as a perpendicular vector, so
D is not dense in
; i.e.,
. Hence, system (
13) is not controllable. □
3.2. DM Theorem
Let E be a locally convex topological vector space and let denote its space of continuous linear functionals.
A subset will be called a cone with vertex at the origin if it is closed under positive scalings; i.e., for every .
The
dual cone associated with ⋀ is
We recall a collection of standard consequences from the Dubovitskii–Milyutin framework (see [
21]). First, given a family
of convex cones that are weakly closed,
where the bar denotes closure in the
-topology. This identity reflects how separation in the dual space converts intersections into (closed) sums.
Next, if
are
open convex cones with a nonempty common part,
then the dual of their intersection decomposes additively:
In other words, under an interiority condition, dualization turns intersections into (finite) sums without requiring additional closure.
Finally, consider convex cones
with vertex at the origin and assume that
are open. Then, the geometric alternative
holds precisely when there exist functionals
for
, not all zero, such that their sum cancels:
This is a dual separation statement for cones: emptiness of the primal intersection is certified by a nontrivial balanced combination of supporting functionals from the corresponding dual cones.
3.3. Euler–Lagrange (EL) Framework
Let
be a functional and let
for
, with
for
. We consider the optimization task
Remark 2. In customary settings, the families for encode inequality-type admissibility regions, whereas represents equality-type restrictions; typically .
Theorem 3. (DM). Let be a (local) minimizer of (17) and suppose the following: - (a)
denotes the decay (descent) cone of at .
- (b)
For , the sets are the feasible direction cones attached to at the point .
- (c)
is the tangent cone to at .
If all the cones , , are convex, then there exist continuous linear functionals for , not all identically zero, such that Remark 3. The relation (18) is often referred to as the abstract Euler–Lagrange (EL) identity. Definitions of the decay/descent, feasible direction, and tangent cones can be found in [21]. Scheme to apply the Dubovitskii–Milyutin Theorem to specific problems:
- (i)
Identify the decay vectors.
- (ii)
Specify the admissible vectors.
- (iii)
Characterize the tangent vectors.
- (iv)
Construct the corresponding dual cones.
We denote by the decay/descent cone and obtain the following important result.
Theorem 4 (see [
21], p. 45).
Let be convex and continuous on a topological vector space E, and let . Then, the (one-sided) directional derivative exists in every direction at , and- (a)
- (b)
We write for the admissible (feasible direction) cone of ℧ at .
Theorem 5 (see [
21], p. 59).
If is convex with , then The collection of all angent directions to ℧ at forms a cone with vertex at the origin; we denote it by and call it the angent cone.
A key device for evaluating tangent cones is the Lyusternik principle, which reduces the computation to a linearized condition at the reference point.
Theorem 6 (Lyusternik; see [
21]).
Let be Banach spaces and suppose the following:- (a)
and is Fréchet-differentiable at ;
- (b)
The derivative is onto.
Then, for the set , the tangent cone at is A detailed proof of this Theorem can be found in [
22] (also, see [
23]).
4. Establishing the Main Result Theorem 1
Proof. Define the functional
by
and set
, where
collect the pairs
that satisfy, respectively, (
3), (
4), (
5), and (
6).
With this notation, Problem 1 can be reformulated as
- (a)
Step (a): Behavior of .
Let
denote the descent cone of
at
. By results presented in the preliminaries and in [
12,
13,
14], we have that
Whenever
, its dual cone is
Moreover, as in [
21], we have that
is given, for all
, as follows:
Consequently, for any
there exists
such that, for all
,
- (b)
Examination of the constraint
We now compute the tangent cone to
at the reference point
:
Assume, for the moment, that the following
variational integral system
is controllable. Invoking Theorem 6 (see [
12,
13,
14,
21]), one obtains
To determine
, consider the linear subspaces
Hence
. By Proposition 2.40 in [
14] (see also [
13]), we have that: a functional
belongs to
if there exists
such that
Moreover, Lemma 2.5 in [
14] ensures that
is
-closed; using basic properties of dual cones, it follows that
Therefore, a functional lies in precisely when it can be decomposed as with and .
- (c)
Handling of the constraint
With this notation, . Because is closed, convex, and has a nonempty interior, we immediately obtain the following:
Let
denote the cone of feasible directions for
at the point
. Then, the product structure carries over to cones:
where
is the feasible direction cone of
at
.
Consequently, any supporting functional
splits as
By Theorem 5, the functional acts as a support to the set at the reference control .
