1. Introduction
It is common practice in the fields of Calculus of Variations and Optimal Control to extend the space of solutions for problems that cannot be solved in a, say, ordinary space, or if the solution is difficult to find, even with numerical approximation. This process, known as extension, involves compactifying and regularizing the problem, resulting in a more manageable structure and the possibility of obtaining necessary and sufficient conditions for optimality. However, for it to be considered a well-posed extension, it is crucial that there is no gap between the infimum of the original problem and the infimum of the extended problem. Otherwise, the extended problem will not provide any useful information about the original problem. This gap also causes issues for the method of dynamic programming, as the solution to the corresponding Hamilton-Jacobi equation typically coincides with the value function of the extended problem. However, even if the set of strict-sense solutions is -dense in the set of extended paths, the presence of constraints often leads to the occurrence of an infimum gap. In particular, this problem arises when all strict-sense solutions close to a feasible extended trajectory, for instance a local minimizer, fail to meet the constraints.
The main purpose of this paper is to establish new necessary conditions for an infimum gap to occur between problem and its extension below. The fulfillment of these conditions requires the use of necessary optimality conditions, expressed through the maximum principle, in abnormal form (i.e., with cost multiplier zero for some set of multipliers).
Fixed
and
, we consider the minimization of
subject to the dynamic constraint
and to the state and endpoint constraints
The data comprise the bounded, but not necessarily closed set of control values
, the compact set
, the closed target set
, and the functions
,
, and
(see
Section 2). We set
, and then introduce two sets of admissible controls:
(
denotes the closure of
V). We refer to any triple
as an
extended process, or simply a
process, when
are the control functions and
is the solution of (
1) associated with
. A process
is a
strict sense process if
. A strict sense or extended process
is
feasible when it satisfies the constraints (
2). Let
and
denote the subsets of feasible strict sense processes and feasible extended processes, respectively. Hence, we consider the following optimal control problems:
and
Clearly,
, so that
When this inequality is strict, one usually says that
there is an infimum gap or that
the Lavrentiev phenomenon occurs. In particular, introducing a notion of local minimizer based on control distance (see
Section 2), we distinguish:
a type-E local infimum gap, when the cost of a local minimizer of the extended problem is strictly smaller than the local infimum of the original, strict sense problem,
type-S local infimum gap, whether a local strict sense minimizer is not a local minimizer of the extended problem.
Under the assumptions specified in
Section 2, the main results of this paper can be summarized as follows:
- (i)
if at there is a type-E local infimum gap, then satisfies the maximum principle in abnormal form;
- (ii)
if is a local minimizer of , then it satisfies the same maximum principle as the extended problem. If in addition at there is a type-S local Ψ-infimum gap, then is an abnormal minimizer.
As we will illustrate in
Section 5, problem
and its extension
include the impulsive extension of a class of non-smooth, constrained optimal control problems. These problems involve unbounded original dynamics and the customary assumptions of coercivity, which prevent minimizing sequences to converge to discontinuous paths, are not invoked. We point out that consideration of impulsive systems is crucial in many applications (see, e.g. [8,12,22]). For example, instances in mechanics are situations where some state parameters
are treated as controls [9,11].
Warga was the first to study the correlation between the presence of an infimum gap and the validity of the maximum principle in abnormal form for a classical extension by relaxation. He announced the result for a type-S local infimum gap in his early paper [42], which focused on state constraint-free optimal control problems with smooth data. In the monograph [43], Warga proved the relationship between gap and abnormality for a type-E local infimum gap in optimal control problems with state constraints. His subsequent work [45] extended this result to include nonsmooth data, utilizing the concept of ’derivative containers’ introduced in [44]. Vinter and Palladino [37] proved the above mentioned correlation both in case of type-E and type-S local infimum gap for the classical extension by convex relaxation of a class of non-smooth state-constrained optimal control problems which subsume those considered by Warga, and under less restrictive hypotheses on data. Their techniques differ significantly from Warga’s, as Warga used approximating cones to reachable sets, while Vinter and Palladino utilize the non-smooth maximum principle, expressed in terms of subdifferentials, originally formulated by Clarke [13]. More recently, following the latter approach, results of this kind were established in [34], [16] for the impulsive extension of optimal control problems with unbounded dynamics (without and with state constraints, respectively), and in [17,18,19] for an abstract extension, including both relaxation and impulsive extension as special cases. In particular, in [16,17,18] we also provide, for the first time, sufficient conditions for the nondegeneracy of the abnormality condition related with a type-E infimum gap.
