22.02.2015 Views

Linear Programming Approach to Deterministic Infinite Horizon ...

Linear Programming Approach to Deterministic Infinite Horizon ...

Linear Programming Approach to Deterministic Infinite Horizon ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

SIAM J. CONTROL OPTIM.<br />

Vol. 48, No. 4, pp. 2480–2512<br />

c○ 2009 Society for Industrial and Applied Mathematics<br />

LINEAR PROGRAMMING APPROACH TO DETERMINISTIC<br />

INFINITE HORIZON OPTIMAL CONTROL PROBLEMS WITH<br />

DISCOUNTING ∗<br />

VLADIMIR GAITSGORY † AND MARC QUINCAMPOIX ‡<br />

Abstract. We investigate relationships between the deterministic infinite time horizon optimal<br />

control problem with discounting, in which the state trajec<strong>to</strong>ries remain in a given compact set Y ,<br />

and a certain infinite dimensional linear programming (IDLP) problem. We introduce the problem<br />

dual with respect <strong>to</strong> this IDLP problem and obtain some duality results. We construct necessary<br />

and sufficient optimality conditions for the optimal control problem under consideration, and we give<br />

a characterization of the viability kernel of Y . We also indicate how one can use finite dimensional<br />

approximations of the IDLP problem and its dual for construction of near optimal feedback controls.<br />

The construction is illustrated with a numerical example.<br />

Key words. optimal control problems with discounting, long run average optimal control,<br />

occupational measures, averaging, linear programming, duality, viability kernels, numerical solution<br />

AMS subject classifications. 34E15, 34C29, 34A60, 93C70<br />

DOI. 10.1137/070696209<br />

1. Introduction and preliminaries. It is well known that the dynamics of a<br />

nonlinear s<strong>to</strong>chastic control system has a linear representation through the dynamics<br />

of the corresponding state-control probability distributions. A different (but related)<br />

idea of “linearizing” nonlinear optimal control problems can be realized through reformulating<br />

these as optimization problems on spaces of occupational measures, which,<br />

under mild conditions, can be shown <strong>to</strong> be “equivalent” (or “asymp<strong>to</strong>tically equivalent”)<br />

<strong>to</strong> certain infinite dimensional (ID) linear programming (LP) problems. This<br />

idea is applicable <strong>to</strong> both s<strong>to</strong>chastic and deterministic settings. It is based on the fact<br />

that the occupational measures generated by admissible controls and the corresponding<br />

solutions of a nonlinear system satisfy certain linear equations representing the<br />

system’s dynamics in a relaxed integral form.<br />

Fundamental results that justify the use of LP formulations in various problems of<br />

optimal control of s<strong>to</strong>chastic systems have been obtained in [11], [24], [34], [39], [54],<br />

[55]. Important advances in the development of IDLP formulations in deterministic<br />

optimal control problems considered on finite time intervals have been made in [35],<br />

[42], [51], [58] (and in some earlier papers mentioned therein). Also various aspects of<br />

the LP approach <strong>to</strong> deterministic problems of optimal control with long run average<br />

criteria were studied in [22] and [30] (important related developments can be found<br />

in [21] and [31]).<br />

This paper is devoted <strong>to</strong> the development of the LP approach <strong>to</strong> the deterministic<br />

infinite horizon optimal control problem with discounting, in which the state trajec-<br />

∗ Received by the edi<strong>to</strong>rs July 4, 2007; accepted for publication (in revised form) May 15, 2009;<br />

published electronically August 14, 2009.<br />

http://www.siam.org/journals/sicon/48-4/69620.html<br />

† Center for Industrial and Applied Mathematics, University of South Australia, Mawson Lakes,<br />

SA 5095, Australia (v.gaitsgory@unisa.edu.au). This author’s work was partially supported by the<br />

Australian Research Council Discovery grants DP0664330 and DP0986696, and by Linkage International<br />

grants LX0560049 and LX0881972.<br />

‡ Labora<strong>to</strong>iredeMathématiques, unité CNRS UMR6205, Université de Bretagne Occidentale, 6<br />

Avenue Vic<strong>to</strong>r Le Gorgeu, 29200 Brest, France (marc.quincampoix@univ-brest.fr). This author’s<br />

work was partially supported by Linkage International grant LX0560049.<br />

2480<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2481<br />

<strong>to</strong>ries remain in a compact set Y (a state constraint). We establish that the optimal<br />

value of this problem coincides with the optimal value of a certain IDLP problem,<br />

the feasible set of the latter coinciding with the convex hull of the set of discounted<br />

occupational measures generated by the control system, and we show that this IDLP<br />

problem and its dual can be used for construction of necessary and sufficient optimality<br />

conditions and for a characterization of the viability kernel of Y . We also indicate<br />

a way of how one can use finite dimensional approximations of the IDLP problem and<br />

its dual for construction of near optimal feedback controls (the construction being<br />

illustrated with a numerical example).<br />

Note that infinite horizon problems of optimal control arise in many applications<br />

(in engineering, economics, environmental modeling, etc.). They have been studied<br />

intensively(see,e.g.,[1],[3],[4],[7],[8],[10],[11],[15],[17],[20],[23],[24],[32],<br />

[34], [36], [39], [40], [41], [44], [47], [50], [56], [60]), and the present paper aims at<br />

contributing <strong>to</strong> this important line of research.<br />

The control system we will be dealing with is of the form<br />

(1.1) y ′ (t) =f(y(t),u(t)), t≥ 0,<br />

where the function f(·) :R m × U ↦→ R m is continuous in (y, u) and satisfies Lipschitz<br />

conditions in y uniformly with respect <strong>to</strong> u. The controls are Lebesgue measurable<br />

functions u(·) :[0,S] ↦→ U or u(·) :[0, +∞) ↦→ U (depending on whether the system<br />

is considered on the finite time interval [0,S] or on the infinite time interval [0, +∞)),<br />

where U is a compact metric space. The sets of these controls are denoted as U S and<br />

U, respectively. A solution of the system (1.1) obtained with a control u(·) andwith<br />

the initial condition y(0) = y 0 will be denoted as y(t, y 0 ,u(·)).<br />

Let Y be a nonempty compact subset of R m . We will be considering the solutions<br />

of the system (1.1), which satisfy the state constraint<br />

(1.2) y(t, y 0 ,u(·)) ∈ Y,<br />

and we will denote by US Y (y 0) ⊂U S and by U Y (y 0 ) ⊂U the sets of controls such<br />

that (1.2) is satisfied for all t ∈ [0,S]andforallt ∈ [0, ∞), respectively. Note that<br />

the set Y is called viable if U Y (y 0 ) ≠ ∅ for all y 0 ∈ Y (see [5]).<br />

Let us consider the optimal control problem<br />

(1.3) inf<br />

u(·)∈U Y (y 0)<br />

∫ +∞<br />

0<br />

e −Ct g(y(t, y 0 ,u(·)),u(t))dt def<br />

= V Y C (y 0),<br />

where g : R m × U ↦→ R is continuous and satisfies Lipschitz conditions in y uniformly<br />

with respect <strong>to</strong> u and where C>0 is the discount fac<strong>to</strong>r. We will be interested in<br />

establishing connections between this problem and the problem<br />

∫<br />

(1.4) inf<br />

γ∈W (C,y 0)<br />

where<br />

(1.5)<br />

W (C, y 0 ) def<br />

=<br />

{<br />

γ ∈P(Y × U) :<br />

Y ×U<br />

∫<br />

Y ×U<br />

g(y, u)γ(dy, du) def<br />

= g ∗ (C, y 0 ),<br />

(∇φ(y) T f(y, u)+C(φ(y 0 ) − φ(y)))γ(dy, du)<br />

=0 ∀φ ∈ C 1 },<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2482 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

with P(Y ×U) standing for the space of probability measures defined on Borel subsets<br />

of Y × U. Note that problem (1.4) is of IDLP since its objective function and its constraints<br />

are linear in γ (see, e.g., [2]). Note that the key element of our consideration<br />

is the fact that the discounted occupational measures generated by the solutions of<br />

system (1.1) satisfy the constraints defining W (C, y 0 ) (see Proposition 2.2).<br />

Along with (1.3), we will also be considering the optimal control problem<br />

(1.6)<br />

∫<br />

1<br />

S<br />

S inf inf g(y(t, y 0 ,u(·)),u(t))dt def<br />

= G S .<br />

y 0∈Y u(·)∈US Y (y0) 0<br />

In [22] and [30] it has been established that this problem (considered with S →∞)<br />

is closely related <strong>to</strong> the IDLP problem<br />

∫<br />

(1.7) inf g(y, u)γ(dy, du) def<br />

= g ∗ ,<br />

γ∈W<br />

where<br />

(1.8) W def<br />

=<br />

{<br />

γ ∈P(Y × U) :<br />

Y ×U<br />

∫<br />

Y ×U<br />

∇φ(y) T f(y, u)γ(dy, du) =0 ∀φ ∈ C 1 }.<br />

The argument used in [22] and [30] was based on results from s<strong>to</strong>chastic control<br />

theory (mainly, on the fundamental result characterizing the set of one-dimensional<br />

marginal stationary distributions of a control martingale problem obtained in [54];<br />

see also [11] and [39]). We will show that some of the results of [22] and [30] that<br />

establish relationships between (1.6) and (1.7) can be obtained on the basis of results<br />

establishing relationships between (1.3) and (1.4) (by letting the discount fac<strong>to</strong>r tend<br />

<strong>to</strong> zero).<br />

The rest of the paper is organized as follows. In section 2, we reformulate the<br />

optimal control problems (1.3) and (1.6) in terms of occupational measures, and we<br />

prove some preliminary results relating these problems with the IDLP problems (1.4)<br />

and (1.7). In section 3, we introduce the problem dual <strong>to</strong> (1.4) and obtain duality<br />

results. In section 4, we use results of section 3 <strong>to</strong> establish the relationships between<br />

(1.3) and (1.4), and we obtain necessary and sufficient optimality conditions for (1.3).<br />

In section 5, we demonstrate the possibility of applying results of section 4 for a<br />

characterization of the viability kernel of Y . In section 6, we use results of section<br />

4 <strong>to</strong> establish the relationships between (1.6) and (1.7). In section 7, we discuss the<br />

possibility of approximation of the IDLP problem and its dual with finite dimensional<br />

LP problems, and we illustrate that the latter can be used for finding a near optimal<br />

control in (1.3) with a numerical example. In section 8, we make some conclusions<br />

about the obtained results. In the appendix, we give some proofs that were omitted<br />

in the previous consideration.<br />

Let us conclude this section with some comments and notation. Note, first of<br />

all, that the space P(Y × U) is known <strong>to</strong> be compact in weak convergence (weak ∗ )<br />

<strong>to</strong>pology (see, e.g., [9] or [46]). Hence, the sets W and W (C, y 0 )arecompactinthis<br />

<strong>to</strong>pology, and a solution <strong>to</strong> problem (1.7) or problem (1.4) exists as soon as W or,<br />

respectively, W (C, y 0 ), is not empty.<br />

Let us endow the space P(Y × U) with a metric ρ,<br />

(1.9) ρ(γ ′ ,γ ′′ ) def<br />

=<br />

∞∑<br />

j=1<br />

∣ ∫<br />

1 ∣∣∣<br />

2<br />

∫U×Y<br />

j q j (y, u)γ ′ (dy, du) −<br />

U×Y<br />

q j (y, u)γ ′′ (dy, du)<br />

∣ ,<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2483<br />

for all γ ′ ,γ ′′ ∈P(Y ×U), where q j (·),j =1, 2,...,is a sequence of Lipschitz continuous<br />

functions which is dense in the unit ball of C(Y ×U) (the space of continuous functions<br />

on Y × U). Note that this metric is consistent with the weak convergence <strong>to</strong>pology<br />

of P(Y × U). Namely, a sequence γ k ∈P(Y × U) converges<strong>to</strong>γ ∈P(Y × U) inthis<br />

metric if and only if<br />

∫<br />

∫<br />

(1.10) lim q(y, u)γ k (dy, du) = q(y, u)γ(dy, du)<br />

k→∞ U×Y<br />

U×Y<br />

for any continuous q(·) ∈ C(Y ×U). Using this metric ρ, one can define the “distance”<br />

ρ(γ,Γ) between γ ∈P(Y ×U)andΓ⊂P(Y ×U), and the Hausdorff metric ρ H (Γ 1 , Γ 2 )<br />

between Γ 1 ⊂P(Y × U) andΓ 2 ⊂P(Y × U), as follows:<br />

(1.11)<br />

ρ(γ,Γ) def<br />

= inf<br />

γ ′ ∈Γ ρ(γ,γ′ ) ,<br />

{<br />

}<br />

ρ H (Γ 1 , Γ 2 ) def<br />

=max sup ρ(γ,Γ 2 ), sup ρ(γ,Γ 1 ) .<br />

γ∈Γ 1 γ∈Γ 2<br />

Note that, although, by some abuse of terminology, we refer <strong>to</strong> ρ H (·, ·) asametric<br />

on the set of subsets of P(Y × U), it is, in fact, a semimetric on this set (since<br />

ρ H (Γ 1 , Γ 2 )=0isequivalent<strong>to</strong>Γ 1 =Γ 2 if and only if Γ 1 and Γ 2 are closed).<br />

2. Occupational measure formulations. In this section we introduce occupational<br />

and discounted occupational measures. We reformulate the optimal control<br />

problems (1.3) and (1.6) in terms of these measures, and we establish some readily<br />

verifiable relationships between (1.3) and (1.4) and between (1.6) and (1.7). The<br />

results of this section are used in the further consideration.<br />

Let u(·) ∈US Y (y 0)andy(t) =y(t, y 0 ,u(·)), t ∈ [0,S]. A probability measure<br />

γ u(·),S ∈P(Y ×U) is called the occupational measure generated by the pair (y(·),u(·))<br />

on the interval [0,S] if, for any Borel set Q ⊂ Y × U,<br />

(2.1) γ u(·),S (Q) = 1 S<br />

∫ S<br />

0<br />

1 Q (y(t),u(t))dt,<br />

where 1 Q (·) is the indica<strong>to</strong>r function of Q. This definition is equivalent <strong>to</strong> the statement<br />

that the equality<br />

(2.2)<br />

∫<br />

Y ×U<br />

q(y, u)γ u(·),S (dy, du) = 1 S<br />

∫ S<br />

0<br />

q(y(t),u(t))dt<br />

is valid for any q(·) ∈ C(Y × U).<br />

Let u(·) ∈U Y (y 0 )andy(t) =y(t, y 0 ,u(·)), t∈ [0, ∞). The pair (y(·),u(·)) is said<br />

<strong>to</strong> generate an occupational measure on the interval [0, ∞) if there exists a limit<br />

