Processing math: 100%
Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2024, Volume 215, Issue 6, Pages 790–822
DOI: https://doi.org/10.4213/sm9967e
(Mi sm9967)
 

This article is cited in 1 scientific paper (total in 1 paper)

Exact formulae for the increment of the objective functional and necessary optimality conditions, alternative to Pontryagin's maximum principle

N. I. Pogodaev, M. V. Staritsyn

Matrosov Institute for System Dynamics and Control Theory of Siberian Branch of Russian Academy of Sciences, Irkutsk, Russia
References:
Abstract: The paper presents elements of the theory of local extremum in the problem of optimal control with free right end and, in general, uncertain initial position of trajectories on the basis of exact formulae for the increment (variations of infinite order) of the objective functional. Necessary conditions for optimality of ‘feedback’ type are obtained: their formulations involve auxiliary feedback controls which generate program descent controls (in the minimum problem). The conditions proposed in this work provide an alternative to the classical Pontryagin principle (and even improve it in some special cases) and open the way to constructing indirect methods for local search without procedures for adjustment of the parameters of ‘descent depth’.
Bibliography: 26 titles.
Keywords: optimal control, exact formulae for the increment of the objective functional, feedback necessary conditions of optimality, Pontryagin's maximum principle, continuity equation.
Funding agency Grant number
Russian Science Foundation 23-21-00161
This work was supported by the Russian Science Foundation under grant no. 23-21-00161, https://rscf.ru/en/project/23-21-00161/.
Received: 13.06.2023 and 01.04.2024
Bibliographic databases:
Document Type: Article
MSC: Primary 49K15; Secondary 49K45, 49N35
Language: English
Original paper language: Russian

§ 1. Introduction

This work elaborates the subject of [1]–[3], which is asymmetric to Krotov’s sufficient optimality conditions (see [4]); it is devoted to so-called feedback necessary optimality conditions for program controls and to the corresponding methods of descent in the classical problem of dynamic optimization

(P)I[u](x[u](1))min,
ddtx(t)˙x(t)=f(x(t),u(t)),x(0)=yRn,
uUL(I;U),URm,
and in one of its qualitative generalizations, namely, the problem of control for an ensemble of trajectories (§ 3). The first condition in (1.1) is assumed to hold almost everywhere (in the sense of the Lebesgue measure) on the interval I[0,1]; we let x=x[u] denote the Carathéodory solution of the Cauchy problem (1.1) which corresponds to the control u.

Assume that a given process ¯σ=(¯x,¯u), ¯uU, ¯x=x[¯u], has to be examined for optimality. The conventional necessary conditions for a strong (as well as a Pontryagin or an L1-) extremum (see [5] and [6]) are based on classes of needle-shaped and weak variations of the control ¯u, which generate processes σ=(x,u) with the property of potential descent with respect to the functional

I[u](x(1))(¯x(1))I[¯u].
We adopt the terminology of [1] and say that such variations reject ¯u, and we refer to the corresponding processes as comparison processes.

In feedback (feedback) necessary conditions the role of rejecting variations is played by feedback controls (and this is already indicated by the term ‘feedback’). These variations widen the set of comparison processes by augmenting it with some sliding modes, namely, admissible process for the convexified problem1 (coP). Trajectories admitted to compare with ¯x are curves (in general, inadmissible in (P)) generated by feedback controls

w(t,x)argminυUx¯φt(x)f(x,υ),
where ¯φ:I×RnR is a sufficiently regular weakly nonincreasing2 solution of the boundary problem for the Hamilton–Jacobi inequality
tφt(x)+minυUxφt(x)f(x,υ)0,φ1=,
associated with ¯σ in one way or another (in the general case the boundary condition can contain an arbitrary majorant of the function ); x is the gradient operator with respect to the variable xRn, and the dot denotes the scalar product of two vectors.

Once ¯φ has been defined, the strategy (1.4) of control for the system that delivers a minimum to (1.5) gives us a formal structure of control for descent from ¯σ, like in dynamic programming. Let us put aside for a while the (nontrivial) aspect of the implementation of this strategy and discuss the problem of the selection of the function ¯φ, which was called in [1] the majorant of the objective functional at the point ¯σ. Among all solutions of inequality (1.5) it is reasonable to select one of those for which the construction (1.4) provides the deepest descent. It should be noted that in general conditions (1.5) themselves do not guarantee any descent at all: it can occur (see Example 2 in § 8.3) that for badly chosen ¯φ all curves x=x[w] synthesized by the rule (1.4) are definitely ‘worse’ than ¯x, that is, (x(1))>(¯x(1)).

The problem of the construction of adequate majorants in the nonlinear problem (P) was qualified in [3] as an open problem of the theory.3 It is a common practice to consider linear or quadratic (in the state variable) functions generated by constructions of Pontryagin’s maximum principle (PMP) and second-order conditions (solutions of the conjugate system), namely, matrix impulses of Gabasov and Riccati type. These functions generate indeed controls for descent in the corresponding particular classes of problems — linear and linear-quadratic in the state variable — however, in the general case their use has heuristic nature.

In this work we propose a universal (and fairly simple) class of nonlinear majorants in problem (P): the sought-for majorant is the function ¯p:I×RnR which takes a constant value along the flow ¯X1,t of the vector field (t,x)f(x,¯u(t)) on the interval [t,1] and coincides with at the terminal instant. In other words, the majorant is defined by the relation ¯pt(x)(¯X1,t(x)) for all (t,x)I×Rn.

Note that under the standard regularity assumptions for the vector field f the mapping (t,x)¯X1,t(x), as well as the majorant ¯p, is absolute continuous in t and continuously differentiable in x. In particular, for each xRn we have the equalities (see § 3.1)

t¯pt(x)+x¯pt(x)f(x,¯u(t))=0for almost all tI;p1(x)=(x).
Below we interpret the transport equation in this ‘pointwise’ sense. Note at the same time that the function ¯p is also the unique weak solution (that is, a solution in the sense of distributions) of problem (1.6); see [7].

The majorant constructed above possesses all required properties. First, it is weakly monotone, namely, ‘weakly constant’. Second, ¯p satisfies the Hamilton–Jacobi inequality (1.5):

t¯pt(x)+minυUx¯pt(x)f(x,υ)0t¯pt(x)+x¯pt(x)f(x,¯u(t))
at all points of I×Rn where the derivative t¯pt(x) exists. Finally, the corresponding feedback strategies
w(t,x)argminυUx¯pt(x)f(x,υ)
generate (at least one) trajectory (of a sliding mode) satisfying (1.3). We establish this result in § 7. This fact is obvious under the assumption that some implementation of the synthesis of w gives a program control uU and the corresponding solution x=x[u] for which
x¯pt(x(t))f(x(t),u(t))=minυUx¯pt(x(t))f(x(t),υ)for a.e. tI.
Then (d/dt)¯pt(x(t))0 and (x(1))=¯p1(x(1))¯p0(y)=¯p1(¯x(1))=(¯x(1)). Here the first and last equalities follow from the boundary condition (1.6), the inequality is due to the nonincreasing behaviour of the composition ¯px, and the second equality is a consequence of the fact that ¯p is constant along ¯x.

Below we will see that the function ¯p is a solution of the boundary problem (1.5) for which the cost of the process synthesized by the rule (1.4) is certainly at most I[¯u]=(¯x(1)). This will be established as a consequence of an exact formula for the increment of the functional, which generalizes such formulae for the linear and linear-quadratic settings (see [8]).

1.1. The main notation and assumptions

In this work we use the following notation:

By P(Rs) we denote the set of all probability measures on Rs and by Pc(Rn) the subset of measures with compact support. The latter set can be endowed with the structure of a metric space by equipping it with the Kantorovich p-metric Wp, p1 (see [9]) (this space is not complete, but it is dense in any complete space (Pp(Rn),Wp), where Pp(Rn) is the subset of P(Rn) consisting of all measures with finite pth moment: see [10]). It is always assumed below that Pc(Rn) denotes the metric space (Pc(Rn),W1) without further qualification.

We distinguish two Borel measures on Rs, namely, Ls is the classical Lebesgue measure and δx is the atomic Dirac measure concentrated at a point xRs. The abbreviation ‘a.e.’ stands for ‘almost everywhere’ or for ‘almost all’ points of the corresponding set with respect to the measure indicated. When the measure is not indicated, it is assumed to be the one-dimensional Lebesgue measure L1.

Given a norm || on Rn, we denote the matrix norm consistent with it by the same symbol. Finally, DxV is the matrix of partial derivatives (xiVj) of the vector field V:RnRn, where the Vj are components of V.

In this work the following basic assumptions are supposed to hold.

Remark 1. We restrict our considerations to the autonomous case, since the approach elaborated here involves (in § 6) elements of the theory developed by Krasovskii and Subbotin [11], in which the continuity of the vector field with respect to the variable t plays an essential role. At the same time the case when the dependence mentioned above is sufficiently regular reduces to the autonomous case by extending the phase space.

§ 2. Variational approach. Bilinear problem

The method of the derivation of feedback necessary conditions described above does not emphasize their variational character. However, the condition corresponding to the feedback variation (1.7) can be derived in a classical way, by considering the increment of the functional. On the whole this approach reproduces the standard algorithm for deriving the PMP; it is most demonstrative in the case of a bilinear problem.

For the sake of simplicity let us make an additional assumption:

(A4) the mapping υf(x,υ) is affine and the set U is convex.

Recall that in this case the operator ux[u] is continuous as a function L(I;Rn)C(I;Rn), where L is endowed with the weak* topology σ(L,L1) of duality to L1. The way the main results can be carried over to the general case (sliding modes of control) is discussed in § 7.