However, all these works focus primarily on type-E local infimum gap and consider -local minimizers (i.e. local minimizers with respect to the -distance of trajectories). Specifically, apart from Warga’s initial work, type-S local infimum gap is only studied in [37], for the extension by convexification of the dynamics, and in proceeding [19], for a more general extension. In both papers, the result is not entirely satisfactory, however, because it is shown that a strict sense -local minimizer that is not also an extended minimizer, satisfies in abnormal form an ’averaged version’ of the maximum principle, which is much less informative than the actual maximum principle.
In this paper, for the extension under consideration, on the one hand, we fill the gap that was left in the previous literature between the results obtained for type-E infimum gap and type-S infimum gap, respectively, by showing that in both cases the local minimizer is abnormal for the maximum principle associated with the extended problem. On the other hand, we extend the previous results for type-E local infimum gap to the case of a notion of local minimizer based on control distance rather than trajectory distance. Note that, by the continuity property of the input-output map associated with the control system, this implies that the present results imply the previous ones.
The paper is organized as follows. In
Section 2 we collect notation, useful definitions, and the precise assumptions. In
Section 3 we rigorously introduce the concepts of type-E and type-S local infimum gap and state our main results, proved in
Section 5.
Section 4 is devoted to apply these results to the impulsive extension of a control-affine system with unbounded controls. We also give an example.
Section 6 contains some concluding remarks.
2. Notation ands basic assumptions
2.1. Notation and preliminaries
Given and a set , we write , , for the space of absolutely continuous functions, Lebesgue integrable functions, essentially bounded functions defined on and with values in X, respectively. For all the classes of functions introduced so far, we will not specify domain and codomain when the meaning is clear and we will use , , or also , to denote the and the ess-sup norm, respectively. Furthermore, we denote by , , , the Lebesgue measure, the convex hull, the closure, and the boundary of X, respectively. As customary, is the characteristic function of X, namely if and if . Given a closed set and a point , we define the distance of z from as . For any , we write . We use to denote the space of monotone non decreasing, real valued functions on of bounded variation, vanishing at the point 0 and right continuous on . Each defines a Borel measure on , still denoted by , its total variation function is indicated by or by , and its support is spt. If , we say that if for any continuous function .
Some standard constructs from non-smooth analysis are employed in this paper. For background material we refer the reader for instance to [13,40]. A set
is a
cone if
for any
, whenever
. Take a closed set
and a point
, the
limiting normal cone of
D at
is given by
in which the notation
is used to indicate that all points in the converging sequence
lay in
D. Take a lower semicontinuous function
and a point
, the
limiting subdifferential of
G at
is
If
is a lower semicontinuous function and
, we write
,
to denote the
partial limiting subdifferential of G at w.r.t. x, y, respectively. When
G is differentiable,
is the usual gradient operator and
,
denote the partial derivatives of
G. Given a locally Lipschitz continuous function
and
, the
hybrid subdifferential of
G at
is
where
is the set of differentiability points of
G. Finally, given a locally Lipschitz continuous function
and
, we write
to denote the
Clarke generalized Jacobian, defined as
where now
denotes the classical Jacobian matrix of
G. If
and
,
,
denote the
Clarke generalized Jacobian of G at w.r.t. x, y, respectively. We recall that it holds
2.2. Basic assumptions
We shall consider the following hypotheses, in which
is a feasible extended process, which we call the
reference process and, for some
, we set
-
(H1)
The Borel set is compact and the Borel set is bounded. Moreover, there exists a sequence of closed subsets of V such that
-
(H2)
The cost function Ψ is Lipschitz continuous on a neighborhood of . The target is closed. The constraint function h is upper semicontinuous and there exists such that
-
(H3)
-
For all , is Lebesgue measurable on . Moreover, there exists such that
for all , . Furthermore, there exists some continuous increasing function with such that for any , we have
Remark 1. Condition
(H1), which is always satisfied when the set
V is relatively open, implies (and in general is stronger than) the density of
in
in the
-norm. In particular, for any
and any
there exists an integer
such that the Hausdorff distance
for every
. Hence, by the selection theorem [7, Theorem 2, p. 91]] there is a measurable function
for a.e.
t, such that
Remark 2. Condition
(H3) is satisfied, for instance, when
where
,
verify hypothesis (
4) and, in addition, the function
is
and
is continuous on the compact set
. Another situation where condition
(H3) is verified, is when the dynamics function has a polynomial dependence on the control variable
w, with locally Lipschitz continuous coefficients in the state variable.
3. Type-E or type-S local infimum gap and abnormality
3.1. Type-E and type-S local infimum gap
We recall that
and
denote the sets of feasible strict sense and feasible extended processes, respectively. Given
,
, we define the following distance:
Definition 1 (Local minimizer).
Let and denote and or and , respectively. A process is called a local
-minimizer
for problem if, for some , one has The process is a-minimizer
for problem if .