(2.3) lim γ def<br />

u(·),S = γ u(·) .<br />

S→∞<br />

Note that γ u(·) is generated by (y(·),u(·)) on [0, ∞) if and only if<br />

(2.4)<br />

∫<br />

Y ×U<br />

for any q(·) ∈ C(Y × U).<br />

∫<br />

1 S<br />

q(y, u)γ u(·) (dy, du) = lim q(y(t),u(t))dt<br />

S→∞ S 0<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2484 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

Let u(·) ∈U Y (y 0 )andy(t) =y(t, y 0 ,u(·)), t ∈ [0, ∞). A probability measure<br />

γu(·) C ∈P(Y × U) is called the discounted occupational measure generated by the pair<br />

(y(·),u(·)) if, for any Borel set Q ⊂ Y × U,<br />

∫ ∞<br />

(2.5) γu(·) C (Q) =C e −Ct 1 Q (y(t),u(t))dt,<br />

where the latter definition is equivalent <strong>to</strong> the equality<br />

∫<br />

∫ ∞<br />

(2.6)<br />

q(y, u)γu(·) C (dy, du) =C e −Ct q(y(t),u(t))dt<br />

Y ×U<br />

0<br />

being valid for any q(·) ∈ C(Y × U).<br />

Proposition 2.1. If γ u(·) is generated by the pair (y(·),u(·)) on [0, ∞), then<br />

(2.7) lim ρ(γu(·) C ,γ C→0<br />

u(·)) =0.<br />

Proof. From the fact that γ u(·) is generated by the pair (y(·),u(·)) on [0, ∞), it<br />

follows that the limit in the right-hand side of (2.4) exists for any q(·) ∈ C(Y × U).<br />

Hence, by the Abelian theorem (see, e.g., Lemma 3.5(i) in [33]),<br />

∫ ∞<br />

∫<br />

lim C e −Ct 1 S<br />

q(y(t),u(t))dt = lim q(y(t),u(t))dt<br />

C→0 0<br />

S→∞ S 0<br />

∫<br />

⇒ lim q(y, u)γu(·) C→0<br />

∫Y C (dy, du) = q(y, u)γ u(·) (dy, du).<br />

×U<br />

Y ×U<br />

The validity of the latter for any q(·) ∈ C(Y × U) is equivalent <strong>to</strong> (2.7).<br />

Let us introduce the following notation:<br />

(2.8) Γ S (y 0 ) def<br />

=<br />

⋃<br />

u(·)∈U Y S (y0) {γ u(·),S },<br />

(2.9) Γ(C, y 0 ) def<br />

=<br />

⋃<br />

0<br />

u(·)∈U Y (y 0)<br />

Γ S<br />

def<br />

= ⋃<br />

{γ C u(·) },<br />

y 0∈Y<br />

{Γ S (y 0 )},<br />

with Γ S (y 0 ) def<br />

= ∅ if US Y (y 0)=∅ and Γ(C, y 0 ) def<br />

= ∅ if U Y (y 0 )=∅. Due <strong>to</strong> (2.6)<br />

and (2.2), respectively, problems (1.3) and (1.6) can be rewritten in this notation as,<br />

respectively,<br />

∫<br />

(2.10) inf g(y, u)γ(dy, du) =CVC Y (y 0 )<br />

γ∈Γ(C,y 0)<br />

and<br />

∫<br />

(2.11) inf<br />

γ∈Γ S<br />

g(y, u)γ(dy, du) =G S .<br />

Note that these problems are not of LP (since Γ(C, y 0 )andΓ S are not defined by<br />

linear constraints), and our immediate aim is <strong>to</strong> relate them <strong>to</strong> IDLP problems (1.4)<br />

and (1.7).<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2485<br />

Proposition 2.2. The following relationships are valid:<br />

(2.12) CV Y C (y 0 ) ≥ g ∗ (C, y 0 );<br />

(2.13) ¯coΓ(C, y 0 ) ⊂ W (C, y 0 ),<br />

where co ¯ stands for the closed convex hull.<br />

Proof. Take arbitrary γ ∈ Γ(C, y 0 ). By definition, there exists u(·) ∈U Y (y 0 )<br />

and y(t) def<br />

= y(t, y 0 ,u(·)) such that γ = γu(·) C (that is, γ is the discounted occupational<br />

measure generated by the pair (y(·),u(·))). Using the fact that (2.6) is valid for any<br />

continuous function q(y, u), one can obtain<br />

∫<br />

∫ ∞<br />

∇φ(y) T f(y, u)γu(·) C (dy, du) =C e −Ct ∇φ(y(t)) T f(y(t),u(t))dt = −Cφ(y 0 )<br />

Y ×U<br />

0<br />

+ C 2 ∫ ∞<br />

0<br />

∫<br />

e −Ct φ(y(t))dt = −C (φ(y 0 ) − φ(y))γu(·) C (dy, du)<br />

Y ×U<br />

∀φ ∈ C1<br />

⇒ γ = γ C u(·) ∈ W (C, y 0) ⇒ Γ(C, y 0 ) ⊂ W (C, y 0 ).<br />

The last inclusion implies (2.12), and also it implies (2.13) (since W (C, y 0 )isconvex<br />

and compact).<br />

Proposition 2.3. The following relationships are valid:<br />

(2.14) lim S→∞ G S ≥ g ∗ ,<br />

(2.15) lim<br />

S→∞<br />

max ρ(γ,W)=0.<br />

γ∈ ¯coΓ S<br />

Proof. The proof of the proposition is contained in the corresponding part of the<br />

proof of Theorem 2.1(i) in [28] (see also Proposition 2 and Corollary 3 in [30]). For<br />

the sake of completeness, we also give a sketch of the proof below.<br />

Let S k ,k=1, 2,..., be such that S k →∞,andlet γ k ∈ Γ(S k ). Let also γ be a<br />

partial limit of {γ k }. That is, there exists a subsequence {k ′ }⊂{k} such that<br />

(2.16) lim<br />

k ′ →∞ γk′ = γ.<br />

By (1.10), it implies that<br />

∫<br />

∫<br />

(2.17) lim (φ ′ (y)) T f(u, y)γ k (du, dy) =<br />

k→∞ U×Y<br />

U×Y<br />

(φ ′ (y)) T f(u, y)γ(du, dy)<br />

for any φ ∈ C 1 . Also, from the fact that γ k ∈ Γ(S k ), it follows that there exist an<br />

initial condition y k ∈ Y and a control u k (·) ∈U Y S k (y k 0 ) such that<br />

∫<br />

U×Y<br />

(φ ′ (y)) T f(u, y)γ k (du, dy) = 1 ∫ S<br />

k<br />

S k (φ ′ (y k (τ))) T f(u k (τ),y k (τ))dτ,<br />

0<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2486 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

where y k (τ) def<br />

= y(t, y k 0 ,uk (·)) is the corresponding solution of system (1.1). The second<br />

integral in the expression above is equal <strong>to</strong><br />

φ(y k (S k )) − φ(y k (0))<br />

S k<br />

and tends <strong>to</strong> zero as S k tends <strong>to</strong> infinity (since y k (τ) ∈ Y ∀ τ ∈ [0,S k ]andY is a<br />

compact set). This and (2.16), (2.17) imply that<br />

∫<br />

(φ ′ (y)) T f(u, y)γ(du, dy) = 0 ∀φ ∈ C 1 ⇒ γ ∈ W,<br />

U×Y<br />

which, in turn, implies that<br />

(2.18) lim S→∞ Γ S ⊂ W.<br />

From (2.18) it follows that (2.14) is valid (see (2.11)) and also that lim S→∞ max γ∈ΓS<br />

ρ(γ,W) =0. The latter implies (2.13) due <strong>to</strong> the definition of ρ (see (1.9)) and due<br />

<strong>to</strong> the convexity of W .<br />

Finally, let us establish the following straightforward relationships between g ∗ (C, y 0 )<br />

and g ∗ and between W (C) def<br />

= ⋃ y W (C, y 0∈Y 0)andW .<br />

Lemma 2.4. The following relationships are valid:<br />

(2.19) lim C→0 inf<br />

y 0∈Y g∗ (C, y 0 ) ≥ g ∗ ;<br />

(2.20) lim C→0 W (C) ⊂ W.<br />

Also<br />

(2.21) lim<br />

max ρ(γ,W)=0.<br />

C→0 γ∈ ¯coW (C)<br />

Proof. Let γ l ∈ W (C l ,y 0 ),l=1, 2,..., and let C l → 0asl →∞.Letalsoγ be a<br />

def<br />

partial limit of γ l . That is, lim l′ →∞ γ l ′ = γ for some {l ′ }⊂{l}. Then, by passing <strong>to</strong><br />

the limit in (1.5), one can obtain that γ ∈ W . This proves (2.20). The inequality (2.19)<br />

follows from (2.20). Also from (2.20) it follows that lim C→0 max γ∈W (C) ρ(γ,W)=0.<br />

The latter implies (2.21).<br />

3. Dual problem and duality relationships. In this section we introduce a<br />

problem dual with respect <strong>to</strong> the IDLP problem (1.4) as a problem of maximization<br />

over functions ψ(·) ∈ C 1 (see (3.1) below), and we establish duality relationships<br />

between (1.4) and (3.1). We then present results allowing one <strong>to</strong> extend the class of<br />

functions used in the formulations of the problems (3.1) and (1.4) from C 1 <strong>to</strong> Lip (the<br />

class of locally Lipschitz continuous functions ψ(·) :R m → R), and we also establish<br />

necessary and sufficient conditions for γ ∗ ∈ W (C, y 0 ) <strong>to</strong> be an optimal solution of (1.4)<br />

(based on duality-type relationships). Results of this section are used <strong>to</strong> establish the<br />

fact that the inequality (2.12) and the inclusion (2.13) take the form of equalities and<br />

<strong>to</strong> construct necessary and sufficient optimality conditions for (1.3) (see section 4).<br />

Let us consider the problem<br />

{<br />

(3.1) sup μ<br />

∣ ∃ψ(·) ∈ }<br />

C1 : ∀(y, u) ∈ Y × U,<br />

μ ≤∇ψ(y) T = μ ⋆ (C, y<br />

f(y, u)+C(ψ(y 0 ) − ψ(y)) + g(y, u)<br />

0 ),<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2487<br />

which we will refer <strong>to</strong> as dual with respect <strong>to</strong> (1.4) (see section 8, where we discuss how<br />

problem (3.1) can be constructed as a “standard” LP dual). Note that, if W (C, y 0 ) ≠<br />

∅, then, for any γ ∈ W (C, y 0 ), from the fact that the pair (μ, ψ(·)) satisfies the<br />

inequality<br />

μ ≤∇ψ(y) T f(y, u)+C(ψ(y 0 ) − ψ(y)) + g(y, u) ∀(y, u) ∈ Y × U,<br />

it follows that μ ≤ ∫ g(y, u)γ(dy, du). Hence, the optimal values of (1.4) and its<br />

Y ×U<br />

dual (3.1) satisfy the inequality<br />

(3.2) μ ∗ (C, y 0 ) ≤ g ∗ (C, y 0 ).<br />

As follows from the theorem stated below, the inequality (3.2) turns in<strong>to</strong> the equality<br />

(that is, there is no duality gap) if and only if W (C, y 0 )isnotempty.<br />

Theorem 3.1. (i) The optimal value of problem (3.1) is bounded (that is,<br />

μ ∗ (C, y 0 ) < ∞) if and only if the set W (C, y 0 ) is not empty; (ii) if the optimal<br />

value of problem (3.1) is bounded, then<br />

(3.3) μ ∗ (C, y 0 )=g ∗ (C, y 0 );<br />

(iii) the optimal value of of problem (3.1) is unbounded (that is, μ ∗ (C, y 0 )=∞) if<br />

and only if there exists a function ψ(·) ∈ C 1 such that<br />

(3.4) max<br />

(y,u)∈Y ×U {∇ψ(y)T f(y, u)+C(ψ(y 0 ) − ψ(y))} < 0.<br />

Proof. The proof of the theorem is in the appendix.<br />

By taking C = 0 in (3.1), one obtains the problem<br />

{<br />

(3.5) sup μ<br />

∣ ∃ψ(·) ∈ }<br />

C1 : ∀(y, u) ∈ Y × U,<br />

μ ≤∇ψ(y) T f(y, u)+g(y, u)<br />

def<br />

= μ ⋆ .<br />

As has been shown in [22] (see Theorem 4.1 in [22]), this problem is dual with respect<br />

<strong>to</strong> (1.7) in the sense that the dual relationships between the latter and (3.5) (similar<br />

<strong>to</strong> those established by Theorem 3.1) are true. The proof of Theorem 3.1 follows<br />

exactly the same steps as those of the above mentioned result of [22], and the latter<br />

can be viewed as a special case (C = 0) of Theorem 3.1. In fact, with C =0,the<br />

relationships (3.3) and (3.4) take the forms<br />

(3.6) μ ∗ def<br />

= μ ∗ (0,y 0 )=g ∗ (0,y 0 ) def<br />

= g ∗ ,<br />

and, respectively,<br />

(3.7) max<br />

(y,u)∈Y ×U {∇ψ(y)T f(y, u)} < 0,<br />

with (3.6) being valid if and only if the set W (0,y 0 )=W is not empty and with (3.7)<br />

being satisfied for some ψ(·) ∈ C 1 if and only if this set is empty (the latter is also<br />

equivalent <strong>to</strong> μ ∗ = ∞).<br />

Note that duality results similar <strong>to</strong> Theorem 3.1(ii) have been obtained in [11]<br />

and [24] in a s<strong>to</strong>chastic setting without state constraints (Y = R m ) and in [58] in the<br />

deterministic setting with state constraints (for IDLP problems related <strong>to</strong> optimal<br />

control problems considered on a finite time interval). Note also that problem (3.1)<br />

is equivalent <strong>to</strong><br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2488 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

(3.8) sup<br />

ψ(·)∈C 1<br />

min<br />

(y,u)∈Y ×U {∇ψ(y)T f(y, u)+C(ψ(y 0 ) − ψ(y)) + g(y, u)} = μ ⋆ (C, y 0 )<br />

and that max-min representations similar <strong>to</strong> (3.8) (with C = 0) have been studied for<br />

optimally controlled diffusions in [43] (see also [31] and references therein).<br />

Let us now extend the class of functions used in (3.1) from C 1 <strong>to</strong> Lip.<br />

