• Aucun résultat trouvé

Second-order sufficient conditions for strong solutions to optimal control problems

N/A
N/A
Protected

Academic year: 2021

Partager "Second-order sufficient conditions for strong solutions to optimal control problems"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-00825260

https://hal.inria.fr/hal-00825260

Submitted on 23 May 2013

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires

optimal control problems

Joseph Frederic Bonnans, Xavier Dupuis, Laurent Pfeiffer

To cite this version:

Joseph Frederic Bonnans, Xavier Dupuis, Laurent Pfeiffer. Second-order sufficient conditions for

strong solutions to optimal control problems. ESAIM: Control, Optimisation and Calculus of Varia-

tions, EDP Sciences, 2014, 20 (03), pp.704-724. �10.1051/cocv/2013080�. �hal-00825260�

(2)

0249-6399ISRNINRIA/RR--8307--FR+ENG

RESEARCH REPORT N° 8307

May 2013

conditions for strong solutions to optimal control problems

J. Frédéric Bonnans, Xavier Dupuis, Laurent Pfeiffer

(3)
(4)

RESEARCH CENTRE SACLAY – ÎLE-DE-FRANCE 1 rue Honoré d’Estienne d’Orves Bâtiment Alan Turing

J. Frédéric Bonnans

, Xavier Dupuis

, Laurent Pfeier

Project-Team COMMANDS

Research Report n° 8307 May 2013 24 pages

Abstract: In this report, given a reference feasible trajectory of an optimal control problem, we say that the quadratic growth property for bounded strong solutions holds if the cost function of the problem has a quadratic growth over the set of feasible trajectories with a bounded control and with a state variable suciently close to the reference state variable. Our sucient second- order optimality conditions in Pontryagin form ensure this property and ensure a fortiori that the reference trajectory is a bounded strong solution. Our proof relies on a decomposition principle, which is a particular second-order expansion of the Lagrangian of the problem.

Key-words: Optimal control; second-order sucient conditions; quadratic growth; bounded strong solutions; Pontryagin multipliers; pure state and mixed control-state constraints; decompo- sition principle.

The research leading to these results has received funding from the EU 7th Framework Programme (FP7- PEOPLE-2010-ITN), under GA number 264735-SADCO, and from the Gaspard Monge Program for Optimization and operations research (PGMO).

Inria-Saclay and CMAP, Ecole Polytechnique, Route de Saclay, 91128 Palaiseau Cedex, France. Emails:

frederic.bonnans@inria.fr, xavier.dupuis@cmap.polytechnique.fr, laurent.pfeier@polytechnique.edu.

(5)

Résumé : Nous considérons dans ce rapport une trajectoire admissible d'un problème de commande optimale et disons que la propriété de croissance quadratique pour des solutions fortes est satisfaite si la fonction coût du problème a une croissance quadratique sur l'ensemble des trajectoires dont la commande est bornée et dont la variable d'état est susamment proche de la variable d'état de référence. Nos conditions d'optimalité du second ordre sous forme Pontryaguine garantissent cette propriété et a fortiori que la trajectoire de référence est une solution forte.

Notre preuve s'appuie sur un principe de décomposition, qui est un développement particulier du lagrangien du problème au second ordre.

Mots-clés : Commande optimale; conditions susantes du second ordre; croissance quadra-

tique; solutions fortes; multiplicateurs de Pontryaguine; contraintes pures sur l'état et contraintes

mixtes sur l'état et la commande; principe de décomposition.

(6)

1 Introduction

In this paper, we consider an optimal control problem with nal-state constraints, pure state constraints, and mixed control-state constraints. Given a feasible control u ¯ and its associated state variable y ¯ , we give second-order conditions ensuring that for all R > k uk ¯

, there exist ε > 0 and α > 0 such that for all feasible trajectory (u, y) with kuk

≤ R and ky − yk ¯

≤ ε ,

J (u, y) − J (¯ u, y) ¯ ≥ α(ku − uk ¯

22

+ |y

0

− y ¯

0

|

2

), (1.1) where J (u, y) is the cost function to minimize. We call this property quadratic growth for bounded strong solutions. Its specicity lies in the fact that the quadratic growth is ensured for controls which may be far from u ¯ in L

norm.