In the PMP theory the canonical class of rejecting controls is formed by needle-shaped variations, however, in the u-affine case it is sufficient to restrict the consideration to so-called weak variations:

¯uλ(t)=¯u(t)+λ(u(t)¯u(t)),uU,λ[0,1).
As is well known, the linear part of the mapping
λΔ¯uλI[¯u]I[¯uλ]I[¯u]=λddλ|λ=0I[¯uλ]+o(λ)
in a neighborhood of the origin (the first variation of the functional I, the Gateaux derivative in the direction u¯u) can be represented in the form
ddλ|λ=0I[¯uλ]=I(H(¯x(t),¯ψ(t),u(t))H(¯x(t),¯ψ(t),¯u(t)))dt,
where H(x,p,υ)pf(x,υ) is the classical Pontryagin function and ¯ψ=ψ[¯u] is the solution of the conjugate problem
˙ψ=xH(¯x,ψ,¯u)[Dxf¯u¯x]ψ,ψ(1)=x(¯x(1))
(a cotrajectory); denotes the operation of taking the matrix transpose. The inequality
I(minυUH(¯x(t),¯ψ(t),υ)H(¯x(t),¯ψ(t),¯u(t)))dtminuUddλ|λ=0I[¯u+λ(u¯u)]
and Filippov’s lemma provide a rule for choosing the function u in the expression (2.1):
H(¯x(t),¯ψ(t),u(t))=minυUH(¯x(t),¯ψ(t),υ)for a.e. tI,
which guarantees that for sufficiently small values of λ controls of the form (2.1) are ‘not worse than’ ¯u. In combination with a certain method for adjusting the parameter λ (the method of linear descent) formulae (2.1) and (2.5) describe an algorithm of consecutive approximations in problem (P), an analogue of the method of gradient descent (see, for example, [8]). The condition of the unimprovability of the control ¯u and the halting criterion is, obviously, the validity of the relation
H(¯x(t),¯ψ(t),¯u(t))=minuUH(¯x(t),¯ψ(t),u)for a.e. tI,
which determines the classical Pontryagin principle (in the form of the minimum principle).

It is obvious that if the initial conditions of the problem are sufficiently regular, then one can go even further and consider the Taylor series expansion of the function λΔ¯uλI[¯u] by taking the second and higher variations (a common practice is not to go beyond the second variation). This gives necessary conditions of higher order and more sophisticated methods for local search.

It is known (see [8]) that in the particular case when problem (P) has a linear-quadratic structure there is an analogue of representation (2.2), (2.3) without remainder terms, hence, without any parameters adjusting the closeness of the controls ¯u and u. Such a representation can be considered an ‘infinite-order’ variation of I. Moreover, if the functions xf(x,υ) and x(x) are linear (that is, the problem is linear in state), then this representation is fully expressed in terms of standard constructions of the PMP:4

ΔuI[¯u]=I(H(x(t),¯ψ(t),u(t))H(x(t),¯ψ(t),¯u(t)))dtuU.
Let us compare (2.7) and (2.3). The only difference between them is that now the first argument is a point on the ‘new’ trajectory x=x[u] rather than on the reference trajectory. As above, one can provide the inequality ΔuI[¯u]0 by choosing u so as to satisfy the condition of pointwise minimization of the integrand
u(t)argminυU{H(x[u](t),¯ψ(t),υ)H(x[u](t),¯ψ(t),¯u(t))},
which is obviously equivalent to the relation
H(x[u](t),¯ψ(t),u(t))=minυUH(x[u](t),¯ψ(t),υ)for a.e. tI.
However, in contrast to the pointwise condition (2.5), which gives us a control in an explicit form, the operator equation (2.8) determines the function u implicitly. If the control ¯u is not extremal, then at first sight it is by no means obvious that there exist any solutions of this equation in the class U.

One could act in the following way: first, distinguish a ‘control construction’ in the form of the feedback w(t,x) as a solution of the problem

H(x,¯ψ(t),w(t,x))=minυUH(x,¯ψ(t),υ),
then substitute this function into the system,
˙x=f(x,w(t,x)),x(0)=y,
and, finally, if the last system has a Carathéodory solution x, put u(t)=w(t,x(t)). This new program control would provide an optimality criterion for ¯u which has the same form as (2.6):
H(x(t),¯ψ(t),¯u(t))=H(x(t),¯ψ(t),u(t))minυUH(x(t),¯ψ(t),υ)for a.e. tI.
This is what a ‘feedback analogue’ of Pontryagin’s principle might look like.

Let us emphasize once again that, in contrast to the PMP, which involves only the process ¯σ to be tested for optimality, the formulation of the feedback condition assumes that there exists an additional comparison process σ=(x=x[u],u) which is not supposed to be close to ¯σ either in control or in trajectory.

Of course, in the case of a general position the function xw(t,x) is not continuous, the system (2.9) has no Carathéodory solutions, and nonclassical solutions (such as Krasovskii–Subbotin motions) are not generated by any control uU (for nonconvex U or a nonconvex set f(x,U){f(x,υ)|υU}). Moreover, formula (2.7) itself is valid only in the case when problem (P) is linear in state.

To carry over the idea presented above to the general nonlinear case it is convenient to embed the classical problem (P) into a weaker statement which is linear in the corresponding state variable, namely, the problem of control for an ensemble of trajectories (on the metric space of probability measures). This relaxation is discussed in § 3. The linearity of the weakened problem allows us to involve elements of the duality theory (§ 4). In § 5 we derive two symmetric exact formulae for the increment of the objective functional in the transformed problem (and, as a corollary, in (P)), which are similar to Weierstrass’ classical formula in the calculus of variations (see [12]). As a corollary of these formulae, we obtain a series of necessary optimality conditions similar to (2.10): in § 6 the corresponding conditions are derived in the case of affine dependence on u and a convex set U; this result is generalized in § 7 to the general statement by going over to sliding modes of control. Section 8 is devoted to the discussion of the status of the necessary conditions obtained in this work among similar results, in particular, their relation to Pontryagin’s principle. In § 9 we formulate the method of descent along the functional, which involves the extremal construction of feedback controls (1.7), and establish its convergence. Sections 10 and 11 contain necessary technical results.

§ 3. Relaxation

Let us show that a representation similar to (2.7) is valid for problem (P) of the general form. Below we use the notation fυ(x)f(x,υ) and Rn.

3.1. Flows of vector fields. The transport equation

We start by recalling some necessary facts. Suppose that the assumptions (A2) and (A3) are satisfied. Then for each uU the function (t,x)fu(t)(x) generates the mapping X=X[u]:(s,t,x)Xs,t(x), which is called the flow of the nonautonomous vector field f. Here tXs,t(x) is the solution of the Cauchy problem

tXs,t(x)=fu(t)(Xs,t(x)),Xs,s(x)=x.
For all s,tR the mapping Xs,t:xXs,t(x) is a C1-diffeomorphism RnRn with the property Xτ,tXs,τ=Xs,t for all s,τ,tR. These facts, in particular, imply the invertibility of Xs,t and the relation (Xs,t)1=Xt,s.

Fixing some s we introduce the shorthand notation Pt=Xs,t and Qt=Xt,s. Clearly,

0=t(id)=t(QtPt)=(tQt+DxQtfu(t))Pt,
where id denotes the identity mapping RnRn. Since the expression in parentheses vanishes for all values z=Pt(x) and the mapping xPt(x), RnRn, is bijective for each tI, it can be concluded that tQt satisfies on I×Rn the conditions
tQt+DxQtfu(t)=0,Qs=id,
which are treated in the same pointwise sense as equation (1.6). This yields a useful expression for the derivative of the flow with respect to the first index:
tXt,s=DxXt,sfu(t)Jt,sfu(t),
where sJt,s[u](x) for each xRn is a solution (see [13], Theorem 2.3.2) of the Cauchy problem for the linear matrix equation
sJt,s=Dxfu(s)Xt,s(x)Jt,s,Jt,t=E;
E=En denotes the identity matrix of size n×n. Let ξC1(Rn;R). Then it turns out that the function pξQ is5 a solution (in the sense indicated above) of the nonconservative transport equation
tpt+xptfu(t)=0
with the intermediate condition ps=ξ.

3.2. The problem of control for a (statistical) ensemble of trajectories

Note that the phase space Rn in problem (P) can be endowed with the natural structure of a probability space (Rn,F,P) by introducing the canonical probability measure6 P=δy. In this case the function tXt(x)X0,t(x)[u]x[u](t) must be treated as a (deterministic) random process. For each tI the distribution of the random variable xXt(x) is determined by the probability measure μt(Xt)δy=δx[u](t)P(Rn). It is well known that under the standard regularity assumptions the function μ=μ[u]:tμt, which describes the behaviour of this measure in time, is a weak solution (see the definition below) of the linear partial differential equation

tμt+x(fu(t)μt)=0.
This formal equation is a direct generalization of the classical continuity equation to the case of arbitrary probability (or nonnegative) measures. If P=δy, then it is equivalent to (its characteristic) ordinary differential equation (1.1): the only weak solution of the continuity equation on I with the initial condition μ0=δy for a control uU is the curve tδx[u](t).

In turn, the quality criterion for the problem (P) can be formulated in terms of the linear mapping P(Rn)R,

(x[u](1))=dμ1[u]μ1[u],,
the minimum of which over all μ=μ[u], uU, coincides with the value of (P). Thus, we arrive at an equivalent statement of the original control problem, which is now linear in the new state variable.

Now we put aside the particular choice of the probability structure (F,P) and consider the extremal problem

(RP)J[u],μ1[u]min,
tμt+x(fu(t)μt)=0,
μ0=ϑ,
uU.
Here the role of states is played by the probability measures μtP(Rn) on the phase space of problem (P); the initial distribution of ϑP(Rn) is specified. The class U of admissible controls remains the same. Assumptions (A1)(A4) are still supposed to be fulfilled, along with the additional assumption

(A5) the measure ϑ has a compact support (ϑPc(Rn)).