Lemma 3.2. The following representation for μ ∗ (C, y 0 ) is valid:<br />

{<br />

(3.9) μ ⋆ (C, y 0 )=sup μ<br />

∣<br />

}<br />

∃ψ(·) ∈ Lip : ∀(y, u) ∈ Y × U,<br />

μ ≤ min ξ∈∂ψ(y) ξ T ,<br />

f(y, u)+g(y, u)+C(ψ(y 0 ) − ψ(y))<br />

where ∂ψ(y) stands for Clarke’s generalized gradient of ψ(y).<br />

Proof. Let us denote by ¯μ(C, y 0 ) the right-hand side of (3.9). It is easy <strong>to</strong> see<br />

that μ ∗ (C, y 0 ) ≤ ¯μ(C, y 0 ) and, hence, the statement is obvious if μ ∗ (C, y 0 )=+∞.<br />

Suppose that μ ∗ (C, y 0 ) < +∞, and take arbitrary μ ∈ R, ψ(·) ∈ Lip such that<br />

(3.10) μ ≤ min<br />

ξ∈∂ψ(y) ξT f(y, u)+g(y, u)+C(ψ(y 0 ) − ψ(y)) ∀(y, u) ∈ Y × U.<br />

Due <strong>to</strong> the fact that y ↦→ ∂φ(y) is upper semicontinuous and has convex and compact<br />

values, and due <strong>to</strong> the continuity of the functions f(·),g(·), ψ(˙), corresponding <strong>to</strong> any<br />

ε>0, there exists ν(ε) > 0 such that lim ε→0 ν(ε) =0andsuchthat<br />

(3.11)<br />

μ − ν(ε) ≤ min<br />

ξ∈∂ψ(y) ξT f(y, u)+g(y, u)+C(ψ(y 0 ) − ψ(y)) ∀(y, u) ∈ (Y + εB) × U,<br />

where B is the open unit ball with the center at the origin in R m .<br />

Fix ε ∈ (0, 1). By Theorem 2.2 in [19] (cf. also [53]), there exists ψ ε (·) ∈ C 1 such<br />

that, for any y ∈ Y + εB,<br />

(3.12) |ψ(y) − ψ ε (y)| ≤ε;<br />

(3.13) ∇ψ ε (y) ∈<br />

⋃<br />

y ′ ∈y+εB<br />

∂ψ(y ′ )+εB.<br />

def<br />

Let ‖f‖ ∞ =max (y,u)∈ Ŷ ×U<br />

||f(y, u)||, where||·|| stands for the Euclidean norm of<br />

a vec<strong>to</strong>r in R m and Ŷ is a compact set containing Y + εB for all ε small enough. Let<br />

L(f),L(g) denote Lipschitz constants in y of f(y, u),g(y, u), and let L(ψ) denotethe<br />

Lipschitz constant of ψ(y) fory from the set Ŷ .<br />

Fix arbitrary (y, u) ∈ Y × U. By (3.13), there exist y ε ∈ y + εB, ξ ε ∈ ∂ψ(y ε ),<br />

and b ε ∈ εB such that<br />

(3.14) ∇ψ ε (y) =ξ ε + b ε .<br />

Using this, one can obtain that<br />

∇ψ ε (y) T f(y, u)+g(y, u)+C(ψ ε (y 0 ) − ψ ε (y))<br />

≥ ξ T ε f(y, u)+g(y, u)+C(ψ ε(y 0 ) − ψ ε (y)) − ε‖f‖ ∞<br />

≥ ξ T ε f(y ε ,u)+g(y ε ,u)+C(ψ ε (y 0 ) − ψ ε (y)) − ε(‖f‖ ∞ + L(f)||ξ ε || + L(g))<br />

≥ ξ T ε f(y ε,u)+g(y ε ,u)+C(ψ ε (y 0 ) − ψ ε (y)) − ε(‖f‖ ∞ + L(f)L(ψ)+L(g))<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2489<br />

(since ξ ε ∈ ∂ψ(y ε ) ⊂ L(ψ) ¯B, with ¯B being the closure of B)<br />

≥ ξ T ε f(y ε,u)+g(y ε ,u)+C(ψ(y 0 ) − ψ(y)) − ε(‖f‖ ∞ + L(f)L(ψ)+L(g)+2C)<br />

(due <strong>to</strong> (3.12))<br />

≥ ξ T ε f(y ε ,u)+g(y ε ,u)+C(ψ(y 0 ) − ψ(y ε )) − ε(‖f‖ ∞ + L(f)L(ψ)+L(g)+2C<br />

+ CεL(ψ))<br />

≥ min ξ∈∂ψ(yε) ξ T f(y ε ,u)+g(y ε ,u)+C(ψ(y 0 ) − ψ(y ε )) − εR,<br />

where R def<br />

= ‖f‖ ∞ + L(f)L(ψ)+L(g)+2C + CL(ψ). By (3.11), this implies<br />

μ − εR − ν(ε) ≤∇ψ ε (y) T f(y, u)+g(y, u)+C(ψ ε (y 0 ) − ψ ε (y)).<br />

Since the latter is valid for any element (y, u) ∈ Y × U, itleads<strong>to</strong><br />

μ − εR − ν(ε) ≤ μ ∗ (C, y 0 ).<br />

This, in turn, leads <strong>to</strong> the inequality μ ≤ μ ∗ (C, y 0 )asε can be arbitrary small, and,<br />

consequently, <strong>to</strong> the inequality ¯μ(C, y 0 ) ≤ μ ∗ (C, y 0 )asμ, ψ(·) were chosen arbitrarily<br />

just <strong>to</strong> satisfy (3.10).<br />

The following lemma and its corollary establish that the set W (C, y 0 ) can also be<br />

characterized with the help of Lipschitz continuous functions.<br />

Lemma 3.3. If γ ∈ W (C, y 0 ), then, for any ψ(·) ∈ Lip,<br />

∫ (<br />

)<br />

(3.15)<br />

min<br />

ξ∈∂ψ(y) ξT f(y, u)+C(ψ(y 0 ) − ψ(y)) γ(dy, du) ≤ 0.<br />

Y ×U<br />

Proof. Let ψ(·) ∈ Lip. As in the proof of Lemma 3.2, corresponding <strong>to</strong> any<br />

ε ∈ (0, 1), there exists a function ψ ε (·) ∈ C 1 such that (3.12) and (3.13) are valid, and<br />

for any (y, u) ∈ Y × U,<br />

∇ψ ε (y) T f(y, u)+C(ψ ε (y 0 ) − ψ ε (y)) ≥<br />

min<br />

ξ∈∂ψ(y ξT f(y ε ,u)+C(ψ(y 0 ) − ψ(y ε )) − ε ˆR,<br />

ε)<br />

where y ε ∈ y + εB and ˆR >0 is a large enough constant. Due <strong>to</strong> the upper semicontinuity<br />

of the map y ↦→ ∂φ(y), due <strong>to</strong> the continuity of the functions f(·), ψ(·),<br />

and also due <strong>to</strong> the fact that the set Y is compact, there exists ˆν(ε) > 0 such that<br />

lim ε→0 ˆν(ε) =0andsuchthat<br />

min<br />

ξ∈∂ψ(y ξT f(y ε ,u) ≥ min<br />

ε) ξ∈∂ψ(y) ξT f(y, u) − ˆν(ε)<br />

⇒ ∇ψ ε (y) T f(y, u)+C(ψ ε (y 0 ) − ψ ε (y)) ≥ min<br />

ξ∈∂ψ(y) ξT f(y, u)+C(ψ(y 0 ) − ψ(y))<br />

−ε ˆR − ˆν(ε).<br />

By integrating the last inequality over γ ∈ W (C, y 0 ), one obtains<br />

∫ (<br />

)<br />

0 ≥ min<br />

ξ∈∂ψ(y) ξT f(y, u)+C(ψ(y 0 ) − ψ(y)) γ(dy, du) − ε ˆR − ˆν(ε).<br />

Y ×U<br />

Consequently, by letting ε → 0, one establishes (3.15).<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2490 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

Corollary 3.4. The set W (C, y 0 ) allows the representations<br />

(3.16) {<br />

∫ (<br />

)<br />

W (C, y 0 )= γ ∈P(Y ×U) : min<br />

ξ∈∂φ(y) ξT f(y, u)+C(φ(y 0 ) − φ(y)) γ(dy, du)<br />

Y ×U<br />

≤ 0<br />

}<br />

∀φ ∈ Lip .<br />

Proof. By Lemma 3.3, W (C, y 0 ) is contained in the set defined by the right-hand<br />

side of (3.16). To prove the opposite inclusion, take an arbitrary γ belonging <strong>to</strong> the<br />

latter set. Then, for any φ(·) ∈ C 1 ,<br />

∫<br />

(∇φ(y) T f(y, u)+C(φ(y 0 ) − φ(y)))γ(dy, du) ≤ 0,<br />

∫<br />

Y ×U<br />

Y ×U<br />

(∇(−φ(y)) T f(y, u)+C((−φ(y 0 )) − (−φ(y))))γ(dy, du) ≤ 0<br />

⇒<br />

∫<br />

Y ×U<br />

(∇φ(y) T f(y, u)+C(φ(y 0 ) − φ(y)))γ(dy, du) =0.<br />

This implies that γ ∈ W (C, y 0 ).<br />

Finally, let us establish necessary and sufficient optimality conditions for γ ∗ ∈<br />

W (C, y 0 ) <strong>to</strong> be optimal in (1.4).<br />

Lemma 3.5. Let γ ∗ ∈ W (C, y 0 ),andletthereexistψ(·) such that<br />

(3.17) μ ⋆ (C, y 0 ) ≤ min<br />

ξ∈∂ψ(y) ξT f(y, u)+g(y, u)+C(ψ(y 0 ) − ψ(y)) ∀(y, u) ∈ Y × U.<br />

Then, for γ ∗ <strong>to</strong> be an optimal solution of the problem (1.4) it is necessary and sufficient<br />

that the following relationships are satisfied:<br />

∫ (<br />

)<br />

(3.18)<br />

min<br />

ξ∈∂ψ(y) ξT f(y, u)+C(ψ(y 0 ) − ψ(y)) γ ∗ (dy, du) =0;<br />

Y ×U<br />

(3.19) γ ∗ (Ω(C, y 0 )) = 1,<br />

where<br />

(3.20) Ω(C, y 0 ) def<br />

=<br />

{<br />

(y, u) ∈ Y × U :<br />

min<br />

ξ∈∂ψ(y) ξT f(y, u)+g(y, u)<br />

}<br />

+ C(ψ(y 0 ) − ψ(y)) = μ ⋆ (C, y 0 ) .<br />

Proof. Let γ ∗ ∈ W (C, y 0 ) be an optimal solution of (1.4). That is,<br />

∫<br />

g(y, u)γ ∗ (dy, du) =g ∗ (C, y 0 ).<br />

Y ×U<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2491<br />

By integrating (3.17) and taking in<strong>to</strong> account (3.3), one can obtain that<br />

∫ (<br />

)<br />

min<br />

ξ∈∂ψ(y) ξT f(y, u)+C(ψ(y 0 ) − ψ(y)) γ ∗ (dy, du) ≥ 0.<br />

Y ×U<br />

The latter <strong>to</strong>gether with Lemma 3.3 imply (3.18). Assume that (3.19) is not true.<br />

Then, by integrating (3.17) and taking in<strong>to</strong> account (3.18), one would obtain that<br />

μ ∗ (C, y 0 )


2492 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

4. Optimal control problem (1.3) and IDLP problem (1.4). In this section<br />

we use results of previous sections <strong>to</strong> show that, under mild conditions, inequality<br />

(2.12) and inclusion (2.13) relating optimal control problem (1.3) and its LP counterpart<br />

(1.4) take the form of equalities. We also use necessary and sufficient optimality<br />

conditions for (1.4) <strong>to</strong> construct necessary and sufficient optimality conditions for<br />

(1.3).<br />

Along with problem (1.3) let us consider the problem<br />

(4.1) inf<br />

u(·)∈U<br />

∫ +∞<br />

0<br />

e −Ct g(y(t, y 0 ,u(·)),u(t))dt def<br />

= V C (y 0 )<br />

and also a family of optimal control problems parameterized by δ (δ >0),<br />

(4.2) inf<br />

u(·)∈U Y δ (y 0)<br />

∫ +∞<br />

0<br />

e −Ct g(y(t, y 0 ,u(·))u(t))dt def<br />

= V Y δ<br />

C (y 0),<br />

where U Y δ<br />

(y 0 ) ⊂ U is the set of controls such that, for any u(·) ∈ U Y δ<br />

(y 0 ), the<br />

associated trajec<strong>to</strong>ry satisfies the inclusion<br />

y(t, y 0 ,u(·)) ∈ Y + δ<br />

¯B<br />

def<br />

= Y δ ∀t ≥ 0,<br />

with ¯B a closed unit ball of R m as above. Note that, for any δ>0,<br />

(4.3) V C (y 0 ) ≤ V Y δ<br />

C (y 0) ≤ V Y C (y 0 ) ∀y 0 ∈ Y.<br />

V Y δ<br />

C<br />

Lemma 4.1. Let U Y δ<br />

(y 0 ) be not empty for any y 0 ∈ Y δ , and let the function<br />

(·) (δ being fixed) satisfy Lipschitz conditions on Y δ.Then<br />

(4.4) CV Y δ<br />

C (y 0) ≤ μ ∗ (C, y 0 ) ∀y 0 ∈ Y.<br />

Proof. Let<br />

(4.5) H(y, ξ) def<br />

=max<br />

u∈U {−ξT f(y, u) − g(y, u)}.<br />

It is known that the function V Y δ<br />

C<br />

(·) is a viscosity solution of the Hamil<strong>to</strong>n–Jacobi–<br />

Bellman (HJB) equation<br />

(4.6) CV Y δ<br />

C<br />

+ H(y, DV Y δ<br />

C )=0,<br />

in intY δ (the interior of Y δ ). See Proposition III.2.8 on page 104 and comments on page<br />

277 in [7]. Being a viscosity solution (and due <strong>to</strong> the fact that it satisfies Lipschitz<br />

conditions on Y δ ), the function V Y δ<br />

C<br />

(·) also solves the HJB in the extended sense in<br />

intY δ (see Proposition II.5.13 on page 85 in [7] and also see results in [25]; further<br />

information on the <strong>to</strong>pic can be found in [16] and [18]). Namely,<br />

(4.7) CV Y δ<br />

C<br />

(y)+ max<br />

ξ∈∂V Y δ<br />

C (y) H(y, ξ) =0 ∀y ∈ intY δ .<br />

As Y ⊂ intY δ , from (4.5) and (4.7) it follows that<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2493<br />