Our approach is based on the theory of second-order optimality conditions for optimization problems in Banach spaces [7, 13, 15]. A local optimal solution satises rst- and second-order necessary conditions; denoting by Ω the Hessian of the Lagrangian, theses conditions state that under the extended polyhedricity condition [6, Section 3.2], the supremum of Ω over the set of Lagrange multipliers is nonnegative for all critical directions. If the supremum of Ω is positive for nonzero critical directions, we say that the second-order sucient optimality conditions hold and under some assumptions, a quadratic growth property is then satised. This approach can be used for optimal control problems with constraints of any kind. For example, Stefani and Zezza [19] dealt with problems with mixed control-state equality constraints and Bonnans and Hermant [4] with problems with pure state and mixed control-state constraints. However, the quadratic growth property which is then satised holds for controls which are suciently close to u ¯ in uniform norm and only ensures that (¯ u, y) ¯ is a weak solution.

For Pontryagin minima, that is to say minima locally optimal in a L

1

neighborhood of u ¯ , the necessary conditions can be strengthened. The rst-order conditions are nothing but the well-known Pontryagin's principle, historically formulated in [18] and extended to problems with various constraints by many authors, such as Hestenes for problems with mixed control-state constraints [11] Dubovitskii and Osmolovskii for problems with pure state and mixed control- state constraints in early Russian references [9, 10], as highlighted by Dmitruk [8]. We refer to the survey by Hartl et al. for more references on this principle.

We say that the second-order necessary condition are in Pontryagin form if the supremum of Ω is taken over the set of Pontryagin multipliers, these multipliers being the Lagrange multipliers for which Pontryagin's principle holds. Maurer and Osmolovskii proved in [17] that the second- order necessary conditions in Pontryagin form were satised for Pontryagin minima to optimal control problems with mixed control-state equality constraints. They also proved that if second- order sucient conditions in Pontryagin form held, then the quadratic growth for bounded strong solutions was satised. The sucient conditions in Pontryagin form are as follows: the supremum of Ω over Pontryagin multipliers only is positive for nonzero critical directions and for all bounded neighborhood of u ¯ , there exists a Pontryagin multiplier which is such such the Hamiltonian has itself a quadratic growth. The results of Maurer and Osmolovskii are true under a restrictive full- rank condition for the mixed equality constraints, which is not satised by pure constraints, and under the Legendre-Clebsch condition, imposing that the Hessian of the augmented Hamiltonian w.r.t. u is positive. The full-rank condition enabled them to reformulate their their problem as a problem with nal-state constraints only. Note that these results were rst stated by Milyutin and Osmolovskii in [16], without proof.

For problems with pure and mixed inequality constraints, we proved the second-order neces-

sary conditions in Pontryagin form [2]; in the present paper, we prove that the sucient conditions

in Pontryagin form ensure the quadratic growth property for bounded strong solutions under the

Legendre-Clebsch condition. Our proof is based on an extension of the decomposition principle

(7)

of Bonnans and Osmolovskii [5] to the constrained case. This principle is a particular second- order expansion of the Lagrangian, which takes into account the fact that the control may have large perturbations in uniform norm. Note that the diculties arising in the extension of the principle and the proof of quadratic growth are mainly due to the presence of mixed control-state constraints.

The outline of the paper is as follows. In Section 2, we set our optimal control problem.

Section 3 is devoted to technical aspects related to the reduction of state constraints. We prove the decomposition principle in Section 4 (Theorem 4.2) and prove the quadratic growth property for bounded strong solutions in Section 5 (Theorem 5.3). In Section 6, we prove that under technical assumptions, the sucient conditions are not only sucient but also necessary to ensure the quadratic growth property (Theorem 6.3).

Notations. For a function h that depends only on time t , we denote by h

t

its value at time t , by h

i,t

the value of its i -th component if h is vector-valued, and by h ˙ its derivative. For a function h that depends on (t, x) , we denote by D

t

h and D

x

h its partial derivatives. We use the symbol D without any subscript for the dierentiation w.r.t. all variables except t , e.g. Dh = D

(u,y)

h for a function h that depends on (t, u, y) . We use the same convention for higher order derivatives.

We identify the dual space of R

n

with the space R

n∗

of n -dimensional horizontal vectors.

Generally, we denote by X

the dual space of a topological vector space X . Given a convex subset K of X and a point x of K , we denote by T

K

(x) and N

K

(x) the tangent and normal cone to K at x , respectively; see [6, Section 2.2.4] for their denition.

We denote by |·| both the Euclidean norm on nite-dimensional vector spaces and the cardinal of nite sets, and by k · k

s

and k · k

q,s

the standard norms on the Lesbesgue spaces L

s

and the Sobolev spaces W

q,s

, respectively.

We denote by BV ([0, T ]) the space of functions of bounded variation on the closed interval [0, T ] . Any h ∈ BV ([0, T ]) has a derivative dh which is a nite Radon measure on [0, T ] and h

0

(resp. h

T

) is dened by h

0

:= h

0+

− dh(0) (resp. h

T

:= h

T

+ dh(T) ). Thus BV ([0, T ]) is endowed with the following norm: khk

BV

:= kdhk

M

+ |h

T

| . See [1, Section 3.2] for a rigorous presentation of BV .