A weak solution of equation (3.6) is a function μC(I;P(Rn)) for which the Newton–Leibniz formula holds:

μτ,φτμs,φs=τsμt,tφt+xφtfu(t)dts,τI,s<τ,φC1c(I×Rn).

Remark 2. We adopt the nonclassical definition of a weak solution (with a wider class of test functions), which is here equivalent to the classical one; see [9], Remark 2.5 and Lemma 2.6.

It is known (see [9]) that under the assumptions (A2), (A3) and ϑP1(Rn) (in particular, (A5)) the weak solution of the Cauchy problem (3.6), (3.7) does exist, is unique and admits the representation

μt=(X0,t)ϑ.
Here the operator F:P(Rn)P(Rn) defines the image of the measure under the action of the Borel measurable vector field F:RnRn:
Fμ,φ=μ,φFφ:φFLμ1(Rn;Rn).

Remark 3. Under assumption (A5) the family of measures (3.9) satisfies condition (3.8) for an arbitrary φC1(I×Rn) (φ does not necessarily has a compact support with respect to x).

The setup (RP) is called the ensemble control problem. It generalizes problem (P) to the case of an uncertain initial state (more realistic from the standpoint of applications). Although the variational analysis of the original problem on the basis of exact formulae for the increment can be performed directly, the approach proposed in our work can almost literally be carried over to the generalized model, and it is reasonable (and is even simpler in a certain sense) to present it in terms of the latter. Moreover, the main advantage of this approach — the absence of variation parameters — is most pronounced in problems of control for distributed systems, in which Pontryagin’s principle is formulated in terms of the conjugate partial differential equation (see [14], Theorem 2) and the ‘computational cost’ of the classical and feedback optimality condition is almost the same.

Note that under assumptions (A1)(A4) the minimum in (RP) is attained in the class U of admissible controls: see [15], Theorem 3.2.

The reader familiar with geometric control theory can draw a parallel between the setup (RP) and the formalism of chronological calculus [16], [17], in which the probability structure is replaced with the algebraic one.

§ 4. Duality

Let ξC1(Rn;R) and uU be fixed, and let X=X[u] be the flow of the system (1.1) corresponding to the control u. Consider the function p=p[u]: [s,1]×RnR,

pt=ξX1s,tξXt,s.
As mentioned above, this function is a solution of the Cauchy problem (3.5). It is easily seen that the action of the measure μt on pt does not depend on t:
μt,ptpt,(X0,t)ϑ=ptX0,t,ϑ=ξX0,s,ϑ,
which allows us to consider the trajectory p as the conjugate of μ. Using this we can get rid of the variable μ and reformulate problem (RP) in terms of the variable p. Indeed, putting s=1 and ξ= we obtain
μ1,μ1,p1=μ0,p0ϑ,p0
and conclude that the optimum in (RP) coincides with the solution of the problem
(RP)J[p]ϑ,p0[u]min,tpt+xptfu(t)=0,p1=,uU.
In particular, taking ϑ=δy gives us the setup
(P)I[p]p0[u](y)min,tpt+xptfu(t)=0,p1=,uU,
which is equivalent to the original (classical nonlinear) problem (P). This result is closely related to Theorem 2.1 in [18] (see also [19]) and it can be looked upon as a generalization of relations of nonclassical duality (Remark 4 in § 5; see [20]).

§ 5. Exact formulae for the increment

5.1. The increment of the functional of the weakened problem

Problem (RP) admits an exact representation of the increment of the functional which is similar to (2.7).

Proposition 1. Suppose that conditions (A1)(A3) hold. Let u,¯uU, u¯u, be arbitrary controls, X=X[u] and ¯X=X[¯u] be the corresponding flows of the characteristic system (1.1) and μ=μ[u] be the solution of equation (3.6) corresponding to the control u.

Then the increment ΔuJ[¯u]J[u]J[¯u] of the objective functional of problem (RP) can be represented in the form

ΔuJ[¯u]=Iμt,¯Ht(,u(t))¯Ht(,¯u(t))dt,
where
¯Ht(x,υ)H(x,¯ξt(x),υ),¯ξt=x¯pt¯Jt¯Xt,1,
H is the classical Pontryagin function and ¯p is defined by condition (4.1) for ξ=, s=1, and u=¯u; the matrix ¯Jt¯Jt,1Dx¯Xt,1 is the solution of the Cauchy problem (3.4) at the instant s=1 for u=¯u.

Proof. The proof of Proposition 1 is based on elementary facts from calculus and the theory of ordinary differential equations.

(1) Let us show that the function s¯Xs,1X0,s(x), IR, is Lipschitz continuous for each xRn. Consider the orbit OI(x)={X0,1[ωs[u,¯u]](x)sI} of the point x under the mapping sX0,1[ωs[u,¯u]](x)¯Xs,1X0,s(x), IRn, where

ωs[u,v]{uon [0,s),von [s,1].
The standard arguments based on the Grönwall–Bellman inequality show that under assumptions (A2) and (A3) the set OI(x) is bounded. Then its closure clOI(x) is a compact subset of Rn. It follows from assumption (A1) that the function is locally Lipschitz continuous, hence (in view of the local compactness of Rn) its restriction to clOI(x) is Lipschitz continuous. Now the required fact follows from the Lipschitz continuity of the functions tX0,t(x) and t¯Xt,1(x) with regard to the uniform (in t) local Lipschitz continuity of the function x¯Xt,1(x) (Lemma 1):
|¯Xt,1X0,t(x)¯Xs,1X0,s(x)|L1|¯Xt,1X0,t(x)¯Xs,1X0,s(x)|L1|¯Xt,1X0,t(x)¯Xt,1X0,s(x)|+L1|¯Xt,1X0,s(x)¯Xs,1X0,s(x)|L1(L2L3+L4)|ts|,
where L1=Lip(;clOI(x)) is the Lipschitz constant of the objective function on clOI(x), L2=Lip(¯Xt,1();{X0,t(x)tI}) is the Lipschitz constant of x¯Xt,1(x) on the phase portrait {X0,t(x)tI}, L3=Lip(X0,(x);I) and L4=maxsILip(¯X,1(X0,s(x));I).

(2) With regard to the definition (3.10) and the equalities Xs,s=¯Xs,s=id for all s we represent the increment of the functional in the form

ΔuJ[¯u]μ1¯μ1,=ϑ,X0,1¯X0,1=ϑ,¯X1,1X0,1¯X1,1¯X0,1ϑ,¯X0,1X0,0¯X0,1¯X0,00.
As the mapping t¯Xt,1X0,t(x) is absolutely continuous, one can extend the chain of equalities and convert the last difference with the use of the Newton–Leibniz formula:
ϑ,It(¯Xt,1X0,t¯Xt,1¯X0,t)dt.
By the semigroup property of the flow ¯X the quantity ¯Xt,1¯X0,t=¯X0,1 does not depend on t. Consequently,
ΔuJ[¯u]=ϑ,It(¯Xt,1X0,t)dt.
Let us calculate the derivative under the integral sign:
t(¯Xt,1X0,t)=[¯Xt,1(t¯Xt,1+Dx¯Xt,1fu(t))]X0,t.
Introducing the notation ¯JtDx¯X1,t and taking the expression (3.3) (for s=1 and u=¯u) into account gives
ΔuJ[¯u]=ϑ,I((¯Xt,1)¯Jt[fu(t)f¯u(t)])X0,tdt.
To finish the proof it remains to apply Fubini’s theorem and take representation (3.9) into account.

The proof of the proposition is complete.

5.2. A ‘direct’ formula for the increment in problem (P)

We refine Proposition 1 for the original setup (P). Putting ϑ=δy (which yields the equality μt[u]=δx[u](t) for each tI) we obtain J[u]=I[u]. Then (5.1) takes the form

ΔuI[¯u]=I(¯Ht(x(t),u(t))¯Ht(x(t),¯u(t)))dt,
where ¯H is defined in (5.2).

Remark 4. It is easily seen that in problem (P), which is linear in state, the composition x¯p¯x coincides on I with the reference cotrajectory ¯ψψ[¯u]. This follows from the representation

x¯pt|x=¯x(t)=([Dx¯Xt,1]¯Xt,1)|x=¯x(t)¯Jt,1(¯x(1)),
where t¯Jt,1 is the fundamental matrix solution in the inverse time of the equation in (2.4) (here ¯Jt,1 does not depend on x). In this case formula (5.1) reduces to (2.7).

In the nonlinear case the equality x¯p¯x=¯ψ holds under an additional regularity assumption:

(A6) the function is twice continuously differentiable, as also is the function xf(x,υ) for each υU.

This fact is established by direct differentiation of the function tx¯pt(¯x(t)).

5.3. The dual formula for the increment

Renaming ¯uu, we obtain a ‘dual’ representation for the increment of the functional in problem (RP):

ΔuJ[¯u]=I¯μt,Ht(,u(t))Ht(,¯u(t))dt,
where Ht(x,υ)H(x,ξt(x),υ), ξt=xpt[u]JtXt,1 and p=p[u] is defined by condition (4.1) for s=1 and the control u (X=X[u] is the corresponding flow (1.1) and tJtDxXt,1 is a solution of (3.4)). Refining this representation of problem (P) we obtain an exact formula for the increment:
ΔuI[¯u]=I(Ht(¯x(t),u(t))Ht(¯x(t),¯u(t)))dt,
which is symmetric to (5.6).