{<br />

}<br />

(4.8) −CV Y δ<br />

C<br />

(y)+min min ξ T f(y, u)+g(y, u) =0 ∀y ∈ Y<br />

u∈U ξ∈∂V Y δ<br />

C (y)<br />

(4.9) ⇒ −CV Y δ<br />

C<br />

(y)+ min<br />

ξ∈∂V Y δ<br />

C (y) ξ T f(y, u)+g(y, u) ≥ 0 ∀(y, u) ∈ Y × U<br />

⇒ C(V Y δ<br />

C<br />

(y 0)−V Y δ<br />

C<br />

(y))+ min<br />

ξ∈∂V Y δ<br />

C<br />

(y) ξ T f(y, u)+g(y, u) ≥ CV Y δ<br />

C (y 0) ∀(y, u) ∈ Y ×U.<br />

By (3.9) in Lemma 3.2, the latter implies (4.4).<br />

Corollary 4.2. If V C (·) satisfies Lipschitz conditions in a neighborhood of Y ,<br />

then<br />

(4.10) CV C (y 0 ) ≤ μ ∗ (C, y 0 ) ∀y 0 ∈ Y ;<br />

(4.11) min<br />

y∈Y CV C(y) ≤ μ ∗ .<br />

Proof. Similarly <strong>to</strong> the proof above (see (4.9)) it is established that<br />

(4.12) −CV C (y)+ min<br />

ξ∈∂V C(y) ξT f(y, u)+g(y, u) ≥ 0<br />

∀(y, u) ∈ Y × U<br />

⇒ C(V C (y 0 ) − V C (y)) + min<br />

ξ∈∂V ξT f(y, u)+g(y, u) ≥ CV C (y 0 ) ∀(y, u) ∈ Y × U.<br />

C(y)<br />

This implies the validity of (4.10). From (4.12) it also follows that<br />

min<br />

ξ∈∂V C (y) ξT f(y, u)+g(y, u) ≥ min<br />

y ′ ∈Y CV C(y ′ ) ∀(y, u) ∈ Y × U,<br />

which, by (3.21) in Remark 3.6, implies (4.11).<br />

Let us now introduce an assumption that would ensure that<br />

(4.13) lim V Y δ<br />

C (y 0)=VC Y (y 0 ) ∀y 0 ∈ Y<br />

δ→0<br />

and, thus, by passing <strong>to</strong> the limit with δ → 0 in (4.4), would lead <strong>to</strong> the inequality<br />

(4.14) CV Y C (y 0) ≤ μ ∗ (C, y 0 ) ∀y 0 ∈ Y.<br />

Note that (4.13) is satisfied au<strong>to</strong>matically if Y is invariant with respect <strong>to</strong> the solutions<br />

of system (1.1), in which case, for any δ>0,<br />

(4.15) V C (y 0 )=V Y δ<br />

C (y 0)=VC Y (y 0 ) ∀y 0 ∈ Y.<br />

Let us denote by P(U) the space of probability measures defined on the Borel<br />

subsets of U, and let us consider the so-called relaxed control system (see [59])<br />

(4.16) ẏ(t) = ¯f(y(t),v(t)), t ≥ 0,<br />

where<br />

(4.17) ¯f(y, v)<br />

def<br />

=<br />

∫<br />

U<br />

f(y, u)v(du)<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2494 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

and the controls (“relaxed controls”) are Lebesgue measurable functions v(·) :[0+<br />

∞) ↦→ P(U). Given a relaxed control v(·), let us denote by t ↦→ y(t, y 0 ,v(·)) the<br />

solution of system (4.16) obtained with this control and with the initial condition<br />

y(0) = y 0 . Let V Y (y 0 ) stand for the set of relaxed controls such that, for any v(·) ∈<br />

V Y (y 0 ), the associated trajec<strong>to</strong>ry satisfies the inclusion y(t, y 0 ,v(·)) ∈ Y for all t ≥ 0.<br />

Note that U Y (y 0 ) can be considered a subset of V Y (y 0 )thatconsistsofonlyDirac<br />

measure-valued functions.<br />

Assumption I. For any q(·) ∈ C(Y × U),<br />

inf<br />

u(·)∈U Y (y 0)<br />

∫ +∞<br />

0<br />

e −Ct q(y(t, y 0 ,u(·)),u(t))dt<br />

= inf<br />

v(·)∈V Y (y 0)<br />

∫ +∞<br />

0<br />

e −Ct¯q(y(t, y 0 ,v(·)),v(t))dt ∀y 0 ∈ Y,<br />

where ¯q(y, v) def<br />

= ∫ q(y, u)v(du).<br />

U<br />

Sufficient conditions for Assumption I <strong>to</strong> be satisfied are those that ensure the<br />

applicability of Filippov–Wazewski type theorems on Y (see [26]). In particular,<br />

Assumption I is satisfied if Y is invariant with respect <strong>to</strong> the solutions of system<br />

(1.1) or if f(y, u) ≡ f(y) (the case of uncontrolled dynamics). Some other types<br />

of conditions, under which Assumption I is satisfied, have been described in [29].<br />

For example, it is satisfied if U and Y are convex, f(y, u) is linear, and there exist<br />

ȳ ∈ intY, ū ∈ U such that f(ȳ, ū) = 0 (see Proposition 4.3 in [29]).<br />

Assumption I is not satisfied if, for example, U Y (y 0 ) = ∅, while V Y δ<br />

(y 0 ) ≠<br />

∅ for all y 0 ∈ Y ,asisthecasewhen m =1, f(y, u) =−y + u, with U consisting<br />

of two points, U = {−1, 1}, and Y consisting of one point, Y = {0}. See<br />

Assumption 1 and Remark 1 in [30].<br />

Lemma 4.3. Let U Y (y 0 ) ≠ ∅ for all y o ∈ Y , and let Assumption I be satisfied.<br />

Then (4.13) is valid.<br />

Proof. The proof follows a standard argument. An outline of the proof is given<br />

in the appendix.<br />

Theorem 4.4. Let U Y (y 0 ) ≠ ∅ for all y 0 ∈ Y and let U Y δ<br />

(y 0 ) ≠ ∅ for all y 0 ∈ Y δ<br />

for any δ ∈ (0,δ 0 ] (δ 0 > 0 being small enough). Let also Assumption I be satisfied<br />

and the function V Y δ<br />

C<br />

(·) satisfy Lipschitz conditions on Y δ.Then<br />

(4.18) CV Y C (y 0 )=g ∗ (C, y 0 ) ∀y 0 ∈ Y.<br />

If, for any Lipschitz continuous q(·) :R m × U → R, the function V Y δ<br />

C,q (·),<br />

(4.19) V Y δ<br />

C,q (y 0) def<br />

= inf<br />

u(·)∈U Y δ (y 0)<br />

∫ +∞<br />

satisfies Lipschitz conditions on Y δ ,then<br />

(4.20) ¯coΓ(C, y 0 )=W (C, y 0 ).<br />

0<br />

e −Ct q(y(t, y 0 ,u(·)),u(t))dt,<br />

Proof. By Lemma 4.3, one can pass <strong>to</strong> the limit with δ → 0 in (4.4) <strong>to</strong> obtain<br />

(4.14). The latter and (2.12) imply (4.18). From the fact that V Y δ<br />

C,q<br />

(·) satisfies Lipschitz<br />

conditions on Y δ (with the other conditions of the theorem assumed <strong>to</strong> be satisfied) it<br />

follows that the equality similar <strong>to</strong> (4.18) is valid with the replacement of g(·) byany<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2495<br />

Lipschitz continuous q(·). In the notation introduced in section 2, this can be written<br />

as<br />

∫<br />

∫<br />

(4.21) inf q(y, u)γ(dy, du) = min q(y, u)γ(dy, du).<br />

γ∈Γ(C,y 0) Y ×U<br />

γ∈W (C,y 0) Y ×U<br />

Since<br />

∫<br />

∫<br />

inf q(y, u)γ(dy, du) = min q(y, u)γ(dy, du),<br />

γ∈Γ(C,y 0) Y ×U<br />

γ∈ ¯coΓ(C,y 0) Y ×U<br />

from (4.21) it follows that<br />

∫<br />

(4.22) min<br />

γ∈ ¯coΓ(C,y 0)<br />

Y ×U<br />

q(y, u)γ(dy, du) =<br />

∫<br />

min q(y, u)γ(dy, du).<br />

γ∈W (C,y 0) Y ×U<br />

Due <strong>to</strong> the fact that the space of Lipschitz continuous functions is dense in C(Y × U),<br />

the latter will be also valid for any q(·) ∈ C(Y × U). This implies (4.20).<br />

Remark 4.5. Note that the introduction of Assumption I would not be necessary<br />

(it would just be satisfied au<strong>to</strong>matically) if the problem (1.4) was formulated in the<br />

relaxed control setting (that is, if, instead of (1.1), we used system (4.16) with controls<br />

v(·) ∈V Y (y 0 )). Note also that the assumption about Lipschitz continuity of V Y δ<br />

C (·)<br />

used in Theorem 4.4 is a technical one (that is, it is related <strong>to</strong> the argument used in the<br />

proof of the theorem). In fact, relationships similar <strong>to</strong> (4.18) and (4.20) can be proved<br />

without this assumption (e.g., by modifying the approach of [58] <strong>to</strong> make it applicable<br />

<strong>to</strong> problems considered on the infinite time horizon). Such a proof, however, is much<br />

more involved, and, hence, is not included in this paper (<strong>to</strong> keep the presentation<br />

exposi<strong>to</strong>ry).<br />

To conclude this section, let us construct necessary and sufficient optimality conditions<br />

for the optimal control problem (1.3) based on the assumption that a function<br />

ψ(·) satisfying (3.17) exists. For every y ∈ Y ,let<br />

{<br />

}<br />

K(y) def<br />

=min<br />

u∈U<br />

min<br />

ξ∈∂ψ(y) ξT f(y, u)+g(y, u)<br />

{<br />

}<br />

(4.23) D(y) def<br />

= u ∈ U : K(y) = min<br />

ξ∈∂ψ(y) ξT f(y, u)+g(y, u) ,<br />

and let<br />

(4.24) Y def<br />

= {y ∈ Y : K(y)+C(ψ(y 0 ) − ψ(y)) = μ ⋆ (C, y 0 )}.<br />

Note that, in this notation, the set Ω(C, y 0 ) introduced in (3.20) is presented in the<br />

form<br />

(4.25) Ω(C, y 0 )={(y, u) ∈ Y × U : u ∈D(y), y ∈Y}.<br />

Proposition 4.6. Let a function ψ(·) satisfying (3.17) exist. Then for a control<br />

u ∗ (·) ∈U Y (y 0 ) <strong>to</strong> be optimal in (1.4) and for the equality (4.18) <strong>to</strong> be true, it is<br />

necessary and sufficient that<br />

∫ ∞<br />

(<br />

)<br />

(4.26) e −Ct min<br />

ξ∈∂ψ(y ∗ (t)) ξT f(y ∗ (t),u ∗ (t)) + C(ψ(y(t)) − ψ(y 0 )) dt =0<br />

0<br />

,<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2496 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

and<br />

(4.27) u ∗ (t) ∈D(y ∗ (t)), y ∗ (t) ∈Y,<br />

for almost all t ∈ [0, ∞), wherey ∗ (t) def<br />

= y(t, y0 ∗,u∗ (·)).<br />

Proof. Let u ∗ (·) ∈U Y (y 0 ) be optimal in (1.4) and let (4.18) be satisfied. Then<br />

the discounted occupational measure γu C def<br />

∗ (·)<br />

= γ ∗ generated by the pair (y ∗ (·),u ∗ (·))<br />

on the interval [0, ∞) is a solution of IDLP problem (1.4). Hence, by Lemma 3.5, the<br />

relationships (3.18) and (3.19) are satisfied, (3.18) is equivalent <strong>to</strong> (4.26), and (3.19)<br />

is equivalent <strong>to</strong> the fact that the inclusion (y ∗ (t),u ∗ (t)) ∈ Ω(C, y 0 ) is valid for almost<br />

all t ∈ [0, ∞). The validity of the latter is equivalent <strong>to</strong> (4.27) (due <strong>to</strong> (4.25)).<br />

Conversely, let (4.26) and (4.27) be satisfied. Then the discounted occupational<br />

measure γ ∗ generated by the pair (y ∗ (·),u ∗ (·)) satisfies (3.18) and (3.19).<br />

By ∫ Lemma 3.5, it implies that γ ∗ is a solution of IDLP problem (1.4). Hence,<br />

∞<br />

e −Ct g(y(t),u(t))dt = g ∗ (C, y<br />

0 0 ). Thisprovesthatu ∗ (·) ∈ U Y (y 0 )isoptimal<br />

and that (4.18) is satisfied.<br />

Note that in a special case when Y is invariant, equalities (4.15) are valid, and<br />

from (4.8) it follows that<br />

(4.28) −CV C (y)+min<br />

u∈U<br />

{<br />

min<br />

ξ∈∂V C (y) ξT f(y, u)+g(y, u)<br />

}<br />

=0 ∀y ∈ Y,<br />

which implies that (3.17) is satisfied with ψ(y) =V (C, y).<br />

5. Characterization of the viability kernel of Y . In this section we demonstrate<br />

the possibility of applying one of the results obtained above (namely, Corollary<br />

4.2 of section 4) for a characterization of the viability kernel of Y .<br />

Let us note, first, that the viability kernel of Y (denoted as Viab f(·,U) ; see [5],<br />

[48], [49]) is defined as a “largest” subset of Y such that, for any point y 0 belonging<br />

<strong>to</strong> this subset, U Y (y 0 ) ≠ ∅. Thatis,<br />

y 0 ∈ Viab f(·,U) ⇔ U Y (y 0 ) ≠ ∅.<br />

We use here an idea of [49] for the characterization of the viability of Y . Todothis,<br />

consider problem (4.1) with<br />

(5.1) g(y, u) =d 2 Y (y), d Y (y) def<br />

= min<br />

y ′ ∈Y ||y − y′ ||.<br />

Proposition 5.1. Let the set f(y, U) def<br />

= {η : η = f(y, u), u ∈ U} be convex for<br />

any y ∈ Y .Theny 0 ∈ Viab f(·,U) if and only if W (C, y 0 ) ≠ ∅.<br />

Proof. If y 0 ∈ Viab f(·,U) (Y ), then Γ(C, y 0 ) ≠ ∅ and, consequently, W (C, y 0 ) ≠ ∅<br />

(see Proposition 2.2). Conversely, if W (C, y 0 ) ≠ ∅, then ∫ Y ×U d2 Y (y)γ(dy, du) =0for<br />

any γ ∈ W (C, y 0 ), and<br />

g ∗ (C, y 0 )=μ ∗ (C, y 0 )=0<br />

if g(·) in problems (1.4) and (3.1) is defined by (5.1). By choosing C large enough,<br />

one may assume (without loss of generality) that the function V C (·) satisfies Lipschitz<br />

conditions on R m . See, e.g., Proposition III.2.1 on page 99 in [7]. Hence, by Corollary<br />