Remark 3. Under hypothesis
(H3), for each extended control
in a suitable
-neighborhood of the reference control
, there is one and only one solution
of (
1). Furthermore, the input-output map
from
to
is continuous in this neighborhood, provided
is equipped with the distance
and
with the distance induced by the sup-norm. Consequently, if the process
is an
-local minimizer, meaning that
reaches the minimum over processes
with
for some
, then it is also a local minimizer according to Def. 1. In general, the contrary is not true.
It is now natural to introduce two notions of local infimum gap, depending on whether the reference process is extended or strict sense.
Definition 2 (Infimum gaps).
Let be a continuous function. (i)
If and there is some such that1
we say that at
there is a type-E local
-infimum gap.(ii)
Let be a local Ψ-minimizer for problem which is not a local Ψ-minimizer for problem , i.e. for any there exists some such that
Then, we say that at
there is a type-S local
-infimum gap.
(iii)We say that there is a-infimum gap if
When is clear from the context, we will often simply write infimum gap instead of -infimum gap.
Remark 4. As it is easy to see, thanks to the continuity of the input-output map
the notion of type-E local
-infimum gap at
is actually independent of the cost function
, as it is equivalent to the fact that
(see [17, Proposition 2.1]). If
satisfies (
6), we say that it is an
isolated process.
3.2. Main results
Now we introduce a Pontryagin maximum principle and a notion of normal and abnormal extremal for the extended optimization problem. Then we establish a link between abnormality and occurrence of a gap phenomenon.
Definition 3 (Pontryagin maximum principle).
Let be a feasible extended process for problem and let hypotheses (H1)-(H2)-(H3) be satisfied. We say that is a-extremal,
or satisfies the Pontryagin maximum principle,
if there exist a path , , and a Borel measurable and μ-integrable function satisfying the following conditions:
where
We will call a Ψ-extremal normal
if all possible choices of as above have , and abnormal
when it is not normal. Since the notion of abnormal Ψ-extremal is actually independent of Ψ, in the following abnormal Ψ-extremals will be simply called abnormal extremals.
Theorem 1.
Let and let hypotheses (H1)-(H2)-(H3)be satisfied. Then,
(i) if is a local Ψ-minimizer for , then is a Ψ-extremal. If at there is a type-E local Ψ-infimum gap, then is an abnormal extremal; (ii) if is a local Ψ-minimizer for , then is a Ψ-extremal. If at there is a type-S local Ψ-infimum gap, then is an abnormal extremal.
The proof of Theorem 1, in which the notion of local minimizer adopted in this work, based on the control distance
, plays a crucial role, is given in
Section 5. The main novelty of Theorem 1 is statement (ii), concerning the case where
is a local minimizer of the original problem which is not a local minimizer of the extended one. Indeed, in the previous literature (see [19,37]) it was proved in this case that
is an abnormal extremal for an ‘averaged version’ of the maximum principle only, meaning that the adjoint equation (8) was replaced by the following weaker differential inclusion
in which all information on optimal control is lost.
Remark 5. It is worth mentioning that, despite hypothesis (H1) implies the density of in in the -norm, it is well-known since the earliest work by Warga [45] and Kaskovz [24] that, in general, if the set of strict sense controls is merely an -dense subset of the set of extended controls, the link between gap and abnormality established in Theorem 1 may fail (see for instance the example in [36]).
As a direct consequence of Theorem 1, we obtain that normality is a sufficient condition for the absence of any type of local infimum gap.
Theorem 2.
Let and let hypotheses (H1)-(H2)-(H3)be satisfied. Then, (i) if is a local Ψ-minimizer for which is a normal Ψ-extremal, at there is no type-E local Ψ-infimum gap. If, in addition, is a Ψ-minimizer for , then there is no Ψ-infimum gap; (ii) if is a local Ψ-minimizer for which is a normal Ψ-extremal, at there is no type-S local Ψ-infimum gap, namely, is a local Ψ-minimizer for as well.
4. An application: the impulsive extension
4.1. An impulsive optimal control problem
Consider the following free end-time optimal control problem with
unbounded, control-affine dynamics:
in which
,
,
,
for any
,
, and
. We make the following assumptions on data:
-
(H4)
is a fixed constant possibly equal to , the (unbounded) set of control values U is a closed cone, the target is a closed set, the dynamics functions f, , the constraint function h, and the cost function Ψ are locally Lipschitz continuous.