5.4. A geometric interpretation

Let u and ¯u be arbitrary admissible controls, and X and ¯X be the corresponding flows. For each s[0,1] consider the intermediate control ωs[u,¯u]U defined by equality (5.3) and the flow Xs of the system (1.1) generated by this control. Note that Xs0,1=¯Xs,1X0,s. It is obvious that the function γ:IRn,

γ(s)=Xs0,1(y),sI,
specifies a parametrization of a curve on the attainability set D1(y) of the system (1.1), (1.2). This curve joins the points ¯X0,1(y)¯x(1)x[¯u](1) and X0,1(y)x(1)x[u](1) (Figure 1). Recall (formula (4.1)) that ¯ps=¯Xs,1. Hence ¯ps(x) is the cost of the reference control ¯u in problem (P) as restricted to the interval of time [s,1] with the initial condition x(s)=x. Now assume that u takes the initial state x(0)=y to a point x in time s, that is, x=X0,s(y). Then ¯ps(x) is the cost of the intermediate control ωs[u,¯u] in (P). A small variation s+Δt of the moment of ‘switching’ between the controls u and ¯u has the cost
¯ps+Δt(Xs,s+Δt(x))¯ps+Δt(x+Δtf(x,u(s)))¯ps(x)+Δt(t¯ps(x)+x¯ps(x)f(x,u(s))).

On the other hand ¯p obeys the equation

t¯ps(x)+x¯ps(x)f(x,¯u(s))=0.
Therefore,
¯ps+Δt(Xs,s+Δt(x))¯ps(x)Δtx¯ps(x)(f(x,u(s))f(x,¯u(s))).
It remains to recall that ¯ps(x)=(Xs0,1(y)) and ¯ps+Δt(Xs,s+Δt(x))=(Xs+Δt0,1(y)). Consequently, the rate of change of the cost along the curve γ is
dds(γ(s))=x¯ps(x(s))[f(x(s),u(s))f(x(s),¯u(s))]¯Hs(x(s),u(s))¯Hs(x(s),¯u(s)),
where x(s)=X0,s(y). Noting that
ΔuI[¯u]=Idds(γ(s))ds
we obtain an exact formula for the increment (5.6).

For a rigorous proof of relation (5.10) we go in the opposite direction: we apply the formula for the increment (5.6) and take into account that the curve γ is Lipschitz continuous (see the proof of Proposition 1). This, in particular, shows that (5.10) holds only for almost all s[0,1].

The dual formula (5.8) has a similar representation involving another class of curves ζ on D1(y) generated by variations ωs[¯u,u] of the control of the form (5.3) with the arguments ¯u and u in the reverse order, and joining the points ¯x(1) and x(1) in the ‘opposite direction’: ζ(0)=x(1) and ζ(1)=¯x(1). It is clear that there exist even more sophisticated parametrizations of the ‘motion’ between ¯x(1) and x(1) inside D1(y) — those akin to ‘packages of needles’ in the PMP theory — for example, parametrizations corresponding to the controls

ωs={uon k1j=0[jk,jk+s),¯uotherwise,s[0,1kr],k=2,3,.

Representation (5.10) suggests an obvious way to organize a monotone descent along the functional I: one must construct a curve γ along which the function does not increase, that is, γ˙γ0 almost everywhere on I. For example, one can take a feedback control w satisfying inclusion (1.7). Then the process (x,u) with control u(t)=w(t,x(t)) (of course, if such a process is well defined) generates the required curve γ and therefore is an improving process. A rigorous implementation of this idea is the subject of the rest of this paper.

§ 6. Necessary conditions of optimality

We turn back to the problem of the generation of descent directions for (P). Throughout this section assumptions (A1)(A4) are supposed to hold.

6.1. The principle of optimality with feedback controls

As discussed in § 2, the sign ΔuI[¯u]0 is guaranteed by the choice of u so as to minimize the integrand in (5.6). This turns us back to the problem of the solvability of the operator equation

¯Ht(x[u](t),u(t))=minυU¯Ht(x[u](t),υ)for a.e. tI,
similar to (2.8), in the class of admissible program controls U.

Let us verify that for each ¯uU the set

S[¯u]{σ=(x,u)x=x[u],uU, and (6.1) holds}
is nonempty. To be more precise, we demonstrate that the required property is characteristic for all processes in which:

x is an arbitrary Krasovskii–Subbotin solution (see § 12) corresponding to some feedback control

w(t,x)argminυU¯Ht(x,υ)argminυUH(x,x¯pt(x),υ)
(the set of such solutions is denoted by ¯KS),

uU is an arbitrary program control generating the function x as a solution of system (1.1) (such a control exists by Proposition 6; clearly, if w is Borel measurable, then the class of such controls includes the Borel equivalence class of u such that u(t)w(t,x(t)) for almost all tI).

Proposition 2. Let x¯KS, uU and x=x[u]. Then the pair σ(x,u) satisfies (6.1).

This fact follows from a more general result of Proposition 4 for sliding modes.

It follows from the representation (5.6) that in problem (P) any control u satisfying (6.1) is ‘not worse’ than the reference one, that is, I[¯u]I[u]. If the process ¯σ is optimal, then it is obvious that any process σS[¯u] is optimal as well. This observation can be considered a necessary optimality condition in the spirit of the feedback conditions [1]. It is clear that if this condition holds, then I[¯u]=I[u] for all σS[¯u].

Note that for any u having the property (6.1) the integrand in (5.6) is nonpositive; then we can reformulate our necessary condition in a form close to the PMP.

Theorem 1 (minimum principle with feedback controls). Let ¯σ=(¯x,¯u) be an optimal process in problem (P). Then the condition

¯Ht(x(t),¯u(t))=minυU¯Ht(x(t),υ) (=¯Ht(x(t),u(t)))for almost all tI
holds for each σ=(x,u)S[¯u].

In fact, Theorem 1 contains a series of necessary conditions parametrized by the (nonempty) set S[¯u], which we still call comparison processes; in contrast to [1] it is only admissible processes that we admit to comparison. This theorem proposes the concept of ‘feedback extremal’ alternative to [3]: a feedback extremal is a pair (¯σ,σ) of processes satisfying condition (6.3) (in particular, the equality in parenthesis). With regard to Remark 4 the claim that the process ¯σ is extremal in the classical sense is equivalent to the inclusion ¯σS[¯u] which means that the pair (¯σ,¯σ) is a feedback extremal. The relationship between the two types of extremality is discussed in § 8.1 below.

In conclusion we give a rigorous interpretation of the feedback condition in terms of the curve of monotone descent on the attainability set of the control system (§ 5.4).

Proposition 3. Let ¯σ=(¯x,¯u) be an admissible process, σ=(x,u)S[¯u] be some comparison process and the curve γ:IRn be defined by condition (5.9). Then the objective function does not increase along γ.

Proof. In view of (5.10) and (6.3) we have
dds(γ(s))=¯Hs(x(s),u(s))¯Hs(x(s),¯u(s))=minυU¯Hs(x(s),υ)¯Hs(x(s),¯u(s))0.
The proof of the proposition is complete.

6.2. Descent controls in the weakened problem. Co- and bifeedback optimality conditions

The method for the generation of comparison controls that was presented in § 6 remains unchanged in the weakened problem (RP): the expression (5.1) gives the structure of a descent control in the form of a feedback with respect to the measure μ (it characterizes the state of the system):

w[μ](t)argminυUμ,¯Ht(,υ).
This representation can be used to derive a feedback necessary condition and to construct nonlocal numerical algorithms in the problem of control for an ensemble of trajectories. Some results in this area were obtained in [21].

The dual representations (5.7) and (5.8) suggest a construction for a cofeedback descent control in the form of a functional feedback

w[p](t)argminυUH(¯x(t),xpt(¯x(t)),υ)
and produce a series of ‘cofeedback’ necessary conditions, which can be combined with Theorem 1. The feedback strategies (6.4) can be implemented with the use of the Krasovskii–Subbotin scheme in reverse time, starting from the point ¯x(1)=ζ(1)Xs,1¯X0,s|s=1 (see § 5.4). We do not elaborate here this idea and limit out considerations to the direct approach.

§ 7. General case. Sliding modes

Now we abandon assumption (A4) and suppose that υf(x,υ) is an arbitrary mapping satisfying (A3) and U is an arbitrary (nonconvex) compact set. Although this case is much more general, from the technical point of view it is little different from the one discussed above, provided that we apply a classical trick originating from the theory of Young measures (see [22]): we relax the class of admissible controls U by identifying functions u with elements ν of the set

Y=Y(U){νP(I×U)[(t,υ)t]ν=L1|I}
whose marginals tνt — families of measures obtained by desintegrating ν with respect to the Lebesgue measure L1 on I — have the form tδu(t) (in control theory the mappings tνt are called controls of Gamkrelidze or Warga–Gamkrelidze type, in differential games they are called mixed strategies). This causes a relaxation of the original control system
x(t)=y+[0,t]×Uf(x(s),υ)dν(s,υ),νY,
which corresponds to the convexification of the set of velocities (1.1), (1.2):
˙x=Uf(x,υ)dνt(υ)˙xco{f(x,υ)υU};x(0)=y.
Let x=x[ν] be the solution of (7.1) that corresponds to νY. Processes (x,ν) are called sliding modes of the control. The mapping νx[ν] is well known to be continuous as a function YC(I;Rn), where Y is endowed with the topology of weak convergence of probability measures. The flow (s,t,x)Xs,t[ν] of system (7.1) and the corresponding mapping (s,t,x)Js,t[ν] can be defined in the same way as in § 3.1.

It is obvious that problem (coRP) of the minimization of the linear form ,μ1[ν] over the curves in the family tμt[ν](X0,t[ν])ϑ, νY, is the convexification of problem (RP). The convexified version (coP) of the original problem (P) is a particular case of (coRP) that corresponds to the initial measure ϑ=δy. Note that equation (7.1) is linear in the variable of generalized control and the corresponding (weakened) problem (coRP) is bilinear in the pair (μ,ν).

Now all results of § 6 can be rewritten (almost word for word) in terms of the generalized control ν and the function

¯Ht(x,ϱ)U¯Ht(x,υ)dϱ(υ),I×Rn×P(U)R.
The following proposition holds.

Proposition 4. Let x¯KS, νY and x=x[ν]. Then

¯Ht(x(t),νt)=minϱP(U)¯Ht(x(t),ϱ)for a.e. tI.