4.2 (see 4.10), CV C (y 0 ) ≤ 0, which implies that V C (y 0 ) = 0 (since the value function<br />

is not negative with g(·) as in (5.1)). Consequently,<br />

(5.2) inf<br />

u(·)∈U<br />

∫ +∞<br />

0<br />

e −Ct d 2 Y (y(t, y 0,u(·)))dt =0.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2497<br />

Due <strong>to</strong> the fact that y ↦→ f(y, U) is compact and convex valued, the optimal control<br />

in the above problem exists. The corresponding solution of (1.1) is contained in Y .<br />

Hence, U Y (y 0 ) ≠ ∅.<br />

Proposition 5.2. Let f(y, U) be convex for any y ∈ Y . Then Viab f(·,U) ≠ ∅ if<br />

and only if W ≠ ∅.<br />

Proof. If Viab f(·,U) (Y ) ≠ ∅, thenΓ S ≠ ∅ for all S>0, and, consequently, W ≠ ∅<br />

(see Proposition 2.3). Conversely, let W ≠ ∅. Then ∫ Y ×U d2 Y (y)γ(dy, du) = 0 for any<br />

γ ∈ W ,and<br />

g ∗ = μ ∗ =0<br />

if g(·) in problems (1.7) and (3.5) is defined by (5.1). For C large enough, the optimal<br />

value function V C (·) of problem (4.1) satisfies Lipschitz conditions on R m , and, hence,<br />

one can use (4.11) <strong>to</strong> obtain<br />

C min<br />

y∈Y V C(y) ≤ 0 ⇒ min<br />

y∈Y V C(y) =0<br />

(the last equality being implied by the fact that V C (·) is nonnegative). It follows that<br />

there exists y 0 ∈ Y such that V C (y 0 ) = 0. That is, (5.2) is valid. Repeating now<br />

the argument used in the proof of Proposition 5.1, one arrives at the conclusion that<br />

Viab f(·,U) (Y ) ≠ ∅.<br />

Remark 5.3. Note that, in the relaxed control setting, Propositions 5.1 and 5.2<br />

are valid without the assumption about the convexity of f(y, U) (see[49]andRemark<br />

4.5 above).<br />

Remark 5.4. From Proposition 5.1 and Theorem 3.1 it follows that y 0 ∈ Viab f(·,U)<br />

if and only if there does not exist a function ψ(·) satisfying (3.4). Also from Proposition<br />

5.2 and Theorem 3.1 (considered with C = 0), it follows that Viab f(·,U) ≠ ∅ if<br />

and only if there does not exist a function ψ(·) satisfying (3.7). The latter function<br />

can be interpreted as a Lyapunov function, its existence forcing the solutions of (1.1)<br />

<strong>to</strong> leave Y in a finite time (see Theorem 4.1 and Remark 4.2 in [22]).<br />

6. Optimal control problems with long run average criteria. In this section<br />

we show that one can use Theorem 4.4 of section 4 and Abelian-type results<br />

obtained in [32], [33] <strong>to</strong> establish relationships between the optimal control problem<br />

with long run average criteria (that is, problem (1.6) with S →∞)andtheIDLP<br />

problem (1.7). Note that similar relationships have been established in Theorem<br />

2.1(i) of [28] and in Proposition 5 of [30] under weaker assumptions than those used<br />

in this section. However, the considerations in [28] and [30] were based on results from<br />

s<strong>to</strong>chastic control theory (see [11], [39], and [54]), the proof of which being rather sophisticated.<br />

The proof presented below is straightforward and is based on a purely<br />

deterministic argument.<br />

Proposition 6.1. Let (i) U Y (y 0 ) ≠ ∅ for all y 0 ∈ Y and U Y δ<br />

(y 0 ) ≠ ∅ for all y 0 ∈<br />

Y δ for any δ ∈ (0,δ 0 ] (δ 0 > 0 being small enough); (ii) Assumption I be valid for every<br />

C ∈ (0,C 0 ] (C 0 being a given positive number); and (iii) the function V Y δ<br />

C<br />

(·) satisfy<br />

Lipschitz conditions on Y δ (with a Lipschitz constant being independent on δ ∈ (0,δ 0 ])<br />

for every C ∈ (0,C 0 ]. Then there exists the limit<br />

(6.1) lim<br />

C→0<br />

min<br />

y∈Y CV Y C (y) =g∗ .<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2498 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

If also (iv) for every C ∈ (0,C 0 ] and any Lipschitz continuous q(·) :R m × U → R,<br />

the function V Y δ<br />

C,q (·) defined by (4.19) satisfies Lipschitz conditions on Y δ,then<br />

(6.2) lim ρ H (¯ coΓ(C),W)=0,<br />

C→0<br />

Proof. By (4.9),<br />

min<br />

ξ∈∂V Y δ<br />

C<br />

Hence, by (3.21) and (3.3),<br />

Γ(C) def<br />

= ⋃<br />

y 0∈Y<br />

Γ(C, y 0 ).<br />

(y)<br />

ξ T f(y, u)+g(y, u) ≥ min<br />

y ′ ∈Y CV Y δ<br />

C (y′ ) ∀(y, u) ∈ Y × U.<br />

(6.3) g ∗ ≥ min<br />

y∈Y CV Y δ<br />

C (y) ⇒ g∗ ≥ min<br />

y∈Y CV Y C (y),<br />

where the latter is obtained via passing <strong>to</strong> the limit in the former (and taking in<strong>to</strong><br />

account (4.13) as well as the fact that V Y δ<br />

C<br />

(·) satisfies Lipschitz conditions with a<br />

constant independent of δ). Since, by (2.19) and (2.12),<br />

(6.4) lim C→0 min<br />

y∈Y CV Y C (y) ≥ g∗ ,<br />

the validity of (6.1) is established.<br />

Under the additional assumption that the function V Y δ<br />

C,q<br />

(·) defined by (4.19) satisfies<br />

Lipschitz conditions on Y δ for every C ∈ (0,C 0 ] and for any Lipschitz continuous<br />

q(·) :R m × U → R, the equality similar <strong>to</strong> (6.1) is valid with the replacement of g(·)<br />

by any Lipschitz continuous q(·). Due <strong>to</strong> the definition of Γ(C) (see (6.2)), this can<br />

be written as<br />

∫<br />

∫<br />

lim inf q(y, u)γ(dy, du) =min q(y, u)γ(dy, du)<br />

C→0 γ∈Γ(C) Y ×U<br />

γ∈W Y ×U<br />

(6.5) ⇒ lim<br />

min<br />

C→0 γ∈ ¯coΓ(C)<br />

∫<br />

Y ×U<br />

∫<br />

q(y, u)γ(dy, du) =min q(y, u)γ(dy, du).<br />

γ∈W Y ×U<br />

Since the space of Lipschitz continuous functions is dense in C(Y × U), the latter is<br />

also valid for any q(·) ∈ C(Y × U). Let us now use this fact <strong>to</strong> prove (6.2).<br />

First, note that, by (2.21) and (2.13),<br />

lim max ρ(γ,W)=0.<br />

C→0 γ∈ ¯coΓ(C)<br />

Thus, <strong>to</strong> prove (6.2), one needs <strong>to</strong> show that<br />

(6.6) lim<br />

max<br />

C→0 γ∈W<br />

ρ(γ, ¯coΓ(C)) = 0.<br />

Assume it is not true. Then there exists a positive number α and sequences C i > 0,<br />

γ i ∈ W , i =1, 2,..., such that lim i→∞ C i =0 and<br />

(6.7) lim ρ(γ i , coΓ(C ¯ i )) ≥ α.<br />

i→∞<br />

Due <strong>to</strong> the fact that W is compact in metric ρ, and due <strong>to</strong> the fact that (by Blaschke’s<br />

selection theorem; see, e.g., [38]) the set of closed subsets of P(Y × U) iscompactin<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2499<br />

metric ρ H , the sequences {γ i } and { ¯coΓ(C i )} have partial limits. More specifically,<br />

there exist ˆγ ∈ W and ˆΓ ⊂P(Y × U) such that, for some subsequence {i ′ }⊂{i},<br />

(6.8) lim ρ(γ i ′, ˆγ) =0, lim<br />

i ′ →∞<br />

ρ H(¯ coΓ(C i ′), ˆΓ) = 0.<br />

i ′ →∞<br />

Then, by passing <strong>to</strong> the limit in (6.7), one obtains that<br />

(6.9) ρ(ˆγ, ˆΓ) ≥ α ⇒ ˆγ /∈ ˆΓ.<br />

By the separation theorem (see, e.g., [52, p. 59]), there exists ˆq(·) ∈ C(Y × U) such<br />

that<br />

∫<br />

∫<br />

ˆq(y, u)ˆγ(dy, du) ≤ min ˆq(y, u)γ(dy, du) − β,<br />

Y ×U<br />

γ∈ˆΓ Y ×U<br />

where β>0 is some constant. Hence,<br />

∫<br />

(6.10) min<br />

γ∈W Y ×U<br />

From (6.8) it follows that<br />

lim<br />

min<br />

i ′ →∞ γ∈ ¯coΓ(C i ′ )<br />

∫<br />

ˆq(y, u)γ(dy, du) ≤ min ˆq(y, u)γ(dy, du) − β.<br />

γ∈ˆΓ Y ×U<br />

∫<br />

Y ×U<br />

∫<br />

ˆq(y, u)γ(dy, du) =min ˆq(y, u)γ(dy, du).<br />

γ∈ˆΓ Y ×U<br />

Consequently, by (6.10),<br />

∫<br />

∫<br />

(6.11) min ˆq(y, u)γ(dy, du) ≤ min ˆq(y, u)γ(dy, du) − β<br />

γ∈W Y ×U<br />

γ∈ ¯coΓ(C i ′ ) Y ×U<br />

2<br />

for i ′ large enough. This contradicts the fact that (6.5) is valid with any q(·) ∈ C(Y ×<br />

U). The contradiction proves (6.6) and, thus, it completes the proof of (6.2).<br />

Proposition 6.2. Let conditions (i), (ii), and(iii) of Proposition 6.1 be satisfied.<br />

Then<br />

(6.12) lim<br />

S→∞ G S = g ∗ .<br />

If also condition (iv) of Proposition 6.1 is satisfied, then<br />

(6.13) lim ρ H(¯ coΓ S ,W)=0.<br />

S→∞<br />

Proof. Let C i → 0. By (6.1), there exists a sequence of controls u i (·) ∈US Y (yi 0)<br />

and the corresponding sequence of solutions y i (t) def<br />

= y(t, y0 i ,ui (·)) of system (1.1) such<br />

that<br />

(6.14) lim<br />

i→∞<br />

ζ i =0,<br />

def<br />

ζ i =<br />

∫ +∞<br />

0<br />

e −Cit g(y i (t),u i (t))dt − g ∗ .<br />

From Lemma 3.5(ii) in [33] it follows that there exists a sequence S i , i =1, 2,...,such<br />

that S i ≥ (K>0 being a constant) and such that<br />

K √ Ci<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2500 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

(6.15)<br />

∫<br />

1<br />

Si<br />

g(y i (t),u i (t))dt ≤ g ∗ + ζ i + √ C i<br />

S i 0<br />

⇒ G Si ≤ g ∗ + ζ i + √ C i ⇒ lim S→∞ G S ≤ g ∗ .<br />

The latter inequality and (2.14) imply that<br />

(6.16) lim S→∞ G S = g ∗ ,<br />

which, by (6.15), implies that<br />

(6.17) lim<br />

i→∞<br />

η i =0,<br />

def<br />

η i = 1 ∫ Si<br />

S i<br />

0<br />

g(y i (t),u i (t))dt − g ∗ .<br />

From Lemma 3.8 in [32] it follows that, for any i, there exists a nonnegative t i ≤<br />

S i − √ S i<br />

L<br />

(L>0 being a constant) such that<br />

∫<br />

1 S<br />

(6.18)<br />

g(y i (t i + t),u i (t i + t))dt ≤ g ∗ + η i + 1 ( √ ]<br />

Si<br />

√ ∀S ∈ 0, .<br />

S 0<br />

Si<br />

L<br />

Let ũ i (·) def<br />

= u i (t i + ·), ỹ i (·) def<br />

= y i (t i + ·). Note that ũ i (·) ∈U Y (y i (t i )) and ỹ i (t) =<br />

y(t, y i (t i ), ũ i (·)). Hence, by (6.18),<br />

G S ≤ 1 S<br />

∫ S<br />

0<br />

g(ỹ i (t), ỹ i (t))dt ≤ g ∗ + η i + 1 √<br />

Si<br />

∀S ∈<br />

(<br />

0,<br />

√<br />

Si<br />

L<br />

]<br />

⇒ lim S→∞ G S ≤ g ∗ .<br />

The latter and (6.16) prove (6.12).<br />

If condition (iv) of Proposition 6.1 is satisfied, then the equality similar <strong>to</strong> (6.12)<br />

is valid with the replacement of g(·) by any Lipschitz continuous q(·). In the notation<br />

of section 2, this can be written as<br />

∫<br />

lim inf q(y, u)γ(dy, du) =min q(y, u)γ(dy, du)<br />

S→∞<br />

γ∈W<br />

(6.19) ⇒ lim<br />

γ∈Γ S<br />

∫Y ×U<br />

min<br />

S→∞ γ∈ ¯coΓ S<br />

∫Y ×U<br />

Y ×U<br />

∫<br />

q(y, u)γ(dy, du) =min q(y, u)γ(dy, du).<br />

γ∈W Y ×U<br />

Since the space of Lipschitz continuous functions is dense in C(Y × U), the latter is<br />

valid for any q(·) ∈ C(Y × U). From this point, the proof of (6.13) follows exactly the<br />

same lines as that of (6.2) in Proposition 6.1.<br />

Let us consider two special cases, in which the conditions of the above results are<br />

readily verifiable.<br />

Special case 1. Let there exist positive definite matrices A 1 and A 2 such that<br />

(6.20) (f(y ′ ,u) − f(y ′′ ,u)) T A 1 (y ′ − y ′′ )<br />

≤−(y ′ − y ′′ ) T A 2 (y ′ − y ′′ ) ∀y ′ ,y ′′ ∈ R m , ∀u ∈ U.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2501<br />

Then, for any u(·) ∈U, the solutions of system (1.1) satisfy the inequality<br />

(6.21) d<br />

dt (y(t, y′ 0 ,u(·)) − y(t, y′′ 0 ,u(·)))T A 1 (y(t, y 0 ′ ,u(·)) − y(t, y′′ 0 ,u(·)))<br />

≤−2(y(t, y 0 ′ ,u(·)) − y(t, y′′ 0 ,u(·)))T A 2 (y(t, y 0 ′ ,u(·)) − y(t, y′′ 0 ,u(·)))<br />

≤−α 1 (y(t, y 0,u(·)) ′ − y(t, y 0 ′′ ,u(·))) T A 1 (y(t, y 0,u(·)) ′ − y(t, y 0 ′′ ,u(·)))<br />