Notice that
(sometimes called
fuel or
energy) coincides with the
-norm of the control function
u on
. Assuming, as usual, the function
merely monotone nondecreasing (see e.g. [32]), this problem is non-coercive, i.e. there are no conditions that prevent a minimizing sequence of trajectories from having increasing velocities and converging to a discontinuous path. Hence, adopting a by now standard extension, we embed the original problem into the
space-time or
extended problem
below, where the extended state variable is
, and extended trajectories are
-paths which are (reparameterized)
-limits of graphs of the original trajectories [10,27,30,38,41]:
2
where
, being
W the control set given by
Notice that, with any process
of the original problem
, by setting
through the time-change
we can associate a process
for
with
a.e.. In particular, problem
can be identified with the restriction of problem
to the set of processes with
a.e.. In the following, we will refer to such restriction as
strict sense problem and to such processes as
strict sense processes.
Therefore, the extension consists in considering extended processes where may be zero on nondegenerate subintervals of . On these intervals, the time variable is constant, while the state variable y evolves according to the ‘fast’ dynamics . This explains why is also called the impulsive extension of problem , although it is a conventional optimization problem with bounded controls. In fact, one could give an equivalent s-based description of this extension using bounded variation trajectories and controls [2,5,26,29,31,33,39,46].
Adopting terminology of the present paper, we say that an extended or strict sense process
is
feasible [resp., an original process
is
feasible] if it satisfies all constraints of problem
[resp.,
]. The sets of feasible original, feasible extended and feasible strict sense processes are denoted by
,
and
, respectively. Given
and
, we define the distance:
3 At this point, the definitions of local minimizer and of type-E and type-S local
-infimum gap (see Def. 1 and Def. 2) can be easily adapted to the impulsive extension by replacing the distance
defined in (
5) with the distance
given by (
13). The unmaximized Hamiltonian associated with problem
above is given by
for all
.
Definition 4.
We say that is a-extremal
if there exist a path , , , and Borel-measurable and μ-integrable functions satisfying the following conditions:
where is given by
Moreover, if and , then . Furthermore, if , then (14) can be strengthened with
We say that a Ψ-extremal is normal
if all sets of multipliers as above have , and abnormal
when it is not normal.
From Theorem 1 we can deduce the following results.
Theorem 3.
Let and assume hypothesis (H4). Then,
(i) if is a local Ψ-minimizer for , then is a Ψ-extremal. If at there is a type-E local Ψ-infimum gap, then is an abnormal extremal; (ii) if is a local Ψ-minimizer for , then is a Ψ-extremal. If at there is a type-S local Ψ-infimum gap, then is an abnormal extremal.
Proof. The impulsive extended problem
has a free end-time, so the theory developed in the previous sections for fixed end-time problems does not apply straightforwardly. However, through a standard time rescaling procedure that applies to free end-time problems with Lipschitz continuous time dependence, we can embed problem
into a fixed end-time optimization problem, satisfying all the assumptions of Theorem 1 and for which, for example,
is still a local minimizer if it was so for
. Precisely, let
,
and consider the
rescaled problem:
where, for any
, we have set
We refer to any element
satisfying all constraints in
as a feasible rescaled extended process. If
a.e., then
is called a feasible rescaled strict sense process. For any pair of feasible rescaled extended processes
,
we define the distance
Let us associate with the given reference process
, the (feasible) rescaled process
. From a straightforward application of the chain rule and standard calculations it follows that for any
there exists some
such that with each feasible rescaled extended process
with
, using the time-change
we can associate the following feasible extended process
satisfying
. Moreover,
.
As a consequence, if is a local -minimizer for for some , then is a local -minimizer for , at which there is a type-E local infimum gap as soon as at there is a type-E local infimum gap. At this point, the proof of Theorem 3 can be derived applying Theorem 1 to the rescaled problem. We omit the details, which follow the same line as the proofs of [40] and [17]. □
Remark 6. With similar arguments as in [17], what we have done in this section can be easily generalized to control-polynomial impulsive problems, by which we mean that the dynamics of the original problem
can be replaced by
where
d is an integer
. This generalization may be relevant for some applications to Lagrangian Mechanics, where dynamics are usually control-polynomial with degree
(see [11]).
4.2. An example
The following example tells us that both a type-S local infimum gap and a type-E local infimum gap may occur. Moreover, we exhibit sets of abnormal multipliers, which exists in accordance with Theorem 3. Consider the optimization problem with scalar, unbounded controls:
Let
, then the space-time extension of the above problem is given by
Type-S local infimum gap. Let
be the following strict sense process, where
, the control
is given by the constant pair
and
It is easy to see that
, which corresponds to the process of
associated with the control
, is trivially a strict sense minimizer, as
is the unique feasible strict sense trajectory. However,
is not a local minimizer for the extend problem
. Indeed, let us fix
sufficiently small and let us consider the extended process
where
and
is given by
so that one has
For any
, this is the description in the state-space of a discontinuous state trajectory
for problem
which first reaches the point
using the control
and then jumps to the position
with an impulse. Notice that
is a feasible extended process that satisfies
and whose cost is strictly less that the cost corresponding to
, because it holds
Thus, by the arbitrariness of
, at
there is a type-S local infimum gap. Indeed, a set of abnormal multipliers corresponding to
is given by
, where
,
,
,
,
and
for any
.