Proof. One must only notice that the function
g(t,x,ω)¯Ht(x,ω)minυU¯Ht(x,υ)
satisfies all the assumptions of Proposition 7 by virtue of Lemma 2. The proof is complete.

We note that the minimum on the right-hand side of (7.2) is attained at any measure ϱP(U) with the property

spt(ϱ)argminυU¯Ht(x(t),υ),
where spt denotes the support of the measure. Proposition 4 claims the existence of (at least one) comparison process in the class of sliding modes; under assumption (A4) it reduces to Proposition 2. In turn, a direct generalization of Theorem 1 is as follows.

Theorem 2. Let (¯x,¯ν) be an optimal process of problem (coP). Then the relation

¯Ht(x(t),¯νt)=minϱP(U)¯Ht(x(t),ϱ)for a.e. tI
holds for any pair (x=x[ν],ν) satisfying (7.2).

Note that for a ‘regular’ process ¯σ=(¯x,¯u) condition (7.3) reduces to

(¯Ht(x(t),δ¯u(t)))¯Ht(x(t),¯u(t))=minϱP(U)¯Ht(x(t),ϱ)for a.e. tI.

The result of Theorem 2 can be refined without using any additional information, based only on the knowledge of the reference process.

Theorem 3. Let (¯x,¯ν) be the optimal process in problem (coP). Then the relation

¯Ht(x(t),¯νt)=minϱP(U)¯Ht(x(t),ϱ)for a.e. tI
holds for each sliding mode (x=x[ν],ν) satisfying the condition
spt(νt)argminυU¯Ht(x(t),υ)spt(¯νt)for a.e. tI.

As above, to check a process ¯σ in the original problem (P) for optimality, one can rewrite conditions (7.4) and (7.5) in the form

¯Ht(x(t),¯u(t))=minϱP(U)¯Ht(x(t),ϱ)for a.e. tI
and
spt(νt)argminυU¯Ht(x(t),υ){¯u(t)}for a.e. tI,
respectively. Since one selector of the multivalued mapping IP(U) generated by the last inclusion is the function tδ¯u(t), condition (7.4), in particular, contains Pontryagin’s principle. The same ‘additive’ inclusion of Pontryagin’s extremals in the class of comparison processes is assumed by the formulation of the feedback minimum principle [1]. However, the relationship between the conditions mentioned above and Theorems 1 and 2 is not that trivial. It is discussed in the next section.

§ 8. Discussion and examples

Here we discuss the status of Theorems 1 and 2 among close results.

8.1. The relation to Pontryagin’s principle

We turn back to problem (P) under assumptions (A1)(A4) and (A6). It follows from the equality x¯p¯x=¯ψ (Remark 4) that the process ¯σ satisfies the PMP once the comparison trajectory x coincides with ¯x on the whole interval I (it is possible, however, that u¯u). This is definitely so, for example, in the case when for all (t,x)I×Rn the extremal problem (6.2) has a unique solution.7 We present several versions of such statements. We introduce a feedback analogue of the regularity property for an extremal, namely, the absence of so-called singular pieces of control components.

Definition 1. Let (¯σ,σ) be a pair of admissible processes. A component ¯uk of a control ¯u=(¯u1,,¯um) is said to be regular if

υk¯Ht|x=x(t)0tI
(except, perhaps, finitely many points tI). The pair (¯σ,σ) is said to be regular if inequality (8.1) holds for every k=1,,m.

We can verify that in some typical situations all regular feedback extremals of problem (P) which is affine in control are extremals of Pontryagin’s principle.

Theorem 4 (relation to PMP). Suppose that assumptions (A1)(A4) and (A6) are satisfied, and let (¯σ,σ) be a feedback extremal. Suppose that one of the following conditions holds.

Then the process ¯σ is an extremal in the sense of the PMP and σ=¯σ.

Proof. (1) It is obvious that in the case under study the finite-dimensional extremal problem minυU¯H is separable, which means that
minυU¯Ht(x,υ)=mk=1minυkUkυk¯Ht(x)υk(t,x)I×Rn.
It is also clear that the regularity of the component ¯uk, in combination with condition (6.3), implies the equality uk(t)=¯uk(t) for almost all tI, and since all components of ¯u are supposed to be regular, we have u=¯u. Now the PMP-extremality of ¯σ follows from the equality x¯p¯x=¯ψ.

(2) By (8.1). for almost all tI the linear form υυυ¯Ht|x=x(t) is nondegenerate and all of its minimum points lie on the boundary of the compact set U; if U is strictly convex, then such a point is unique. Then, as above, condition (6.3) implies the equality u=¯u.

The proof of the theorem is complete.

In the regular case the feedback condition is not weaker than the PMP (and Example 3 for α=0 illustrates the phenomenon of strict strengthening). A natural question arises: what are the types of PMP-extremal processes excluded/not excluded by Theorems 1 and 2? We have already seen that processes in the class S[¯u] cannot be better than the PMP-extremal ¯σ corresponding to the point of local minimum of the function on the attainability set of the (convexified) system, since the principle of the construction of γ as a curve of monotone nonincreasing does not assume the ‘ascension’ along level lines of to get out of ¯x(1). However, feedback extremals can correspond to other types of stationary points (Example 3).

8.2. The relation to the feedback minimum principle

The centerpiece of the theory of feedback necessary conditions is the so-called feedback minimum principle (FMP). This condition was originally formulated in terms of a linear majorant of the objective functional with a reference cotrajectory (that is, in the framework of the standard objects of the PMP; see [1]), while the most general result of this type — with a nonlinear majorant — was presented in [3]. Recall the original formulation of the FMP in problem (P): the optimality of a pair ¯σ=(¯x,¯u) implies the optimality of the curve ¯x in the so-called adjoint problem

(AP¯σ)min(x(T)),xX[¯u],
where X[¯u] is the set of all Carathéodory and Krasovskii–Subbotin solutions corresponding to the selectors w(t,x)argminυUH(x,p(t,x),υ) and the function p is defined by the expression
p(t,x)¯ψ(t)+x(x)x(¯x(t)).
Here the construction of feedback controls is the same as in (6.2) up to replacing x¯p by p. In contrast to the conditions obtained above, the FMP has a variational form. In this setting it is assumed that the optimal process ¯σ

Here condition (a) yields directly (see [2]) the extremality of the process ¯σ, as the PMP is certainly ‘included’ in the FMP; this (slightly unnatural) inclusion is provided by the use of feedback Carathéodory solutions, which can be absent if ¯σ is not extremal. As illustrated by examples below, the PMP by no means follows from the ‘essential’ part of the FMP, condition (b).

Let us compare the FMP with Theorems 1 and 2. To do this we reveal the relationship between the function p and the reference solution ¯p of the transport equation. If is linear and the assumption (A6) holds, then p=¯ψ is obviously the gradient x of the linear approximation of the solution ¯p in a neighbourhood of the characteristic curve ¯x:

¯pt(x)¯pt(¯x(t))+x¯pt(¯x(t))(x¯x(t))(¯x(1))+¯ψ(t)(x¯x(t)).
In particular, in the bilinear problem we have p=¯ψx¯pt (Remark 4). In the general case we have p=xη, where
ηt(x)(x)+(¯ψ(t)x(¯x(t)))(x¯x(t))+(¯x(1))(¯x(t))¯pt(x)+(x)(¯x(t))x(¯x(t))(x¯x(t)).
The function η is a rather crude approximation of the ‘exact majorant’ ¯p, which combines (8.2) with the linear approximation of the objective function: (x)(¯x(t))+x(¯x(t))(x¯x(t)).

If (P) is linear in state, then Theorem 2 can be weakened by restricting it to the sliding modes, which are applied only to the set of curves X[¯u] (our formulation also allows other curves). Then the result obtained reduces to (8.2). Hence the proposed necessary condition is not weaker than the FMP. In the bilinear problem the FMP (in its essential part (b)) coincides with the statement of Theorem 1.

8.3. Examples

We begin by illustrating an application of Theorem 2 in comparison with the PMP and FMP.

Example 1. Consider the problem

I[u]=z(1)y2(1)2inf,˙x=u,˙y=u2,˙z=x22,x(0)=y(0)=z(0)=0;|u|1,
in which the infimum is attained at the sliding mode (ˇx,ˇν), ˇνt12(δ1+δ1). We write out the auxiliary constructions of the PMP. The Pontryagian has the form
H(x,p,υ)=H(x,px,py,pz,υ)=pxυ+pyυ2+12pzx2,
and the conjugate trajectory ψ=(ψx,ψy,ψz) is defined by the conditions
ψyy(1),ψz1;˙ψx=xandψx(1)=0.

(1) Improving an extremal process. Consider the control ¯u0 that generates the direct and conjugate trajectories

¯x=¯y=¯z0;¯ψ=(¯ψx,¯ψy,¯ψz)(0,0,1).
Clearly, the process ¯σ=(¯x,¯y,¯z;¯u) is a singular extremal and I[¯u]=0. We apply the FMP and Theorem 2 to this process. Here p(0,y,1) and ¯pt=(1t)x2/2y2/2+z. The FMP offers for comparison all ‘feedback’ controls
w(t,x)argmin|υ|1(yυ2)={[1,1]U,y0,{1,1},y=0,
and condition (6.2) takes the form
w(t,x)argmin|υ|1((1t)xυyυ2)={sign(t1)x2y,y0,sign(t1)x,y=0.
Both these classes include descent strategies generating, in particular, the optimal sliding mode by the Krasovskii–Subbotin scheme. However, it is only the second class that provides important information, namely, the explicit structure of the feedback: since y(0)=0, the FMP gives back the original set of (all admissible) program controls, which means that it actually degenerates.