⇒ ||y(t, y 0 ′ ,u(·)) − y(t, y′′ 0 ,u(·))|| ≤ α 2||y 0 ′ − y′′ 0 ||e−α3t ,<br />

where α i > 0,i =1, 2, 3, are appropriate constants. Hence, for any q(y, u) satisfying<br />

Lipschitz conditions in y with a constant L q ,<br />

∫ ∞<br />

∫ ∞<br />

∥ ∥∥∥<br />

∥ e −Ct q(y(t, y 0 ′ ,u(·)),u(t))dt − e −Ct q(y(t, y 0 ′′ ,u(·)),u(t))dt<br />

0<br />

0<br />

≤<br />

∫ ∞<br />

0<br />

e −Ct L q ||y(t, y ′ 0,u(·)) − y(t, y ′′<br />

0 ,u(·))||dt ≤ ˆL q ||y ′ 0 − y ′′<br />

0 ||,<br />

where ˆL<br />

def<br />

q = α2Lq<br />

α 3<br />

> α2Lq<br />

(6.22) V C,q (y 0 ) def<br />

= inf<br />

u(·)∈U<br />

C+α 3<br />

. It follows that the optimal value function V C,q (·),<br />

∫ +∞<br />

0<br />

e −Ct q(y(t, y 0 ,u(·)),u(t))dt,<br />

satisfies Lipschitz conditions on R m with a constant ˆL q . Also, from the validity of<br />

the estimate (6.22) it follows that system (1.1) has an invariant set, which is a global<br />

attrac<strong>to</strong>r <strong>to</strong> all its solutions (see Theorem 3.1(ii) in [27]). Taking this set as Y ,one<br />

can arrive at the conclusion that the statements of Theorem 4.4 and Propositions<br />

6.1 and 6.2 are valid. Note that (6.20) is a Liapunov-type stability condition. It is<br />

satisfied, for example, if f(y, u) =Ay + f 1 (u), where A is a “stable” matrix (that is,<br />

its eigenvalues have negative real parts) or if, for all y and u, the eigenvalues of the<br />

) T are less than some negative constant.<br />

Special case 2. Assume that the set Y is convex and that there exist δ 0 > 0and<br />

r>0 such that<br />

matrix ∂f(y,u)<br />

∂y<br />

+( ∂f(y,u)<br />

∂y<br />

(6.23) y + rB ∈ f(y, U) ∀y ∈ Y δ0 .<br />

This is a controllability-type condition implying that any two points y 0 ′ and y 0 ′′ in Y δ<br />

and in Y are connectable by a trajec<strong>to</strong>ry of (1.1) lying in y δ (respectively, in Y ), with<br />

the time required for the transition from one point <strong>to</strong> another being equal <strong>to</strong> ‖y′ 0 −y′′ 0 ‖<br />

r<br />

.<br />

It can be verified that, under this condition, the value function V Y δ<br />

C,q<br />

(·) satisfies Lipschitz<br />

conditions on Y δ with a constant ˆL q = Lq<br />

r ,whereL def<br />

q =max (y,u)∈Yδ0 ×U |q(y, u)|<br />

(see, e.g., page 398 in [7]). If, in addition, Assumption I is satisfied, then again the<br />

statements of Theorem 4.4 and Propositions 6.1 and 6.2 are valid. Note that in the<br />

relaxed control setting, the condition (6.23) takes the form y + rB ∈ ¯cof(y, U) for all<br />

y ∈ Y δ0 .<br />

7. Finite dimensional approximations. Numerical example. In this section<br />

we show that problems (1.4) and (3.1) can be written in a “standard” LP form,<br />

and we discuss the possibility of approximating these problems with finite dimensional<br />

LP problems. We also illustrate that the latter can be used for finding a near optimal<br />

control in (1.3) with a numerical example.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2502 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

Let {φ i (·) ∈ C 1 ,i=1, 2,...} be a sequence of functions having continuous partial<br />

derivatives of the second order such that any function ψ(·) ∈ C 1 and its gradient<br />

∇ψ(·) can be simultaneously approximated on Y by linear combinations of functions<br />

from {φ i (·),i=1, 2,...} and their corresponding gradients. That is, for any ψ(·) ∈ C 1<br />

and any δ>0, there exist β 1 ,...,β k (real numbers) such that<br />

{∣ ∥ }<br />

∣∣∣∣ k∑<br />

∥∥∥∥ k∑<br />

(7.1) max ψ(y) − β i φ i (y)<br />

y∈Y<br />

∣ + ∇ψ(y) − β i ∇φ i (y)<br />

≤ δ,<br />

∥<br />

1<br />

with || · || being a norm in R m . An example of such an approximating sequence is the<br />

sequence of monomials y i1<br />

1 ...yim m , i 1,...,i m =0, 1,...,wherey j (j =1,...,m) stands<br />

for the jth component of y (see, e.g., [45]).<br />

Due <strong>to</strong> the above property of the sequence of the functions φ i (·),i=1, 2,..., the<br />

set W (C, y 0 ) can be presented in the form<br />

(7.2) {<br />

∫<br />

W (C, y 0 ) def<br />

= γ ∈P(Y × U) : (∇φ i (y) T f(y, u)+C(φ i (y 0 ) − φ i (y)))γ(dy, du)<br />

Y ×U<br />

1<br />

}<br />

=0 i =1, 2,... ,<br />

where, without loss of generality, one may assume that the functions φ i (·) satisfythe<br />

following normalization conditions:<br />

(7.3) max y∈ ˆD{|φ i (y)|, ||∇φ i (y)||} ≤ 1 , i =1, 2,...,<br />

2i where ||∇φ i (y)|| is a norm of ∇φ i (y) inR m ,and ˆD is a closed ball in R m that contains<br />

Y in its interior.<br />

Let l 1 and l ∞ stand for the Banach spaces of infinite sequences such that, for<br />

def<br />

any x =(x 1 ,x 2 ,...) ∈ l 1 , ||x|| l1 = ∑ i |x i| < ∞ and, for any λ =(λ 1 ,λ 2 ,...) ∈ l ∞ ,<br />

def<br />

||λ|| l∞ =sup i |λ i | < ∞. It is easy <strong>to</strong> see that, given an element λ ∈ l ∞ , one can define<br />

a linear continuous functional λ(·) :l 1 → R 1 by the equation<br />

(7.4) λ(x) = ∑ i<br />

λ i x i ∀x ∈ l 1 , ||λ(·)|| = ||λ|| l∞ .<br />

It is also known (see, e.g., [52, p. 86]) that any continuous linear functional λ(·) :<br />

l 1 → R 1 can be presented in the form (7.4) with some λ ∈ l ∞ . Note that from (7.3)<br />

it follows that (φ 1 (y),φ 2 (y),...) ∈ l 1 and ( ∂φ1 ∂φ<br />

∂y j<br />

, 2<br />

∂y j<br />

,...) ∈ l 1 for any y ∈ Y , and,<br />

hence, for any λ =(λ 1 ,λ 2 ,...) ∈ l ∞ , the function ψ λ (y),<br />

(7.5) ψ λ (y) def<br />

= ∑ i<br />

λ i φ i (y),<br />

is continuously differentiable, with ∇ψ λ (y) = ∑ i λ i∇φ i (y).<br />

Let us now rewrite problem (1.4) in a “standard” LP form by using the representation<br />

(7.2). Let M(Y × U) (respectively, M + (Y × U)) stand for the space of<br />

all (respectively, all nonnegative) measures with bounded variations defined on Borel<br />

subsets of Y × U, andletA(·) :M(U × Y ) ↦→ R 1 × l 1 stand for the linear opera<strong>to</strong>r<br />

defined for any γ ∈M(U × Y ) by the equation<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2503<br />

(∫<br />

∫<br />

A(γ) def<br />

= 1 Y ×U (y, u)γ(dy, du), (∇φ i (y) T f(y, u)<br />

Y ×U<br />

Y ×U<br />

)<br />

+ C(φ i (y 0 )−φ i (y)))γ(dy, du), i =1, 2,... .<br />

In this notation problem (1.4) takes the form<br />

(7.6) min<br />

γ<br />

{〈g, γ〉 | A(γ) =(1, 0), γ ∈M + },<br />

where 0 is the zero element of l 1 ,and〈·,γ〉, here and in what follows, stands for the<br />

integral of the corresponding function over γ.<br />

Define now the linear opera<strong>to</strong>r A ∗ (·) :R 1 × l ∞ ↦→ C(Y × U) ⊂M ∗ (Y × U) by<br />

the equation<br />

(7.7)<br />

A ∗ (μ, λ) def<br />

= μ + ∇ψ λ (·) T f(·, ·)+C(ψ λ (y 0 ) − ψ λ (·)) ∀ μ ∈ R 1 , ∀ λ =(λ i ) ∈ l ∞ ,<br />

where ψ λ (·) is as defined in (7.5). Note that from (7.7) it follows that, for any<br />

γ ∈M(Y × U),<br />

∫<br />

〈A ∗ (μ, λ),γ〉 = (μ1 Y ×U (y, u)+∇ψ λ (y) T f(y, u)+C(ψ λ (y 0 )−ψ λ (y)))γ(dy, du)<br />

Y ×U<br />

(7.8)<br />

def<br />

= 〈(μ, λ),A(γ)〉.<br />

That is, the opera<strong>to</strong>r A ∗ (·) istheadjoin<strong>to</strong>fA(·), and, hence, the problem dual <strong>to</strong><br />

(7.6) can be written in the form (see page 39 in [2])<br />

(7.9) sup<br />

(μ,λ)∈R 1 ×l ∞<br />

{μ |−A ∗ (μ, λ)+g(·) ≥ 0}<br />

and, by (7.7), is equivalent <strong>to</strong><br />

sup<br />

(μ,λ)∈R 1 ×l ∞<br />

{μ |−μ −∇ψ λ (y) T f(y, u) − C(ψ λ (y 0 ) − ψ λ (y))<br />

(7.10) + g(u, y) ≥ 0 ∀(y, u) ∈ Y × U}.<br />

Due <strong>to</strong> the approximation property (7.1), the optimal value in (7.10) will be the same<br />

as in the problem<br />

sup {μ |−μ−∇ψ(y) T f(y, u)−C(ψ(y 0 )−ψ(y))+g(u, y) ≥ 0 ∀(y, u) ∈ Y ×U}<br />

(μ,ψ(·))∈R 1 ×C 1<br />

= sup {μ |μ ≤∇ψ(y) T f(y, u)+C(ψ(y 0 ) − ψ(y)) + g(u, y) ∀(y, u) ∈ Y × U}<br />

(μ,ψ(·))∈R 1 ×C 1 = μ ⋆ (C, y 0 ),<br />

the latter being equivalent <strong>to</strong> (3.1).<br />

Let us now discuss the possibility of approximating the IDLP problem (1.4) and<br />

its dual (3.1) with certain finite dimensional LP problems obtained by truncating the<br />

infinite system of constraints in (7.2) <strong>to</strong> a system with a finite number of constraints<br />

and by considering probability measures concentrated on a finite number of points.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2504 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

Let the points (y l ,u k ) ∈ Y × U, l =1,...,L Δ , k =1,...,K Δ , define a grid on<br />

Y ×U parameterized by Δ > 0 in such a way that, for any (y, u) ∈ Y ×U, thereexists<br />

agridpoint(y l ′,u k ′) satisfying the inequality ||(y, u) − (y l ′,u k ′)|| ≤ aΔ, a=const.<br />

Let us define the polyhedral set W N,Δ (C, y 0 ) ⊂ R LΔ +K Δ<br />

by the equation<br />

{<br />

∑<br />

W N,Δ (C, y 0 ) def<br />

= γ = {γ l,k }≥0 : γ l,k =1, ∑ (∇φ i (y l ) T f(y l ,u k )<br />

l,k<br />

l,k<br />

}<br />

(7.11) + C(φ i (y 0 ) − φ i (y l )))γ l,k =0 ∀i =1, 2,...,N ,<br />

∑ K<br />

Δ<br />

where ∑ def<br />

l,k<br />

= ∑ L Δ<br />

l=1 k=1 and the indexation of the components of γ ∈ W N,Δ (C, y 0 )<br />

corresponds <strong>to</strong> the indexation of the grid points, and let us consider the following finite<br />

dimensional LP problem (in what follows we call this the (N,Δ)-problem)<br />

∑<br />

(7.12) min γ l,k g(y l ,u k ) def<br />

= g N,Δ (C, y 0 ).<br />

γ∈W N,Δ (C,y 0)<br />

l,k<br />

Let us also consider the finite dimensional LP problem, which is dual <strong>to</strong> (7.12),<br />

{<br />

(<br />

N∑<br />

N<br />

)<br />

∑<br />

max μ : μ ≤ λ i ∇φ i (y l ) T f(y l ,u k )+C λ i (φ<br />

(μ,λ)∈R 1 ×R } i (y 0 ) − φ i (y l ))<br />

N<br />

i=1<br />

i=1<br />

(7.13) + g(y l ,u k ) ∀(u l ,y k ) .<br />

Similarly <strong>to</strong> [22] and [30] (where the case C = 0 was studied) it can be established<br />

that, under natural conditions, the following results are true. First (cf. Propositions<br />

7and9in[30]),<br />

(7.14) lim lim ρ H(W N,Δ (C, y 0 ),W(C, y 0 )) = 0<br />

N→∞ Δ→0<br />

(7.15) ⇒ lim lim<br />

N→∞ Δ→0 gN,Δ (C, y 0 )=g ∗ (C, y 0 ).<br />

Second (cf. Proposition 6.3 in [22]), the function<br />

(7.16) ψ N,Δ (y) def<br />

=<br />

N∑<br />

i=1<br />

λ N,Δ<br />

i φ i (y),<br />

where λ N,Δ =(λ N,Δ<br />

i ) is a solution of the finite dimensional dual (7.13), solves problem<br />

(3.1) approximately in the sense that, for any δ>0,<br />

μ ∗ (C, y 0 ) − δ ≤ ∇ψ N,Δ (y) T f(y, u)+C(ψ N,Δ (y 0 ) − ψ N,Δ (y))<br />

(7.17) + g(y, u) ∀(y, u) ∈ Y × U<br />

if N is large enough and Δ is small enough. Third, under some additional conditions<br />

(see Theorem 7.1 in [22]), the feedback control defined by the equation<br />

(7.18) u N,Δ (y) def<br />

=argmin<br />

u∈U {∇ψN,Δ (y) T f(y, u)+g(y, u)}<br />

is near optimal in (1.3) (in the sense that the value of the objective function obtained<br />

with this control tends <strong>to</strong> the optimal one if Δ → 0andN →∞).<br />

We do not give the proofs of the above mentioned results in the present paper.<br />