Type-E local infimum gap. Consider now the following extended process
, where
and
is given by
so that one has
It is easy to see that
is a minimizer for
, as it is feasible and its corresponding cost is equal to zero. Moreover, at
there is type-E local infimum gap, since
defined in the previous step is the unique feasible strict sense process. Indeed, a set of abnormal multipliers corresponding to
is given by
, where
,
,
,
,
,
,
and
for any
.
5. Proof of Theorem 1
Preliminarily, let us observe that, since the proofs of statements (i)-(ii) involve only extended processes with trajectories close to the reference trajectory
and the controls take values in compact sets, using standard cut-off techniques we can assume that hypotheses
(H2) and
(H3) are satisfied in the whole space
. Therefore, the input-output map
associated with (
1) is well-defined and continuous (actually, uniformly continuous).
5.1. Proof of statement (i)
If is a local -minimizer for , the fact that it satisfies the Pontryagin maximum principle in Def. 3 can be easily derived by [40]. The proof that whenever at there is a type-E local infimum gap, then it is an abnormal extremal, requires instead a careful adaptation of the arguments used in the proof of [17], where the same result is obtained for a notion of type-E local infimum gap in which the distance between controls is replaced by the distance of the trajectories. The proof is divided into several steps in which successive sequences of optimization problems are introduced that have as admissible controls the strict sense ones, and costs that measure how much a process violates the constraints. Thanks to the Ekeland principle, for these problems it is possible to find a sequence of minimizers which converge to the reference process . Furthermore, applying a maximum principle to these approximating problems with reference to the above mentioned minimizers, we obtain in the limit a set of multipliers with for the extended problem with reference to .
Step 1. Define the function
, given by
and for any
, introduce the payoff
Fix a sequence
satisfying
and let
be such that
By the uniform continuity of the input-output map and the Lipschitz continuity of
, it follows that
. Moreover,
for every
i large enough, since
is an isolated process in view of Remark 4.
According to hypothesis
(H1) and Remark 1, for any
i there exist an element of the sequence
, which we denote by
, and some
such that
. Hence, let
be such that
and
. As a consequence,
is a
-minimizer for the optimization problem (
) given by
where
It is an easy task to show that, if we equip
with the distance
, then it turns out to be a complete metric space. Accordingly, in view of the Ekeland’s variational principle, there exists
which is a minimizer for problem
given by
where
is defined as
Moreover, one has
, so that
. In particular, it holds
Furthermore, in view of the continuity of the input-output map associated with control system (
1), one has
By the previous convergence analysis and since
is isolated, it follows that
for any
i. Therefore, possibly passing to a subsequence, for any
i we have
Step 2. From the above reasonings it follows that
is a minimizer for the optimal control problem
, given by
Possibly passing to a subsequence, only one of the following two cases occurs:
Let us first analyze
Case (a). Since in this case
implies
, one has
. Moreover, in view of the max rule for subdifferentials (see e.g. [40]), if
, then there exist
,
such that
,
and
. Furthermore,
for
whenever
is strictly greater than the
k-th term in the maximization. Thanks to the above reasonings and applying the maximum principle to problem
with reference to its minimizer
we deduce that there exist
,
,
,
,
such that
, and a Borel-measurable and
-integrable map
satisfying conditions (i)′–(vi)′ below:
- (i)′
;
- (ii)′
and for a.e. ;
- (iii)′
, , ;
- (iv)′
-a.e. ;
- (v)′
spt;
- (vi)′
-
l
l
for any ,
where
is the diameter of the compact set
and
is defined as
From (ii)′ and (iii)′ we deduce that
and
. Since
, from (iii)′ we also have
. By summing up these relations and (i)′ we get
so that
. By rescaling the multipliers, one obtains
and
.
If instead
Case (b) occurs, then
for any
i in view of (
23). Hence, for
small, the process
still is a minimizer for
and
for all
. By applying the maximum principle to
with reference to
, we deduce the existence of
and
4 satisfying conditions (i)′–(vi)′ above for
,
(hence,
). In this case, by (iii)′ we deduce
. By summing up this relation with (i)′ we get
, so that
. By rescaling the multipliers, we have
and
.