(2) The FMP: lack of improvement of a nonextremal process. The control ¯u1 as the reference control gives

¯x(t)=¯y(t)=t,  ¯z(t)=t36;¯ψx(t)1t22,  ¯ψy¯y(1)=1;I[¯u]=13.
It is easily seen that this process is not extremal in the sense of Pontryagin. We consider the function
p(t,x,y,z)¯ψ(t)+(x,y,z)(¯x(t),¯y(t),¯z(t))=(1t22,(y+1t),1)
and define the multivalued mapping ¯U(t,x,y,z) as the solution of the minimization problem
H(x,p(t,x,y,z),υ)x22=1t22υ(y+1t)υ2min,|υ|1.
We are interested in the case when y0 and t[0,1]. Note that the quantity γ=γ(t,y)(y+1t) is strictly positive for t[0,1) (the point t=1 can be ignored). Hence we have to minimize the concave function
βυγυ2=γυ(βγυ),γ>0.
Here the quantity β=β(t)(1t2)/2 is also strictly positive for t[0,1), and therefore the minimization over υ[1,1] gives us the extremal mapping ¯U{1}. In other words, all feedback controls of the FMP are exhausted by the single function w1, which generates the unique (in any sense) solution
x(t)=y(t)=¯x(t)=¯y(t)=t,z(t)=¯z(t)=t36
with the same cost as the reference process: I[u1]=1/3. We arrive at the conclusion that the FMP does not improve the nonextremal process under consideration.

Thus, the sets of processes satisfying the PMP and the comparison condition from the FMP are not proper subsets of each other. This means that the PMP and FMP are in fact two independent necessary conditions of optimality.

(3) Theorem 2: improvement of the control ¯u1. To apply Theorem 2 we find that

¯pt=(1t+x)3x36(1t+y)22+z(x,y,z)¯pt=((1t+x)2x22,(1t+y),1).
The minimization of the function
H(x,¯pt(x,y,z),υ)x22=(1t)2+2(1t)x2υ(y+1t)υ2
over υ[1,1] for small values of t gives us a single control δ1, which is realized till the moment of time t=1/3. On the interval [1/3,1] the sliding mode ν=34δ1+14δ1 is applied, for which x(t)=(t1)/2 and ¯U={±1}. As a result, the new process looks as follows:
x(t)={t,t[0,13),t12,t[13,1],y(t)=t,z(t)={t36,t[0,13),(t1)324+154,t[13,1],
and it has cost I[ν]=13/27<1/3. We see that in the case under study Theorem 2 improves both the PMP and FMP.

As shown in [3], in this example even the second-order FMP ‘does not work’. The improvement is due to a certain strengthening of the FMP by passing to the extremal of the convexified problem on a series of controls which is equivalent to ¯u (see [3], Proposition 1).

In Example 1 we see ‘looping’ iterations of FMP, however, the feedback controls obtained there does not disimprove the reference process in any case. Let us show that the use of the linear majorant in the nonlinear problem can cause strict disimprovement of a nonextremal control.

Example 2. The problem has the form

I[u]=y(1)inf,˙x=u,˙y=x2x3;x(0)=ε>0,y(0)=0;|u|1.
The dynamics of the system is organized so that the strategy u1 is optimal for large values of the parameter ε and counteroptimal for small values. Consider a nonextremal control ¯u0 and write out all objects of the PMP:
H=pxυ+py(x2x3);ψy1;˙ψx=3x22x,ψx(1)=0.
The reference ‘bitrajectory’ has the form
¯xε,¯y=(ε2ε3)t;¯ψx=ε(3ε2)(t1),
and I[¯u]=ε2ε3. For 0<ε<2/3 the only solution of the problem
pxυ=¯ψx(t)υε(3ε2)(t1)υmin,|υ|1,
on [0,1) is the program w1, which generates the trajectory
x[w](t)=εt,y[w](t)=(tε)33+(tε)44+ε33ε44
with cost
I[w]=(1ε)33+(1ε)44+ε33ε44>ε2ε3=I[¯u],0<ε1
(this inequality becomes obvious as ε tends to zero).

As in the previous problem, here the phase variables and control variable are separated and the functional is linear. Therefore, the control generated by the FMP turns out to be a program control. At the initial instant the FMP ‘determines’ correctly the direction of local descent, but the absence of feedback makes it impossible to adapt the strategy afterwards. As a result, the FMP ‘makes a mistake’, and this mistake is stable under small variations of the parameter ε.

In contrast to the FMP, Theorem 1 ‘introduces’ the lacking feedback and generates a locally optimal synthesis w(t,x)=sign(23x)x(1t) (here w(0,ε)=1 for small values of ε and w(0,ε)=1 for large values); we leave out the calculations and only indicate that ¯pt(x,y)=(x2x3)(1t)+y).

The last example illustrates the phenomenon of the absence of feedback improvement of a PMP-extremal not corresponding to a point of local minimum of the objective function.

Example 3. Consider the following problem:

I[u]=y(1)+z(1)min,˙x=u,˙y=xu,˙z=α2u2,x(0)=y(0)=z(0)=0;|u|1.
Here H=(pxx)υ+(α/2)υ2. For any α0 the control ¯u0 corresponds to an extremal (singular for α=0 and strict otherwise) with cotrajectory ¯ψx0, ¯ψy=¯ψz1. It is easily seen that for α(0,1) the vector (¯y(1),¯z(1)) is neither a point of local minimum, nor a point of local maximum of the function (x,y,z)=y+z on the attainability set of the convexified system10 (for example, for a constant u0 we have I[u]<0, and for a sliding mode νλ,ε=(1λ)δ0+(λ/2)(δε+δε), where ε,λ(0,1], it turns out that I[νλ,ε]>0). Since the problem is linear in state, the FMP and Theorem 1 produce the same result (¯py+z). For α=0 (when the problem is bilinear) this result is the feedback control w(x)=signx generating, in particular, the global solutions u±1. However, for α>0 the only feedback strategy
w(x)={1,x<α,xα,|x|α,1,x>α,
admitted for comparison leaves the reference point unchanged.

In conclusion, we comment on the role of feedback conditions in calculus of variations: in contrast to the PMP, which sometimes provides information about the structure and properties of an unknown optimal process, Theorems 1 and 2 (as well as the FMP) cannot be applied without the knowledge of the reference approximation ¯u. From the analytic point of view it is reasonable to consider these results as an additional test for the optimality of the already constructed PMP-extremal control. The direct iterative application of these theorems produces an algorithm for the numerical solution of the optimization problems under study. The formulation and properties of this algorithm are discussed in the next section.

§ 9. Descent method

Let us turn again to problem (P) that is affine in control. On the set U×U we introduce the nonnegative functional

E[u,v]I(H(x,xpt[u](x),u(t))minυUH(x,xpt[u](x),υ))|x=x[v](t)dt.
Clearly, the equality E[u,v]=0 is equivalent to the claim that (x[v],v) is a comparison process for (x[u],u), and E[u,u]=0 means the PMP-extremality of u.

Let us describe an iteration of the conceptual descent algorithm based on Theorem 1.

Descent method. Put u0=¯u and suppose that ukU has already been calculated.

Remark 5. Finding the function ¯p as a solution of the transport partial differential equation is a well-known computational problem, which can be solved by the classical grid methods of integration only in spaces of low dimension (actually, for n3). However, with the use of the explicit representation ¯pt=¯Xt,1 and the Krasovskii–Subbotin method this obstacle can be avoided: let (ti,xi) be the current node of the polygonal approximation of the synthesized trajectory (§ 12). In accordance with the method of descent, the computation of the (i+1)st node assumes the knowledge of the gradient x¯pt(x)¯Jt(x)(Xt,1(x)) at the point (ti,xi). To achieve this it is sufficient to solve the phase system (1.1) and the linearized system (3.4) with the Cauchy conditions x(ti)=xi and Jti,ti=E, respectively.

The convergence of the method of descent is established by the following result.

Proposition 5. Suppose that assumptions (A1)(A4) are fulfilled, and let (uk) be a sequence of controls produced by the method. Then the following hold.

(1) There exists a subsequence ωkjωk{(u2k,u2k+1)} that converges in the space U×U with the direct product topology, where U is equipped with the topology σ(L,L1).

(2) Let (u,v) be a partial limit of the sequence ωk. Then E[u,v]=0.

Proof. (1) Since U is a convex compact set in Rm, the family U is compact in the space L with the topology σ(L,L1) by the Banach–Alaoglu theorem. Then by Tychonoff’s classical theorem the space U×U is compact in the direct product topology. Since L1 is separable, the weak* topology σ(L,L1) on U is metrizable, and therefore compactness is equivalent to sequential compactness, that is, the existence of a convergent subsequence.

(2) Under the above assumptions the operators uX[u](x) and uJ[u](x) are continuous as functions UC(I;Rn) for any xRn (as systems (3.1) and (3.4) are affine in control). Hence for each xRn the mapping uxp[u](x)X,1[u](x)J[u](x) is a continuous function UC(I;Rn). Since the function xxp[u](x), RnRn, is also continuous, the mapping (u,v)xp[u](x[v]), U×UC(I;Rn), is continuous as a composition of continuous functions. Then the operator (u,v)E[u,v], U×UR, is also continuous as a composition of continuous operators.

By the definition of the residual E the equality E[uk,uk+1]=I[uk+1]I[uk] holds for every k. The sequence of numbers {I[uk]} is monotone and bounded, hence convergent. Consequently, limkE[uk,uk+1]=0. Then the subsequence ωkj{(u2kj,u2kj+1)}ωk having the limit (u,v) satisfies the equality 0=limjE[u2kj,u2kj+1]=E[u,v].

The proof of the proposition is complete.