- (d)
Treatment of the constraint
Introduce the scalar functional
By Example 7.5 from [
21], its directional derivative at
in the direction
is
where
Since along the active pathwise constraint, we have
, this set reduces to
On the other hand, the feasible direction cone for
at
contains the descent cone:
By Theorem 4, this latter cone admits the representation
Consequently, from (
20), we obtain the inclusion of dual cones
Then, by Example 10.3 (see [
21], p. 73), we have that for all
there is a non-negative Borel measure
on
such that
and
has support in
- (e)
Euler–Lagrange relation.
Note first that each of the sets
is a convex cone. Consequently, by Theorem 3 there exist functionals
for
, not all trivial, such that
Expanding (
21) using the representations introduced earlier, we obtain
Next, given any
, there exists
solving (
19) with initial condition
. Hence
and, in particular,
. Therefore the previous identity reduces to
Let
denote the solution of the adjoint relation (
12); that is,
This is a Volterra-type integral equation of the second kind, which admits a unique solution
(see [
24], p. 519). Consequently, multiplying both sides by
and integrating with respect to
t over the interval
, we obtain
By integrating by parts in
t and exchanging the order of integration, the first contribution in (
23) can be rewritten as follows:
Similarly, the second term of (
23) can be written as follows:
Likewise, we have that for the third term
The fourth term in (
23) can be streamlined by applying the integration by parts formula for Stieltjes integrals (see [
24]), together with the assumptions
and
. In particular, since the boundary points are inactive (
and
), it follows that
.
Then, by (
24), (
25), (
26), and (
27), we have that
Then, rewriting last equality, we have
Since
satisfies the variational Equation (
19), we have that
Consequently, from the identity (
22), we deduce
for every
. Since
acts as a supporting functional of
at the point
, Example 10.5 (see [
21], p. 76) yields
for all
and for almost every
.
We now argue that the alternative
together with
cannot occur. Indeed, if
, then
. Hence
so
. From (
12) with
, it follows that
which forces
. In addition, by (
22), we also obtain
for
; plugging this into (
21) implies
, i.e.,
contradicting Theorem 3.
Up to this point, two auxiliary hypotheses have been invoked:
First, we assumed
. Second, we postulated that
is controllable.
We now show that these assumptions are dispensable. If
, by the definition of
, it follows that
Take
and
. Then, from (
29), we infer
for all
where
x solves (
19). Hence, for each
,
which implies
for all
and almost every
.
Finally, suppose (
19) is not controllable. Then, by Lemma 1, there exists a nontrivial
such that
and, for all
,
Choosing
and
, the function
satisfies (
12), and therefore
for every
and almost all
.
The proof of Theorem 1 is now complete. □
5. Sufficient Conditions for Optimality
The first-order requirement in Theorem 1 (the maximum principle) becomes a sufficient criterion under additional hypotheses. To make this explicit, we focus on the specialization of Problem 1 in which the underlying Volterra dynamics are linear.
Problem 2. Here is the state and is the control. The kernels and have sizes and , respectively, with entries of class , whereIn other words, A and B are continuously differentiable in both arguments on the triangular domain Δ. Let be any feasible pair satisfying (33)–(36). Theorem 7. Assume that the hypotheses – together with items and of Theorem 1 are satisfied. In addition, suppose the following:
- (A)
The linear Volterra system (33) is controllable. - (B)
There exists a control with for all .
- (C)
Let be the state associated with via (33). Then and for every . - (D)
The functions h, g, and ψ are convex in their first two arguments.
Then, the pair is a global solution of Problem 2.
Proof. Define the mapping
by
Set
, where
is given by (
33) and (
34),
by (
35), and
by (
36), as in Theorem 1.
With this notation, Problem 2 is equivalent to
Observe that each
(
) is convex; moreover, by assumptions (C)–(D) the functional
is convex and
.
Therefore, by Theorem 2.17 in [
14], we have the following:
minimizes over ℧ if and only if there exist (), not all vanishing, such that
Here
(
) are the approximation cones defined exactly as in Theorem 1.
Let
denote the feasible direction cone to
at
. Then
where
and
Hence, by dual-cone monotonicity,
Consequently, any
has the representation
for some non-negative Borel measure
supported on
R.
Now, assume the maximum principle of Theorem 1 holds. Then, there exist
,
, and a non-negative Borel measure
supported on
R, together with a function
solving
with
, and for every
and almost every
,
To conclude, it suffices to exhibit
(
), not all zero, such that
. Define
and the functionals
From (
38), we obtain
so
supports
at
, and therefore
. Define
by
We claim
, where
as in Theorem 1. Indeed, for
, multiply (
37) by
and integrate over
to obtain
Hence
so that
and therefore
.
Finally, define
by
Then
,
,
, and
with at least one of these functionals being nonzero because, by hypothesis,
.
The claimed global minimality of follows from the convexity assumptions. □
6. A Mathematical Model
In this part, we outline a representative real-world model where our results apply; the subsequent section states an open question.