Step 3. For both Case (a) and Case (b) we have proved that for any i there exist , and a Borel-measurable and -integrable map satisfying relations (i)–(vi) below
- (i)
;
- (ii)
a.e. ;
- (iii)
;
- (iv)
-a.e. ;
- (v)
spt;
- (vi)
-
l
for any ,
where
is as in (
24). Employing a standard convergence analysis (see [16] for more details) we deduce that there exist
and a Borel-measurable and
-integrable map
such that, up to a subsequence, we have
Therefore, using (
22) and passing to the limit in conditions (i), (iv) and (v) we obtain
Moreover, using basic properties of subdifferentials and the fact that
for any
(see [40]), passing to the limit in (iii) we deduce that
where
is given by
Let us now derive the adjoint equation (8). Let
, so that
in view of (
21). Using (
3) and hypothesis
(H3), for a.e.
, we get
where, since
, the map
is given by
By the continuity of
, (
21) and (
25) we deduce that, up to a subsequence,
for a.e.
. Moreoveor, it holds
Hence, by the dominated convergence theorem,
in
(in particular,
in
). From the compactness of trajectories theorem (see [40, Theorem 2.5.3]) it follows that for a.e.
it holds
Now it remains to prove (10). Let
and, as a consequence of hypothesis
(H1), let
be such that
for any
i and
. Condition (vi) implies that
Up to a subsequence, the right hand side of the above inequality converges to
, by the dominated convergence theorem. At the same time it holds
But now the second term in the right hand side of the equality above tends to zero in view of the dominated convergence theorem, while the third one converges to zero because of (
22) and the fact that
. Therefore, we have proved that for any
one has
From a measurable selection theorem (10) immediately follows.
5.2. Proof of statement (ii)
Let be a local -minimizer for . We can derive that it is an extremal of the Pontryagin maximum principle from ([40] Theorem 9.3.1). In particular, the maximality condition (10) still holds with the maximum taken over , since we assume that the dynamics function is continuous with respect to the w-variable.
If
is a local
-minimizer for
which is not a local
-minimizer for
, then, on the one hand, there exists
such that
for any
such that
. On the other hand, taken
with
, for each
i there exists some
such that
and
. Hence, for any
such that
, one has
, so that we have by construction
Since the strict sense process
z is arbitrary, this proves that at
there is a type-E local infimum gap for any
i. Hence, in view of Theorem 1, (i) for any
i there exist
,
and a Borel-measurable and
-integrable map
satisfying conditions (i)–(vi) below:
- (i)
;
- (ii)
a.e. ;
- (iii)
;
- (iv)
-a.e. ;
- (v)
spt;
- (vi)
a.e. t,
where
is as in (
24). We do observe that our construction implies
, so that (
21) and (
22) hold true. We can thus conclude the proof employing a standard convergence analysis similar to that in the
Step 3 of the proof of Theorem 1, (i).
6. Concluding remarks
In this paper we investigate infimum gap phenomena that may occur when we pass from an optimal control problem with non-smooth data, endpoints, and state constraints, to an extended version of it, in a framework that includes the impulsive extension of a class of non-coercive problems with unbounded dynamics. In particular, we consider a type-E and a type-S local infimum gap: in the former an extended minimizer has cost which is strictly smaller than the infimum cost over close feasible strict sense processes, in the latter a local strict sense minimizer does not locally minimize the extended problem. Following on from Warga’s previous research, but utilizing more recent perturbation techniques from non-smooth analysis, we prove that whenever at a process there is either a type-E or a type-S local infimum gap for a notion of local minimizer based on the control distance , then it satisfies a non-smooth, constrained version of the Pontryagin maximum principle in abnormal form. Compared to previous results, in which there was an ’asymmetry’ between the necessary abnormality conditions derived for type-E and type-S local infimum gap, for the extension under consideration we obtain the same condition for both.
As a corollary, we provide sufficient conditions in the form of a normality test for the absence of local infimum gap phenomena. Although a normality test for gap avoidance might seem completely theoretical and hardly verifiable, it can actually be very useful because in certain situations normality follows from easily verifiable criteria. These criteria take the form of constraint and endpoint qualification conditions for normality and have been extensively explored in the literature (see e.g. [6,14,15,25] and references therein). As shown in [16,33,34], where several explicit conditions for normality in control-affine impulsive extensions are presented, these criteria are generally weaker than those previously established to directly determine the absence of a gap, as in [1,28].
The framework introduced in this paper may have implications for future infimum gap research in several directions. On the one hand, it may be the starting point for some generalizations, such as, for instance: (i) determine a higher-order maximum principle also for local minimizers of the strict sense problem and prove that in the case of a type-S local infimum gap, there is abnormality of the higher-order conditions as well. So far, results of this kind are only known for extended minimizers and for type-E infimum gap, limited to the impulsive extension case (see [3,4,35]); (ii) explore infimum gap phenomena for the impulsive extension of optimal control problems involving control-affine systems with time delays, for which necessary optimality conditions have very recently been established by Fusco, Motta, and Vinter in [20,21].