The output of the method proposed above is a feedback (in a particular case, Pontryagin’s) extremal or a sequence of controls lying on the level set of the functional I corresponding to such an extremal (the case of ‘looping’). Its implementation with the use of the Krasovskii–Subbotin scheme produces an algorithm of dynamical optimization,11 which, in contrast to indirect algorithms based on the PMP [8], does not contain parameters of the ‘descent depth’12 (hence no internal procedures for linear search). Of course, in practice this algorithm is applicable to the general problem (P) (without assumptions (A4)), as well as to the convexified problem (coP), without any modification. In the latter problem, to approximate sliding modes with property (7.5) one can use feedback controls of the form

wλ(t,x)(1λ)¯u(t)+λw(t,x),λ(0,1],
where w obeys inclusion (6.2). This is consistent with the informal idea of the ‘locally optimal’ synthesis of control (see [23]).

§ 10. Appendix: auxiliary assertions

Below Lip(φ;A) is the minimum Lipschitz constant of the function φ:RnR on the set ARn and Lip(φ)Lip(φ;Rn); Br is the closed ball of radius r centred at the origin.

Lemma 1. Suppose that assumptions (A1)(A3) are fulfilled. Then the following hold.

(1) For any compact set KRn the attainability set {Xs,t[ν](x)s,tI,xK,νY} of system (7.1) from the set K on the time interval [s,t] is contained in a closed ball BRf whose radius Rf=Rf(K,U)>0 depends only on Cf, K and U (and is the same for all s and t).

(2) For any compact set KRn and any s,tI, xK and νY the following estimate is valid:

|Js,t[ν](x)|exp{max(x,υ)BRf(K,U)×U|Dxfυ(x)|}CJ(K,U).

(3) For any compact set KRn the functions tXs,t[ν](x), tXt,s[ν](x) and tJt,s[ν](x) are Lipschitz continuous on I with Lipschitz constants LX(K,U) and LJ(K,U) which are common to all sI, xK and νY.

(4) The functions xXt,s[ν](x) and xJt,s(x)DxXt,s[ν](x) are locally Lipschitz continuous on Rn uniformly in s,tI and νY.

Proof. We restrict our consideration to the proof of the uniform local Lipschitz continuity of the family Jt,s[ν], s,tI, νY. All other facts are well known in the theory of ordinary differential equations.

Let x,zKRn, where K is a compact set; suppose for definiteness that s>t. It follows from (3.4) that

|Jt,s(x)Jt,s(z)|=|[t,s]×U[Jτ,s(x)DxfυXτ,s(x)Jτ,s(z)DxfυXτ,s(z)]dν|[t,s]×U|Jτ,s(x)||DxfυXτ,s(x)DxfυXτ,s(z)|dν(τ,υ)+[t,s]×U|Jτ,s(x)Jτ,s(z)||DxfυXτ,s(z)|dν(τ,υ)CJ(K)maxυULip(Dxfυ;BRf(K))|xz|+max(x,υ)BRf(K)×U|Dxfυ(x)|st|Jτ,s(x)Jτ,s(z)|dτ.
Then Grönwall’s lemma suggests the estimate
|Jt,s(x)Jt,s(z)|Mf(K;t,s)|xz|Mf(K;0,1)|xz|,Mf(K;t,s)CJ(K)maxυULip(Dxfυ;BRf(K))×exp{max(x,υ)BRf(K)×U|Dxfυ(x)|(st)}.

The proof of the lemma is complete.

Lemma 2. Suppose that conditions (A1)(A3) hold, ¯uU and KRn is a compact set. Then the restriction of the family of functions (t,x)¯Ht(x,υ), υU, to the set I×K is a Lipschitz continuous function whose Lipschitz constant depends only on Cf, Lip(;K), LX(K,U), LJ(K,U), CJ(K,U), Rf(K,U) and Lf(K,U), where Lf(K,U) is a common Lipschitz constant for the mappings xf(x,υ), υU, on the set K and LX, LJ, CJ and Rf were defined in Lemma 1.

Proof. In view of Lemma 1 the claim of the lemma follows from the representation x¯pt=¯Jt¯Xt,1 and the following estimates, which are valid for any s,tI, x,zK and υU:
|¯Ht(x,υ)¯Hs(x,υ)||x¯pt(x)f(x,υ)x¯ps(x)f(x,υ)||x¯pt(x)x¯ps(x)||f(x,υ)||¯Jt(x)¯Xt,1(x)¯Js(x)¯Xs,1||f(x,υ)||¯Jt(x)¯Js(x)||¯Xs,1||f(x,υ)|+|¯Jt(x)||¯Xt,1(x)¯Xs,1(x)||f(x,υ)|LJ(K)|ts|Lip(;BRf(K,U))maxK×U|f|+CJ(K,U)Lip(;BRf(K,U))LX(K,U)|ts|maxK×U|f|Lt¯H|ts|,|¯Ht(x,υ)¯Ht(z,υ)||x¯pt(x)f(x,υ)x¯pt(z)f(z,υ)||x¯pt(x)||f(x,υ)f(z,υ)|+|x¯pt(x)x¯pt(z)||f(z,υ)|Lf(K,U)|x(¯Xt,1(x))||¯Jt(x)||xz|+|x(¯Xt,1(x))||¯Jt(x)¯Jt(z)||f(z,υ)|+|x(¯Xt,1(x))x(¯Xt,1(z))||¯Jt(z)||f(z,υ)|Lf(K,U)Lip(;BRf(K,U))CJ(K,U)|xz|+Lip(;BRf(K,U))LJ(K,U)|xz|maxK×U|f|+Lip(;BRf(K,U))|xz|CJ(K,U)maxK×U|f|Lx¯H|xz|,
where the Lipschitz constants are defined by the expressions
Lt¯Hmax{LJ(K)Lip(;BRf(K,U));CJ(K,U)Lip(;BRf(K,U))LX(K,U)}×maxI×K|f|
and
Lx¯Hmax{Lf(K,U)Lip(;BRf(K,U))CJ(K,U);Lip(;BRf(K,U))LJ(K,U)×maxK×U|f|;Lip(;BRf(K))CJ(K,U)maxK×U|f|}.

The proof of the lemma is complete.

§ 11. Appendix: feedback controls and sliding modes

This section contains a brief review of the main facts about feedback controls and their relation to sliding modes. A feedback control for system (1.1) can be an arbitrary function (t,x)w(t,x), I×RnU. In general, this function is not even supposed to be measurable, which makes impossible its direct substitution (for u) into (1.1). The action of w on the system can be defined by use of various ‘sampling’ schemes, the most widely known (and simplest) of which is the Krasovskii–Subbotin scheme.

Definition 2. Let w:I×RnU be a given function. Consider the sequence {πN}I of partitions of the interval I by points

{tNi}Ni=0,0tN0<tNi<<tNN1<tNN1,N1,
with the properties πN+1πN (the sequence {πN} is nondecreasing by inclusion) and
limNmin1iN|tNitNi1|=0.
We define the sequence of ‘Euler’s polygons’ iteratively: xN(t)=x[w(tNi,xN(tNi))](t), t[tNi,tNi+1), i=0,,N1, and define xN at the point t=1 by continuity. Let {xNs}{xN} be a subsequence of the sequence NxN that has a uniform limit x on I as s. All such partial limits x=x[w] are called Krasovskii–Subbotin solutions of the closed equation (2.9).

By construction each of the indicated solutions is contained in the tube of trajectories of the convexified system (7.1), that is, presents a sliding mode. In other words, x[w]=x[ν] for some νY.

Suppose that assumptions (A2) and (A3) hold, and let KS[w] be the system of all Krasovskii–Subbotin solutions corresponding to w. It is clear that any subset of KS[w] is relatively compact in C(I;Rn), since all the trajectories of sliding modes are uniformly bounded and Lipschitz continuous with a common constant, and therefore satisfy the hypotheses of the Arzelà–Ascoli theorem. This yields the existence of (at least one) solution x[w] for each w. This solution is not unique even in simplest cases (see [11]).

If the operator ux[u] is continuous on U, then among the generalized controls ν that generate the curve x[w] as a trajectory of a sliding mode there is a measure with the property νt=δu(t) for almost all tI, uU (see [24], Corollary 1), which yields the following result.

Proposition 6. Suppose that assumptions (A2)(A4) hold, and let w:I×RnU be an arbitrary feedback control and x=x[w] be one of the Krasovskii–Subbotin solutions of system (2.9) generated by it. Then there exists a program control uU such that x[u]=x[w].

If the piecewise constant strategies in the Krasovskii–Subbotin scheme are chosen in accordance with some (sufficiently regular) rule, then it is natural to expect that the last assertion also holds for the limiting sliding mode.

Proposition 7. Let

gC(I×Rn×U;R)
be a given nonnegative function such that the mapping (t,x)g(t,x,υ) is locally Lipschitz continuous uniformly in υU. Further, let w:I×RnU be an arbitrary function, {πN}I be a sequence of partitions (11.1) of the interval I which is nondecreasing by inclusion, {uN} be a sequence of piecewise constant controls of the form
uN(t)uN,iU for tΔtNi(tNi1,tNi),i=1,,N,N1,
and {xNx[uN]} be the corresponding sequence of Euler polygons. Finally, let ν be a partial limit of the sequence (νN) of Young measures for which νNt=δuN(t) and x=x[ν] be the corresponding partial uniform limit of the sequence {xN}. Suppose that the condition
g(tNi,xN(tNi),uN,i)=0tNiπN
holds for every N1. Then
I×Ug(t,x(t),υ)dν(t,υ)=0.