They are very similar <strong>to</strong> the proofs of the corresponding results of [22] and [30] (also,<br />

they will be included in a separate publication). However, we will illustrate how these<br />

results can be used for a construction of near optimal control in (1.3) with a numerical<br />

example.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2505<br />

Consider problem (1.3) with the following data (a periodic optimization problem<br />

with similar data was considered in Example 1 of [30]; see also [6]):<br />

u ∈ U def<br />

=[−1, 1] ⊂ R 1 ,<br />

y =(y 1 ,y 2 ) ∈ Y def<br />

= {(y 1 ,y 2 ) | y 1 ∈ [−5, 5], y 2 ∈ [−8, 8]} ⊂ R 2 ;<br />

f(y, u) def<br />

=(y 2 , −4y 1 − 0.3y 2 + u ),<br />

g(u, y) def<br />

= u 2 − y1 2 .<br />

Let us formulate the (N,Δ)-problem (7.12) with the use of the monomials<br />

φ i1,i 2<br />

(y 1 ,y 2 ) def<br />

= y i1<br />

1 yi2 2 , i 1,i 2 =0, 1,...,K,<br />

as the functions φ i (y) (note that the number N in (7.11) is equal <strong>to</strong> (K +1) 2 − 1<br />

in this case) and with the use of the rectangular grid of size Δ. This problem was<br />

solved by the CPLEX solver for K =10(N = 120) and Δ = 0.0125, and for C =0.1,<br />

y 0 =(−3, −5). The optimal value obtained is g N,Δ (C, y 0 ) ≈−3.6014 . Using the<br />

solution of the dual problem, one can construct the function ψ N,Δ (y) and find the<br />

feedback control u N,Δ (y) (according <strong>to</strong> (7.16) and (7.18)). It can be easily seen that<br />

in this case, u N,Δ (y) allows an explicit representation,<br />

⎧<br />

⎪⎨<br />

(7.19) u N,Δ (y) =<br />

⎪⎩<br />

− 1 ∂ψ N,Δ (y 1,y 2)<br />

2 ∂y 2<br />

if |− 1 2<br />

−1 if − 1 2<br />

1 if − 1 2<br />

∂ψ N,Δ (y 1,y 2)<br />

∂y 2<br />

∂ψ N,Δ (y 1,y 2)<br />

|≤1,<br />

∂y 2<br />

< −1,<br />

∂ψ N,Δ (y 1,y 2)<br />

∂y 2<br />

> 1.<br />

Substituting this control in<strong>to</strong> system (1.1) and integrating it with MATLAB allows<br />

one <strong>to</strong> obtain the state-control trajec<strong>to</strong>ry. This trajec<strong>to</strong>ry and its projection on<strong>to</strong><br />

the state space (y 1 ,y 2 ) are depicted in Figures 1 and 2. Note that, as can be seen<br />

from these figures, the trajec<strong>to</strong>ries appear <strong>to</strong> converge <strong>to</strong> some periodic regime (“limit<br />

cycle”).<br />

The value of the objective function numerically evaluated on the state-control<br />

trajec<strong>to</strong>ry is −3.5921, which is close <strong>to</strong> the value of g N,Δ (α, y 0 ) indicated above (the<br />

error being of the order of the grid size). Thus, by (7.15), the feedback control defined<br />

by (7.19) is likely <strong>to</strong> be near optimal in the problem under consideration. Note that<br />

the points marked with dots in Figures 1 and 2 are the grid points that correspond<br />

<strong>to</strong> the positive (basic) components of the found optimal solution of (N,Δ)-problem<br />

(7.12) (denote it as γ ∗ = {γl,k ∗ }). This solution can serve as an approximation <strong>to</strong> a<br />

solution of the IDLP problem (1.4), which (provided that it is unique) coincides with<br />

the discounted occupational measure generated by the optimal state-control trajec<strong>to</strong>ry<br />

of (1.3). Hence, the fact that a certain component of γ ∗ is positive indicates<br />

that the optimal state-control trajec<strong>to</strong>ry attends a “small” vicinity of the grid point<br />

corresponding <strong>to</strong> this particular component. Consequently, the fact that the trajec<strong>to</strong>ries<br />

depicted in Figures 1 and 2 pass close <strong>to</strong> the marked points can be interpreted<br />

as another indication that these trajec<strong>to</strong>ries are close <strong>to</strong> the optimal ones.<br />

8. Conclusions. We have established the relationships between the optimal control<br />

problem (1.3) and the IDLP problem (1.4) (Theorem 4.4) based on the duality<br />

results (Theorem 3.1 and Lemma 3.2). We have shown that the IDLP problem (1.4)<br />

and its dual can be used for the analysis and solution of the optimal control problem<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2506 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

1<br />

0.5<br />

u 0<br />

−0.5<br />

−1<br />

−6<br />

−4−2<br />

0<br />

2<br />

4<br />

6<br />

y 1<br />

−8<br />

−6<br />

−4<br />

−2<br />

0<br />

y 2<br />

2<br />

4<br />

6<br />

8<br />

Fig. 1. Near optimal state-control trajec<strong>to</strong>ry.<br />

8<br />

6<br />

4<br />

2<br />

y 2<br />

0<br />

−2<br />

−4<br />

−6<br />

−8<br />

−6 −4 −2 0 2 4 6<br />

y 1<br />

Fig. 2. Near optimal state trajec<strong>to</strong>ry.<br />

(1.3). In particular we constructed necessary and sufficient optimality conditions for<br />

this problem (Proposition 4.6), we obtained results characterizing the viability kernel<br />

of Y (Propositions 5.1 and 5.2), and we indicated a way <strong>to</strong> use finite dimensional approximations<br />

of the IDLP problem (1.4) and its dual for finding a numerical solution<br />

of (1.3) (see section 7). Also, we have shown that the relationships between (1.3) and<br />

(1.4) can be used for establishing similar relationships between the optimal control<br />

problem (1.6) (with S →∞) and the IDLP problem (1.7) (Propositions 6.1 and 6.2).<br />

Appendix. In this section we will prove Theorem 3.1 and give an outline of the<br />

proof of Lemma 4.3.<br />

Proof of Theorem 3.1(iii). If the function ψ(·) satisfying (3.4) exists, then<br />

and, hence,<br />

min<br />

(y,u)∈Y ×U {−∇ψ(y)T f(y, u) − C(ψ(y 0 ) − ψ(y))} > 0<br />

(A.1)<br />

lim<br />

α→∞<br />

min {g(y, (y,u)∈Y ×U u)+α(−∇ψ(y)T f(y, u) − C(ψ(y 0 ) − ψ(y)))} = ∞.<br />

This implies that the optimal value of problem (3.1) is unbounded (μ ∗ (C, y 0 )=∞).<br />

Assume now that the optimal value of problem (3.1) is unbounded. That is, there<br />

exists a sequence (μ k ,ψ k (·)) such that lim k→∞ μ k = ∞,<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2507<br />

(A.2) μ k ≤ g(y, u)+∇ψ k (y) T f(y, u)+C(ψ k (y 0 ) − ψ k (y)) ∀(y, u) ∈ Y × U<br />

(A.3) ⇒ 1 ≤ 1 μ k<br />

g(y, u)+ 1 μ k<br />

(∇ψ k (y) T f(y, u)+C(ψ k (y 0 )−ψ k (y)))<br />

∀(y, u) ∈ Y ×U.<br />

For k large enough,<br />

1<br />

μ k<br />

|g(y, u)| ≤ 1 2<br />

for all (y, u) ∈ Y × U. Hence<br />

1<br />

2 ≤ 1 μ k<br />

(∇ψ k (y) T f(y, u)+C(ψ k (y 0 ) − ψ k (y))) ∀(y, u) ∈ Y × U.<br />

That is, the function ψ(y) def<br />

= − 1<br />

μ k<br />

ψ k (y) satisfies (3.4).<br />

ProofofTheorem3.1(i). From (3.2) it follows that, if W (C, y 0 ) is not empty, then<br />

the optimal value of problem (3.1) is bounded.<br />

Conversely, let us assume that the optimal value μ ∗ (C, y 0 )ofproblem(3.1)is<br />

bounded, and let us establish that W (C, y 0 ) is not empty. Assume that it is not true<br />

and W (C, y 0 ) is empty. Define the set Q by the equation<br />

{<br />

∫<br />

Q def<br />

= x =(x 1 ,x 2 ,...):x i = (∇φ i (y) T f(y, u)+C(φ i (y 0 ) − φ i (y)))γ(du, dy),<br />

Y ×U<br />

}<br />

(A.4) γ ∈P(Y × U) .<br />

It is easy <strong>to</strong> see that the set Q is a convex and compact subset of l 1 (the fact that Q<br />

is relatively compact in l 1 is implied by (7.3); the fact that it is closed follows from<br />

P(Y × U) being compact in the weak convergence <strong>to</strong>pology).<br />

By (7.2), the assumption that W (C, y 0 ) is empty is equivalent <strong>to</strong> the assumption<br />

that the set Q does not contain the “zero element” (0 ∉Q). Hence, by a separation<br />

theorem (see, e.g., [52, p. 59]), there exists ¯λ =(¯λ 1 , ¯λ 2 ,...) ∈ l ∞ such that<br />

∑<br />

0=¯λ(0) > max x∈Q<br />

¯λ i x i<br />

=max γ∈P(Y ×U)<br />

∫<br />

Y ×U<br />

i<br />

(∇ψ¯λ(y) T f(y, u)+C(ψ¯λ(y 0 ) − ψ¯λ(y)))γ(dy, du)<br />

=max (y,u)∈Y ×U {∇ψ¯λ(y) T f(y, u)+C(ψ¯λ(y 0 ) − ψ¯λ(y))},<br />

where ψ¯λ(y) = ∑ ¯λ i i φ i (y) (see (7.5)). This implies that the function ψ(y) def<br />

= ψ¯λ(y)<br />

satisfies (3.4), and, by Theorem 3.1(iii), μ ∗ (C, y 0 ) is unbounded. Thus, we have<br />

obtained a contradiction that proves that W (C, y 0 )isnotempty.<br />

Proof of Theorem 3.1(ii). By Theorem 3.1(i), if the optimal value of problem (3.1)<br />

is bounded, then W is not empty, and hence a solution <strong>to</strong> problem (1.4) exists.<br />

Define the set ˆQ ⊂R 1 × l 1 by the equation<br />

∫<br />

ˆQ def<br />

= {(θ, x) :θ ≥ g(y, u)γ(du, dy), x =(x 1 ,x 2 ,...),<br />

∫<br />

(A.5) x i =<br />

Y ×U<br />

Y ×U<br />

(∇φ i (y) T f(y, u)+C(φ i (y 0 ) − φ i (y)))γ(dy, du), γ ∈P(Y × U)}.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2508 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

The set ˆQ is convex and closed. Also, for any j =1, 2,..., the point (θ j , 0) ∉ ˆQ,<br />

def<br />

where θ j = g ∗ (C, y 0 ) − 1 j and 0 is the zero element of l 1. On the basis of a separation<br />

theorem (see [52, p. 59]), one may conclude that there exists a sequence (κ j ,λ j ) ∈<br />

R 1 × l ∞ ,j=1, 2,... (with λ j def<br />

=(λ j 1 ,λj 2 ,...)) such that<br />

(<br />

κ j g ∗ (C, y 0 ) − 1 )<br />

+ δ j ≤ inf<br />

j<br />

(θ,x)∈ ˆQ<br />

{<br />

κ j θ + ∑ i<br />

λ j i x i<br />

{ ∫<br />

= inf κ j θ + (∇ψ λ j (y) T f(y, u)+C(ψ λ j (y) − ψ λ j (y 0 )))γ(du, dy)<br />

γ∈P(Y ×U)<br />

Y ×U<br />

∫<br />

}<br />

(A.6)<br />

s.t. θ ≥ g(y, u)γ(du, dy) ,<br />

Y ×U<br />

where δ j > 0 for all j and ψ λ j (y) = ∑ i λj i φ i(y). From (A.6) it immediately follows<br />

that κ j ≥ 0. Let us show that κ j > 0. In fact, if it was not the case, one would obtain<br />

that<br />

∫<br />

0 0.<br />

Dividing (A.6) by κ j one can obtain that<br />

g ∗ (C, y 0 ) − 1 (<br />

j < g ∗ (C, y 0 ) − 1 )<br />

+ δj<br />

j κ j<br />

≤<br />

{∫ (<br />

min<br />

g(y, u)+ 1 ) }<br />

γ∈P(Y ×U) Y ×U κ j (∇ψ λ j (y)T f(y, u)+C(ψ λ j (y 0 ) − ψ λ j (y))) γ(dy, du)<br />

{<br />

= min g(y, u)+ 1 }<br />

(y,u)∈Y ×U<br />

κ j (∇ψ λ j (y)T f(y, u)+C(ψ λ j (y 0 ) − ψ λ j (y))) ≤ μ ∗ (C, y 0 )<br />

⇒ g ∗ (C, y 0 ) ≤ μ ∗ (C, y 0 ).<br />

The latter and (3.2) prove (3.3).<br />

Outline of the proof of Lemma 4.3. Let V stands for the set of all Lebesgue<br />

measurable relaxed controls v(·) :[0+∞) ↦→ P(U), and let V Y δ<br />

(y 0 ) ⊂Vbe such<br />

that, for every v(·) ∈V Y δ<br />

(y 0 ), the solution y(t, y 0 ,v(·)) of (4.16) satisfies the inclusion<br />

y(t, y 0 ,v(·)) ∈ Y δ for all t>0.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2509<br />

For any v ′ (·),v ′′ (·) ∈V, define<br />

∫<br />

∫<br />

(A.7) α ′ i(t) def<br />

= φ i (u)v ′ (t, du), α ′′<br />

i (t) def<br />

=<br />

U<br />

U<br />

φ i (u)v ′′ (t, du),<br />

where φ i (u),i =1, 2,..., is a sequence of Lipschitz continuous functions, which is<br />

dense in the unit ball of C(U). Let e m (·) :[0,T] → R 1 , m =1, 2,..., be a sequence<br />

of square integrable functions which is dense in L 2 [0,T], and let<br />

)<br />

(A.8)<br />

¯ζT (α ′ i (·),α′′<br />

def<br />

i (·)) =<br />

(∣<br />

∞∑ ∣∣∣∣ ∫ T<br />

∫ T<br />

2 −m e m (t)α ′ i (t)dt −<br />

m=1<br />

0<br />

0<br />

e m (t)α ′′<br />

i (t)dt ∣ ∣∣∣∣<br />

∧ 1<br />

,<br />

(A.9)<br />

ζ T (v ′ (·),v ′′ (·)) def<br />

=<br />

∞∑<br />

i=1<br />

2 −i ḡ T (α ′ i (·),α′′ i (·)),<br />

(A.10)<br />

ζ(v ′ (·),v ′′ (·)) def<br />

=<br />

∞∑<br />

2 −l ζ l (v ′ (·),v ′′ (·)).<br />

l=1<br />

It can be shown that ζ(·, ·) introduced in (A.10) defines a metric on V and that V is<br />

compact in this metric (see, e.g., Lemma 2.2 in [12]). Also it can be shown that the<br />

following two statements are valid: (i) V Y δ<br />

(y 0 ), V Y (y 0 )arecompactand<br />

(A.11) lim δ→0 V Y δ<br />

(y 0 ) ⊂V Y (y 0 );<br />

(ii) for any continuous q(y, u) :R m × U → R 1 and ¯q(y, u) :R m ×P(U) → R 1 defined<br />

as in (4.17), the functional<br />

(A.12)<br />

Ψ q (v(·)) def<br />

=<br />

∫ +∞<br />

0<br />

e −Ct¯q(y(t, y 0 ,v(·)),v(t))dt<br />

is continuous in v(·). From (A.11) and continuity of Ψ(v(·)) it follows that<br />