Another interesting problem might be to consider different extension procedures for classes of control systems not considered in this paper (such as distributed parameters systems or multistage problems).
Funding
This research was funded by the INdAM-GNAMPA Project 2023, CUP E53C22001930001, and by PRIN 2022, Prot. 2022238YY5, CUP C53D23002370006
Conflicts of Interest
The authors declare no conflicts of interest.
References
- Aronna M.S., Motta M., Rampazzo F., Infimum gaps for limit solutions. Set-Valued Var. Anal. 2015, 23, no. 1, 3–22.
- Aronna M.S., Rampazzo F., L1 limit solutions for control systems, J. Differential Equations 2015, 258, 954–979.
- Aronna M.S., Motta M., Rampazzo F., Necessary conditions involving Lie brackets for impulsive optimal control problems. In Proceedings of the 58th IEEE Conference on Decision and Control (CDC) December 11-13, 2019, Nice, France, 1474-1479.
- M.S. Aronna, M. Motta, F. Rampazzo, A Higher-Order maximum principle for Impulsive Optimal Control Problems, SIAM J. Control Optim. 2020, 58(2), 814–844.
- Arutyunov A.V., Karamzin D.Y., Pereira F.L., State constraints in impulsive control problems: Gamkrelidze-like conditions of optimality, J. Optim. Theory Appl. (2015), 166, no. 2, 440–459.
- Arutyunov A.V., Karamzin D.Y., A survey on regularity conditions for state-constrained optimal control problems and the non-degenerate maximum principle, J. Optim. Theory Appl. (2020), 184, no. 3, 697–723.
- Aubin, J.-P., Cellina, A., Differential inclusions. Set-valued maps and viability theory. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 264. Springer-Verlag, Berlin, 1984.
- Azimov D., Bishop R., New trends in astrodynamics and applications: Optimal trajectories for space guidance, Ann. New York Acad. Sci. 2005, 1065(1), 189–209.
- [Bressan(1991)] Bressan Aldo Hyper-impulsive motions and controllizable coordinates for Lagrangean systems, Atti Accad. Naz. Lincei, Memorie, 1991, Serie VIII, Vol. XIX, 197–246.
- Bressan A., Rampazzo F., On differential systems with vector-valued impulsive controls, Boll. Un. Mat. Ital. B, 1988, (7) 2, no. 3, 641–656.
- Bressan A., Rampazzo F., Moving constraints as stabilizing controls in classical mechanics, Arch. Ration. Mech. Anal., 2010, 196, 97–141.
- Catllá A., Schaeffer D., Witelski T., Monson E., Lin A., On spiking models for synaptic activity and impulsive differential equations, SIAM Rev., 2008, 50(3), 553–569.
- Clarke F.H., Optimization and Nonsmooth Analysis, Wiley-Interscience, New York, 1983, reprinted as vol. 5 of Classics in Applied Mathematics, SIAM, Philadelphia, 1990.
- Fontes F.A.C.C., Frankowska H., Normality and nondegeneracy for optimal control problems with state contraints, J. Opt. Theory. Appl., 2015, 166, no. 1, 115–136.
- Frankowska H., Tonon D., Inward pointing trajectories, normality of the maximum principle and the non occurrence of the Lavrentieff phenomenon in optimal control under state constraints, Journal of Convex Analysis, 2013, Vol. 20, No. 4, 1147-1180.
- Fusco G. and Motta M., No Infimum Gap and Normality in Optimal Impulsive Control Under State Constraints. Set-Valued Var. Anal., 2021, 29, no. 2, 519–550.
- Fusco G. and Motta M., Nondegenerate abnormality, controllability, and gap phenomena in optimal control with state constraints, SIAM J. Control Optim., 2022, 60, no. 1, 280–309.
- Fusco G. and Motta M., Gap phenomena and controllability in free end-time problems with active state constraints, J. Math. Anal. Appl., 2022, 510 no. 2, Paper No. 126021, 25 pp.
- Fusco G. and Motta M., Strict sense minimizers which are relaxed extended minimizers in general optimal control problems, In Proceedings of the 60th IEEE Conference on Decision and Control (CDC) December 13-15, 2021. Austin, Texas.
- Fusco G. and Motta M., Impulsive optimal control problems with time delays in the drift term, submitted,http://arxiv.org/abs/2307.12806.
- Fusco G., Motta M., Vinter R., Optimal impulsive control for time delay systems, submitted,http://arxiv.org/abs/2402.11591.
- P. Gajardo, H. Ramirez C., A. Rapaport, Minimal time sequential batch reactors with bounded and impulse controls for one or more species, SIAM J. Control Optim., 2008, 47(6), 2827–2856.