Optimal Control for an Epidemic: The SIR Scheme
Consider a community facing an infectious disease that we aim to mitigate by vaccination. We introduce the state components:
: Number of infectious individuals capable of transmitting the disease;
: Number of susceptible (yet noninfected) individuals;
: Number of removed/recovered individuals who are no longer susceptible.
Let
denote the transmission coefficient,
the recovery rate, and
the vaccination/control input, subject to the bound
. The associated optimal control problem for the SIR dynamics is
where
.
In contrast with classical formulations using differential equations, we model the epidemic process through an integral representation. This approach naturally accounts for the cumulative nature of epidemic transitions and facilitates numerical implementation.
The functions , , and should not be interpreted merely as constant initial values. Instead, they may represent either
Pre-processed or interpolated empirical data on the susceptible, infected, and recovered subpopulations prior to the onset of control;
Accumulated trajectories obtained from previous simulation stages or observational time series;
Non-constant baseline trajectories reflecting uncertainty or memory effects.
In epidemic models such as the SIR system, various forms of time-dependent state constraints of the form
may arise to reflect practical limitations, policy objectives, or ethical considerations. Below, we describe several representative types of such constraints.
Healthcare Capacity Constraints:
- –
Limit on active infections:
- –
Maximum proportion infected:
Containment or Safety Constraints:
- –
Infected individuals less than susceptible:
- –
Infection rate constraint:
Final-Time Constraints:
- –
Minimum number of recovered individuals at final time:
- –
Near eradication at the final time:
Mixed-Variable Constraints:
- –
Combined restriction on susceptible and recovered populations:
- –
Preservation of a minimum level of susceptibles:
Cost- or Resource-Based Constraints:
- –
Indirect economic/social cost limitation:
This formulation allows greater flexibility in modeling and aligns well with the structure of optimal control theory in Banach spaces, where integral equations provide a natural framework for weak formulations and the application of variational methods.
The aim is to select a vaccination policy that minimizes the functional objective over the fixed horizon
T. With the notation introduced below, this problem can be put into the abstract framework of Theorem 1.
where
.
Therefore, the adjoint equation becomes the following equation:
Since no terminal constraint is imposed on
, we may set
. For convenience, we also normalize the multiplier to
. We now state the Pontryagin maximum condition.
From
and since
, the variational inequality reads
Then
Hence, optimal control is the projection of
onto
; i.e.,
At the terminal time, we have
. Therefore, if
, the control saturates at the upper bound near
T, i.e.,
in a neighborhood of the final time.
7. Open Problem
We close with an open question that outlines a fertile line of inquiry, one that could naturally evolve into a doctoral dissertation. The challenge concerns an optimal control setting that, at the same time, accommodates impulsive interventions and enforces pathwise constraints on the state. Our aim is to investigate the following formulation in forthcoming work.
Problem 3. Determine a control and a trajectory that locally
minimize the performance indexThis is subject to the following requirements. Volterra-type state evolution:where the kernel K satisfies suitable regularity and growth hypotheses (to be specified in the sequel). Impulsive dynamics:with prescribed impulse times and given jump maps . Admissible controls:where Ω is compact and convex. Pathwise state restriction:
8. Conclusions and Final Remarks
We have developed a fresh criterion for the controllability of linear systems driven by Volterra integral equations, obtained via a new lemma that is of standalone mathematical interest. A direct consequence is that the customary hypothesis requiring controllability of the linearized (variational) dynamics near an optimal pair turns out to be unnecessary.
On this foundation, we extended Pontryagin’s maximum principle to a wide class of optimal control problems with Volterra-type dynamics, allowing for both terminal constraints and time-dependent path constraints. Leveraging the Dubovitskii–Milyutin framework, we established
necessary optimality conditions under mild regularity. In addition,
Section 5 (Theorem 7) provides
sufficient conditions for optimality: under convexity of the cost functional and linear Volterra dynamics, the maximum principle becomes a sufficient criterion for
global optimality, recovering the classical sufficiency in the differential case (Corollary 1). Two complementary adjoint representations were derived: (i) a Volterra adjoint equation when only terminal constraints are present and (ii) an adjoint relation involving a Stieltjes integral against a non-negative Borel measure in the presence of state constraints.
When the Volterra kernel collapses to a differential operator, our statements specialize to the classical Pontryagin maximum principle, thereby unifying the integral and differential settings within a single theory. The case study on optimal vaccination in an SIR epidemic model highlights the practical reach of the results.
We also proposed an open problem that naturally follows from this work, pointing to fertile directions for future research and potential thesis-level investigations. Progress on these questions could further broaden the scope of the Dubovitskii–Milyutin approach and stimulate additional applications.