{<br />

}<br />

(A.13)<br />

min Ψ q (v(·)) ≤ lim δ→0 min Ψ q (v(·)) .<br />

v(·)∈V Y (y 0)<br />

By Assumption I,<br />

v(·)∈V Y δ (y 0)<br />

(A.14)<br />

min Ψ g (v(·)) = V<br />

v(·)∈V Y C Y (y 0),<br />

(y 0)<br />

where Ψ g (v(·)) is defined by (A.12) with q(·) =g(·). Since<br />

one now can use (A.13) <strong>to</strong> obtain that<br />

min Ψ g (v(·)) ≤ V Y δ<br />

v(·)∈V Y C (y 0) ∀ δ>0,<br />

δ (y 0)<br />

(A.15)<br />

V Y C (y 0) ≤ lim δ→0 V Y δ<br />

C (y 0).<br />

Since V Y δ<br />

C (y 0) ≤ V Y C (y 0) ∀δ >0, (A.15) implies (4.13).<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2510 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

Acknowledgment. We thank S. Rossomakhine for his help with solving the<br />

numerical example of section 7.<br />

REFERENCES<br />

[1] O. Alvarez and M. Bardi, Ergodic problems in differential games, in Advances in Dynamic<br />

Games Theory, S. Jorgensen, M. Quincampoix, and T.L. Vincent, eds., Birkhäuser Bos<strong>to</strong>n,<br />

Bos<strong>to</strong>n, MA, 2007, pp. 131–152.<br />

[2] E.J. Anderson and P. Nash, <strong>Linear</strong> <strong>Programming</strong> in <strong>Infinite</strong>-Dimensional Spaces, Wiley,<br />

Chichester, 1987.<br />

[3] M. Arisawa, H. Ishii, and P.-L. Lions, A characterization of the existence of solutions for<br />

Hamil<strong>to</strong>n–Jacobi equations in ergodic control problems with applications, Appl. Math. Optim.,<br />

42 (2000), pp 35–50.<br />

[4] Z. Artstein, Invariant measures and their projections in nonau<strong>to</strong>nomous dynamical systems,<br />

S<strong>to</strong>ch. Dyn., 4 (2004), pp. 439–459.<br />

[5] J.-P. Aubin, Viability Theory, Birkhäuser Bos<strong>to</strong>n, Bos<strong>to</strong>n, MA, 1991.<br />

[6] J. Azza<strong>to</strong> and J.B. Krawczyk, Applying a finite horizon numerical optimization method <strong>to</strong><br />

a periodic optimal control problem, Au<strong>to</strong>matica, 44 (2008), pp. 1642–1651.<br />

[7] M. Bardi and I. Capuzzo-Dolcetta, Optimal Control and Viscosity Solutions of Hamil<strong>to</strong>n–<br />

Jacobi–Bellman Equations, Systems Control Found. Appl., Birkhäuser Bos<strong>to</strong>n, Bos<strong>to</strong>n,<br />

MA, 1997.<br />

[8] A. Bensoussan, Perturbation Methods in Optimal Control, John Wiley, New York, 1989.<br />

[9] D.P. Bertsekas and S.E. Shreve, S<strong>to</strong>chastic Optimal Control: The Discrete Time Case,<br />

Academic Press, New York, 1978.<br />

[10] G.K. Basak, V.S. Borkar, and M.K. Ghosh, Ergodic control of degenerate diffusions,<br />

S<strong>to</strong>chastic Anal. Appl., 15 (1997), pp. 1–17.<br />

[11] A.G. Bhatt and V.S. Borkar, Occupation measures for controlled Markov processes: Characterization<br />

and optimality, Ann. Probab., 24 (1996), pp. 1531–1562.<br />

[12] V.S. Borkar and V. Gaitsgory, On averaging of singularly perturbed controlled s<strong>to</strong>chastic<br />

differential equations, Appl. Math. Optim., 56 (2007), pp. 169–209.<br />

[13] H. Brézis, Analyse fonctionnelle. Théorie et applications, Collection Mathématiques Appliquées<br />

pour la Maîtrise, Masson, Paris, 1983.<br />

[14] P. Cardaliaguet, P. Bettiol, and M. Quincampoix, Zero-sum state constraint differential<br />

game: Existence of a value for Bolza problem, Internet. J. Game Theory, 34 (2006), pp.<br />

495–527.<br />

[15] D.A. Carlson, A.B. Haurie, and A. Leizarowitz, Optimal Control on <strong>Infinite</strong> Time <strong>Horizon</strong>,<br />

2nd ed., Springer-Verlag, Berlin, 1991.<br />

[16] F.H. Clarke, Yu.S. Ledyaev, R.J. Stern, and P.R. Wolenski, Qualitative properties of<br />

trajec<strong>to</strong>ries of control systems: A survey, J. Dynam. Control Systems, 1 (1995), pp. 1–48.<br />

[17] F. Colonius and W. Kliemann, The Dynamics of Control, Birkhäuser Bos<strong>to</strong>n, Bos<strong>to</strong>n, MA,<br />

2000.<br />

[18] M. Crandall, H. Ishii, and P.-L. Lions, User’s guide <strong>to</strong> the viscosity solutions of second<br />

order partial differential equations, Bull. AMS, 27 (1992), pp. 1–67.<br />

[19] M.-O. Czarnecki and L.L. Rifford, Approximation and regularization of Lipschitz functions:<br />

Convergence of the gradients, Trans. Amer. Math. Soc., 358 (2006), pp. 4467–4520.<br />

[20] T.D. Donchev and A.L. Dontchev, Singular perturbations in infinite-dimensional control<br />

systems, SIAM J. Control Optim., 42 (2003), pp. 1795–1812.<br />

[21] L.C. Evans and D. Gomes, <strong>Linear</strong> programming interpretations of Mather’s variational principle,<br />

ESIAM Control Optim. Calc. Var., 8 (2002), pp. 693–702 (electronic).<br />

[22] L. Finlay, V. Gaitsgory, and I. Lebedev, Duality In linear programming problems related<br />

<strong>to</strong> deterministic long run average problems of optimal control, SIAM J. Control Optim.,<br />

47 (2008), pp. 1667–1700.<br />

[23] W.H. Fleming and H.M. Soner, Controlled Markov Processes and Viscosity Solutions,<br />

Springer-Verlag, New York, 1991.<br />

[24] W.H. Fleming and D. Vermes, Convex duality approach <strong>to</strong> the optimal control of diffusions,<br />

SIAM J. Control Optim., 27 (1989), pp. 1136–1155.<br />

[25] H. Frankowska, Hamil<strong>to</strong>n–Jacobi equations: Viscosity solutions and generalized gradients, J.<br />

Math. Anal. Appl., 141 (1989), pp. 21–26.<br />

[26] H. Frankowska and F. Rampazzo, Filippov’s and Filippov–Wazewski’s theorems on closed<br />

domains, J. Differential Equations, 161 (2000), pp. 449–478.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


LP IN OPTIMAL CONTROL PROBLEMS WITH DISCOUNTING 2511<br />

[27] V. Gaitsgory, Suboptimization of singularly perturbed control problems, SIAM J. Control<br />

Optim., 30 (1992), pp. 1228–1249.<br />

[28] V. Gaitsgory, On a representation of the limit occupational measures set of a control systems<br />

with applications <strong>to</strong> singularly perturbed control systems, SIAM J. Control Optim., 43<br />

(2004), pp. 325–340.<br />

[29] V. Gaitsgory, Averaging and near viability of singularly perturbed control systems, J.Convex<br />

Anal., 13 (2006), pp. 329–352.<br />

[30] V. Gaitsgory and S. Rossomakhine, <strong>Linear</strong> programming approach <strong>to</strong> deterministic long run<br />

average problems of optimal control, SIAM J. Control Optim., 44 (2006), pp. 2006–2037.<br />

[31] D. A. Gomes and A. M. Oberman, Computing the effective Hamil<strong>to</strong>nian using a variational<br />

approach, SIAM J. Control Optim., 43 (2004), pp. 792–812.<br />

[32] L. Grüne, On the relation between discounted and average optimal value functions, J.Differential<br />

Equations, 148 (1998), pp. 65–99.<br />

[33] L. Grüne, Asymp<strong>to</strong>tic controllability and exponential stabilization of nonlinear control systems<br />

at singular points, SIAM J. Control Optim., 36 (1998), pp. 1485–1503.<br />

[34] K. Helmes and R. H. S<strong>to</strong>ckbridge, Numerical comparison of controls and verification of<br />

optimality for s<strong>to</strong>chastic control problems, J. Optim. Theory Appl., 106 (2000), pp. 107–<br />

127.<br />

[35] D. Hernandez-Hernandez, O. Hernandez-Lerma, and M. Taksar, The linear programming<br />

approach <strong>to</strong> deterministic optimal control problems, Appl. Math., 24 (1996), pp. 17–33.<br />

[36] O. Hernandez-Lerma and J.B. Lasserre, Markov Chains and Invariant Probabilities,<br />

Birkhäuser-Verlag, Basel, 2003.<br />

[37] M. Kamenskii, P. Nistri, and M. Quincampoix, Singularly perturbed second order differential<br />

inclusions by an averaging method, Int. J. Pure Appl. Math., 3 (2002), pp. 343–355.<br />

[38] E. Klein and A.C. Thompson, Theory of Correspondence, Wiley-Interscience, New York,<br />

1984.<br />

[39] T.G. Kurtz and R.H. S<strong>to</strong>ckbridge, Existence of Markov controls and characterization of<br />

optimal Markov controls, SIAM J. Control Optim., 36 (1998), pp. 609–653.<br />

[40] H.J. Kushner, Weak Convergence Methods and Singularly Perturbed S<strong>to</strong>chastic Control and<br />

Filtering Problems, Birkhäuser Bos<strong>to</strong>n, Bos<strong>to</strong>n, MA, 1990.<br />

[41] H.J. Kushner and P.G. Dupuis, Numerical Methods for S<strong>to</strong>chastic Control Problems in Continuous<br />

Time, 2nd ed., Springer-Verlag, New York, 2001.<br />

[42] J.B. Lasserre, D. Henrion, C. Prieur, and E. Trélat, Nonlinear optimal control via occupation<br />

measures and LMI-relaxations, SIAM J. Control Optim., 47 (2008), pp. 1643–1666.<br />

[43] A. Leizarowitz, Optimal controls for diffusions in R d —A min-max, max-min formula for the<br />

minimal cost growth rate, J. Math. Anal. Appl., 149 (1990), pp. 180–209.<br />

[44] A. Leizarowitz and A.J. Zaslavski, <strong>Infinite</strong>-horizon discrete-time optimal control problems,<br />

J. Math. Sci., 116 (2003), pp. 3369–3386.<br />

[45] J.G. Llavona, Approximation of Continuously Differentiable Functions, Math. Stud. 130,<br />

North–Holland, Amsterdam, The Netherlands, 1986.<br />

[46] K.R. Parthasarathy, Probability Measures on Metric Spaces, Academic Press, New York,<br />

1967.<br />

[47] S. Pickenhain and V. Lykina, Sufficiency conditions for infinite horizon optimal control<br />

problems, in Recent Advances in Optimization, Lecture Notes in Econom. and Math. Systems<br />

563, Springer, Berlin, 2006, pp. 217–232.<br />

[48] M. Quincampoix, Differential inclusions and target problems, SIAM J. Control Optim., 30<br />

(1992), pp. 324–335.<br />

[49] M. Quincampoix and C. Rainer, S<strong>to</strong>chastic control and compatible subsets of constraints,<br />

Bull. Sci. Math., 129 (2005), pp. 39–55.<br />

[50] M. Quincampoix and F. Watbled, Averaging method for discontinuous Mayer’s problem of<br />

singularly perturbed control systems, Nonlinear Anal. Theory Methods Appl., 54 (2003),<br />

pp. 819–837.<br />

[51] J.E. Rubio, Control and Optimization. The <strong>Linear</strong> Treatment of Nonlinear Problems, Manchester<br />

University Press, Manchester, UK, 1985.<br />

[52] W. Rudin, Real and Complex Analysis, 3rd ed., McGraw-Hill, New York, 1987.<br />

[53] T.R. Rockafellar and R. Wets, Variational Analysis, Grundlehren Math. Wiss. 317,<br />

Springer-Verlag, Berlin, 1998.<br />

[54] R.H. S<strong>to</strong>ckbridge, Time-average control of Martingale problems. Existence of a stationary<br />

solution, Ann. Probab., 18 (1990), pp. 190–205.<br />

[55] R.H. S<strong>to</strong>ckbridge, Time-average control of martingale problems: A linear programming formulation,<br />

Ann. Probab., 18 (1990), pp. 206–217.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.


2512 VLADIMIR GAITSGORY AND MARC QUINCAMPOIX<br />

[56] A. Vigodner, Limits of Singularly Perturbed Control Problems: Dynamical Systems <strong>Approach</strong>,<br />

Thesis for the Degree of Doc<strong>to</strong>r of Philosophy, The Weizmann Institute of Science, Rehovot,<br />

Israel, 1995.<br />

[57] C. Villani, Topics in Optimal Transportation, Grad. Stud. Math. 58, AMS, Providence, RI,<br />

2003.<br />

[58] R. Vinter, Convex duality and nonlinear optimal control, SIAM J. Control Optim., 31 (1993),<br />

pp. 518–538.<br />

[59] J. Warga, Optimal Control of Differential and Functional Equations, Academic Press, New<br />

York, 1972.<br />

[60] G.G. Yin and Q. Zhang, Continuous-Time Markov Chains and Applications. A Singular<br />

Perturbation <strong>Approach</strong>, Springer, New York, 1997.<br />

Copyright © by SIAM. Unauthorized reproduction of this article is prohibited.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!