- Hájec O., Book review: Differential systems involving impulses, Bull. Amer. Math. Soc., 12, 1985, pp. 272–279.
- Kaśkosz, B., Extremality, controllability, and abundant subsets of generalized control systems, J. Optim. Theory Appl., 1999, 101, no. 1, 73–108.
- Lopes S.O., Fontes F.A.C.C., de Pinho M.d.R., On constraint qualifications for nondegenerate necessary conditions of optimality applied to optimal control problems, Discrete Contin. Dyn. Syst., 2011, 29, no. 2, 559–575.
- Karamzin D.Y., de Oliveira V.A., Pereira F.L., Silva G.N., On the properness of an impulsive control extension of dynamic optimization problems, ESAIM Control Optim. Calc. Var., 2015, 21, no. 3, 857–875.
- Miller B.M., The method of discontinuous time substitution in problems of the optimal control of impulse and discrete-continuous systems. (Russian) Avtomat. i Telemekh., 1993, no. 12, 3–32; Translation in Automat. Remote Control, 1994, 54, no. 12, part 1, 1727–1750.
- Motta M., Minimum time problem with impulsive and ordinary controls, Discrete Contin. Dyn. Syst., 2018, 38, no. 11, 5781–5809.
- Miller B.M., Rubinovich E. Y., Impulsive control in continuous and discrete-continuous systems. Kluwer Academic/Plenum Publishers, 2003, New York.
- Motta M., Rampazzo F., Space-time trajectories of non linear systems driven by ordinary and impulsive controls. Differ. Int. Eq., 1995, 8, 269–288.
- Motta M., Sartori C., On L1 limit solutions in impulsive control, Discrete Contin. Dyn. Syst. Ser. S, 2018, 11, 1201–1218.
- Motta, M., Sartori, C., On asymptotic exit-time control problems lacking coercivity, ESAIM Control Optim. Calc. Var., 2014, 20, no. 4, 957–982.
- Motta M., Sartori C., Normality and nondegeneracy of the maximum principle in optimal impulsive control under state constraints, Journal of Optimization Theory and Applications, 2020, Vol. 185, 44–71.
- Motta M., Rampazzo F., Vinter R.B., Normality and gap phenomena in optimal unbounded control, ESAIM: Control, Optimisation and Calculus of Variations, 2018, 24, no. 4, 1645–1673.
- Motta M., Palladino M., Rampazzo F., Unbounded Control, Infimum Gaps, and Higher Order Normality, SIAM J. Control Optim., 2022, 60, no. 3, 1436–1462.
- Palladino, M. and Rampazzo, F., A geometrically based criterion to avoid infimum gaps in optimal control, J. Differential Equations, 2020, 269, no. 11, 10107–10142.
- Palladino M., Vinter R.B., When are minimizing controls also minimizing extended controls?, Discrete Continuous Dynamical System, 2015, 35(9), 4573-4592.
- Rishel R.W., An extended Pontryagin principle for control systems whose control laws contain measures, SIAM Journal of Control, 1965, 3, no. 2, 191–205.
- A. Sarychev, Nonlinear systems with impulsive and generalized function controls, in Nonlinear Synthesis, Sopron, 1989, in: Progr. Systems Control Theory, vol. 9, Birkhäuser Boston, Boston, MA, 1991, 244–257.
- Vinter R.B., Optimal control. Birkhäuser, Boston, 2000.
- Warga J., Variational problems with unbounded controls, J. Soc. Indust. Appl. Math. Ser. A Control, 1965, 3, 424–438.
- Warga J., Normal Control Problems have no Minimizing Strictly Original Solutions, Bulletin of the Amer. Math. Soc., 1971, 77, 4, 625–628.
- Warga J., Optimal Control of Differential and Functional Equations, Academic Press, New York, 1972.
- Warga, J., Optimization and controllability without differentiability assumptions, SIAM J. Control and Optimization, 1983, 21, 837–855.
- Warga, J., Controllability, extremality, and abnormality in nonsmooth optimal control, J. Optim. Theory Appl., 1983, 41, no. 1, 239–260.
- Wolenski P., Žabić S., A sampling method and approximation results for impulsive systems, SIAM J. Control Optim., 2007, 46 (3, 983–998.
1 |
As customary, when the set is empty we define the infimum equal to . |
2 |
As it is well-known, a distributional approach, where u is replaced by a Radon measure, does not work unless and the Lie brackets for every (see e.g. [10,23]). |
3 |
Notice that is equivalent to the distance obtained replacing with in the -norm (possibly extending the controls to constantly equal to 0), as for some constant . |
4 |
If it were , then , so that the linearity of the adjoint equation (ii)′ implies , contradicting (i)′. |
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2024 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).