Proof. For the sake of simplicity we restrict our consideration to uniform partitions of the interval I by the points tNi=i/N, i=0,,N. It follows from the continuity of the function g, disintegration theorem and the definition of a Young measure that
I×Ug(t,x(t),υ)dν(t,υ)IdtUg(t,x(t),υ)dνt(υ)=limNIdtUg(t,xN(t),υ)dνNt(υ).
Consider the expression under the integral sign. By construction we have
IdtUg(t,xN(t),u)dνNt(u)Ni=1clΔtNidtUg(t,xN(t),u)dδuN,i=Ni=1clΔtNig(t,xN(t),uN,i)dt.
Let KRn be a compact set containing the tube of trajectories (1.1), (1.2), and Lg(K) be the minimal common Lipschitz constant for the functions (t,x)g(t,x,υ), υU, on I×K. Each term in the last sum can be estimated as follows:
|clΔtNig(t,xN(t),uN,i)dt|clΔtNi|g(t,xN(t),uN,i)g(tNi,xN(tNi),uN,i)|dtLg(K)((tNitNi1)22+clΔtNi|xN(t)xN(tNi)|dt)Lg(K)(1+LX(K,U))(tNitNi1)22=Lg(K)(1+LX(K,U))2N2M(K,U)N2.
Here the constant LX(K,U) is the same as in Lemma 1. Note that M does not depend on N. Now recalling that g0 and summing over i=1,,N we obtain the estimate
0IdtUg(t,xN(t),υ)dνNt(υ)O(1N).
Finally, taking the limit as N completes the proof.

In conclusion we pay special attention to one detail concerned with the use of program controls as feedback ones. Each representative of the class uU can formally be interpreted as a feedback. Then different realizations of u — functions which differ from one another on a set of Lebesgue measure zero — generate, in general, different sets of Krasovskii–Subbotin solutions (since the set of functions w is subject to no factorization). This effect can be avoided by adopting a more accurate rule of selection, for example, by restricting the set of values u(tk) at points of the partition (11.1) to elements of the closure (the suitable one-sided closure at the endpoints) of the function u with respect to the Lebesgue measure [25].

Another way consists in abandoning the Krasovskii–Subbotin scheme in favour of an alternative sampling algorithm in which piecewise constant strategies for constructing the polygons are replaced by ‘piecewise program’ strategies (see, for example, [26]).

§ 12. Conclusion

This work presents a new approach to the theory of local extremum in problems of optimal control, which is alternative (but related) to Pontryagin’s principle. Advantages of the approach developed include, along with improving the classical necessary condition and a natural algorithmization, rather simple proofs of the main results. Further research will be devoted to aspects of the practical implementation of the descent method proposed here. The results obtained can easily (as we believe) be carried over to some classes of problems of optimal stochastic control and mean-field control.

Acknowledgments

The authors are grateful to V. A. Dykhta for the discussions of this work and a series of valuable suggestions that he made and to anonymous referees, whose constructive observations contributed to an essential improvement of the original version of this paper.


Bibliography

1. V. A. Dykhta, “Weakly monotone solutions of the Hamilton–Jacobi inequality and optimality conditions with feedback controls”, Autom. Remote Control, 75:5 (2014), 829–844  mathnet  crossref  mathscinet  zmath
2. V. A. Dykhta, “Variational necessary optimality conditions with feedback descent controls for optimal control problems”, Dokl. Math., 91:3 (2015), 394–396  crossref  crossref
3. V. A. Dykhta, “Feedback minimum principle: variational strengthening of the concept of extremality in optimal control”, Izv. Irkutsk. Gos. Univ. Ser. Mat., 41 (2022), 19–39 (Russian)  mathnet  crossref  mathscinet  zmath
4. V. F. Krotov, “Global methods to improve control and optimal control of resonance interaction of light and matter”, Modeling and control of systems in engineering, quantum mechanics, economics and biosciences (Sophia–Antipolis 1988), Lect. Notes Control Inf. Sci., 121, Springer, Berlin, 1989, 267–298  crossref  mathscinet
5. L. S. Pontryagin, V. G. Boltyanskii, R. V. Gamkrelidze and E. F. Mishchenko, The mathematical theory of optimal processes, Intersci. Publ. John Wiley & Sons, Inc., New York–London, 1962, viii+360 pp.  mathscinet  zmath
6. A. Ya. Dubovitskii and A. A. Milyutin, “Extremum problems in the presence of restrictions”, U.S.S.R. Comput. Math. Math. Phys., 5:3 (1965), 1–80  mathnet  crossref  mathscinet  zmath
7. R. J. DiPerna and P. L. Lions, “Ordinary differential equations, transport theory and Sobolev spaces”, Invent. Math., 98:3 (1989), 511–547  crossref  mathscinet  zmath  adsnasa
8. V. A. Srochko, Iterative methods of solution of optimal control problems, Fizmatlit, Moscow, 2000, 160 pp. (Russian)
9. L. Ambrosio and G. Savaré, “Gradient flows of probability measures”, Handbook of differential equations: evolutionary equations, v. III, Handb. Differ. Equ., Elsevier/North-Holland, Amsterdam, 2007, 1–136  crossref  mathscinet  zmath
10. V. I. Bogachev, Weak convergence of measures, Math. Surveys Monogr., 234, Amer. Math. Soc., Providence, RI, 2018, xii+286 pp.  crossref  mathscinet  zmath
11. N. N. Krasovskiĭ and A. I. Subbotin, Game-theoretical control problems, Springer Ser. Soviet Math., Springer-Verlag, New York, 1988, xii+517 pp.  mathscinet  zmath
12. M. A. Lavrentyev and L. A. Lyusternik, A course of variational calculus, GONTI–NKTI, Moscow–Leningrad, 1938, 192 pp. (Russian)
13. A. Bressan and B. Piccoli, Introduction to the mathematical theory of control, AIMS Ser. Appl. Math., 2, Amer. Inst. Math. Sci. (AIMS), Springfield, MO, 2007, xiv+312 pp.  mathscinet  zmath
14. N. Pogodaev, “Program strategies for a dynamic game in the space of measures”, Optim. Lett., 13:8 (2019), 1913–1925  crossref  mathscinet  zmath
15. N. Pogodaev and M. Staritsyn, “Impulsive control of nonlocal transport equations”, J. Differential Equations, 269:4 (2020), 3585–3623  crossref  mathscinet  zmath  adsnasa
16. A. A. Agrachev and Yu. L. Sachkov, Control theory from the geometric viewpoint, Encyclopaedia Math. Sci., 87, Control Theory Optim., II, Springer-Verlag, Berlin, 2004, xiv+412 pp.  crossref  mathscinet  zmath
17. R. J. Kipka and Yu. S. Ledyaev, “Extension of chronological calculus for dynamical systems on manifolds”, J. Differential Equations, 258:5 (2015), 1765–1790  crossref  mathscinet  zmath
18. R. Vinter, “Convex duality and nonlinear optimal control”, SIAM J. Control Optim., 31:2 (1993), 518–538  crossref  mathscinet  zmath
19. F. H. Clarke and C. Nour, “Nonconvex duality in optimal control”, SIAM J. Control Optim., 43:6 (2005), 2036–2048  crossref  mathscinet  zmath
20. V. A. Dykhta, “Nonstandard duality and nonlocal necessary optimality conditions in nonconvex optimal control problems”, Autom. Remote Control, 75:11 (2014), 1906–1921  mathnet  crossref  mathscinet  zmath
21. M. Staritsyn, N. Pogodaev, R. Chertovskih and F. Lobo Pereira, “Feedback maximum principle for ensemble control of local continuity equations: an application to supervised machine learning”, IEEE Control Syst. Lett., 6 (2022), 1046–1051  crossref  mathscinet
22. C. Castaing, P. Raynaud de Fitte and M. Valadier, Young measures on topological spaces. With applications in control theory and probability theory, Math. Appl., 571, Kluwer Acad. Publ., Dordrecht, 2004, xii+320 pp.  crossref  mathscinet  zmath
23. V. I. Gurman, The extension principle in control problems, 2nd revised and augmented ed., Fizmatlit, Moscow, 1997, 288 pp. (Russian)  mathscinet  zmath
24. N. Pogodaev, “Optimal control of continuity equations”, NoDEA Nonlinear Differential Equations Appl., 23:2 (2016), 21, 24 pp.  crossref  mathscinet  zmath
25. A. V. Arutyunov, D. Yu. Karamzin and F. L. Pereira, “Conditions for the absence of jumps of the solution to the adjoint system of the maximum principle for optimal control problems with state constraints”, Proc. Steklov Inst. Math. (Suppl.), 292, suppl. 1 (2016), 27–35  mathnet  crossref  mathscinet  zmath
26. M. Staritsyn and S. Sorokin, “On feedback strengthening of the maximum principle for measure differential equations”, J. Global Optim., 76:3 (2020), 587–612  crossref  mathscinet  zmath

Citation: N. I. Pogodaev, M. V. Staritsyn, “Exact formulae for the increment of the objective functional and necessary optimality conditions, alternative to Pontryagin's maximum principle”, Sb. Math., 215:6 (2024), 790–822
Citation in format AMSBIB
\Bibitem{PogSta24}
\by N.~I.~Pogodaev, M.~V.~Staritsyn
\paper Exact formulae for the increment of the objective functional and necessary optimality conditions, alternative to Pontryagin's maximum principle
\jour Sb. Math.
\yr 2024
\vol 215
\issue 6
\pages 790--822
\mathnet{http://mi.mathnet.ru/eng/sm9967}
\crossref{https://doi.org/10.4213/sm9967e}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4804039}
\zmath{https://zbmath.org/?q=an:07945696}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2024SbMat.215..790P}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001334620600005}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85206914941}
Linking options:
  • https://www.mathnet.ru/eng/sm9967
  • https://doi.org/10.4213/sm9967e
  • https://www.mathnet.ru/eng/sm/v215/i6/p77
  • This publication is cited in the following 1 articles:
    1. E. V. Goncharova, N. I. Pogodaev, M. V. Staritsyn, “Tochnye formuly prirascheniya tselevogo funktsionala v zadache optimalnogo upravleniya lineinym uravneniem balansa”, Izvestiya Irkutskogo gosudarstvennogo universiteta. Seriya Matematika, 51 (2025), 3–20  mathnet  crossref
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:271
    Russian version PDF:5
    English version PDF:28
    Russian version HTML:31
    English version HTML:111
    References:35
    First page:14
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2025