All vector-valued inequalities have to be understood coordinate-wise.

2 Setting

2.1 The optimal control problem

We formulate in this section the optimal control problem under study and we use the same frame- work as in [2]. We refer to this article for supplementary comments on the dierent assumptions made. Consider the state equation

˙

y

t

= f (t, u

t

, y

t

) for a.a. t ∈ (0, T ). (2.1) Here, u is a control which belongs to U , y is a state which belongs to Y , where

U := L

(0, T ; R

m

), Y := W

1,∞

(0, T ; R

n

), (2.2) and f : [0, T ] × R

m

× R

n

→ R

n

is the dynamics. Consider constraints of various types on the system: the mixed control-state constraints, or mixed constraints

c(t, u

t

, y

t

) ≤ 0 for a.a. t ∈ (0, T ), (2.3) the pure state constraints, or state constraints

g(t, y

t

) ≤ 0 for a.a. t ∈ (0, T ), (2.4)

(8)

and the initial-nal state constraints

( Φ

E

(y

0

, y

T

) = 0,

Φ

I

(y

0

, y

T

) ≤ 0. (2.5)

Here c : [0, T ] × R

m

× R

n

→ R

nc

, g : [0, T ] × R

n

→ R

ng

, Φ

E

: R

n

× R

n

→ R

nΦE

, Φ

I

: R

n

× R

n

→ R

nΦI

. Finally, consider the cost function φ : R

n

× R

n

→ R. The optimal control problem is then

(u,y)∈U ×Y

min φ(y

0

, y

T

) subject to (2.1)-(2.5) . ( P ) We call a trajectory any pair (u, y) ∈ U × Y such that (2.1) holds. We say that a trajectory is feasible for problem ( P ) if it satises constraints (2.3)-(2.5), and denote by F(P ) the set of feasible trajectories. From now on, we x a feasible trajectory (¯ u, y) ¯ .

Similarly to [19, Denition 2.1], we introduce the following Carathéodory-type regularity notion:

Denition 2.1. We say that ϕ : [0, T ] × R

m

× R

n

→ R

s

is uniformly quasi- C

k

i

(i) for a.a. t , (u, y) 7→ ϕ(t, u, y) is of class C

k

, and the modulus of continuity of (u, y) 7→

D

k

ϕ(t, u, y) on any compact of R

m

× R

n

is uniform w.r.t. t .

(ii) for j = 0, . . . , k , for all (u, y) , t 7→ D

j

ϕ(t, u, y) is essentially bounded.

Remark 2.2. If ϕ is uniformly quasi- C

k

, then D

j

ϕ for j = 0, . . . , k are essentially bounded on any compact, and (u, y) 7→ D

j

ϕ(t, u, y) for j = 0, . . . , k − 1 are locally Lipschitz, uniformly w.r.t.

t .

The regularity assumption that we need for the quadratic growth property is the following:

Assumption 1. The mappings f , c and g are uniformly quasi- C

2

, g is dierentiable, D

t

g is uniformly quasi- C

1

, Φ

E

, Φ

I

, and φ are C

2

.

Note that this assumption will be strengthened in Section 6.

Denition 2.3. We say that the inward condition for the mixed constraints holds i there exist γ > 0 and v ¯ ∈ U such that

c(t, u ¯

t

, y ¯

t

) + D

u

c(t, u ¯

t

, y ¯

t

)¯ v

t

≤ −γ, for a.a. t. (2.6) In the sequel, we will always make the following assumption:

Assumption 2. The inward condition for the mixed constraints holds.

Assumption 2 ensures that the component of the Lagrange multipliers associated with the mixed constraints belongs to L

(0, T ; R

nc

) , see e.g. [5, Theorem 3.1]. This assumption will also play a role in the decomposition principle.

2.2 Bounded strong optimality and quadratic growth

Let us introduce various notions of minima, following [16].

(9)

Denition 2.4. We say that (¯ u, y) ¯ is a bounded strong minimum i for any R > k¯ uk

, there exists ε > 0 such that

φ(¯ y

0

, y ¯

T

) ≤ φ(y

0

, y

T

), for all (u, y) ∈ F(P ) such that (2.7) ky − yk ¯

≤ ε and kuk

≤ R,

a Pontryagin minimum i for any R > k uk ¯

, there exists ε > 0 such that

φ(¯ y

0

, y ¯

T

) ≤ φ(y

0

, y

T

), for all (u, y) ∈ F(P ) such that (2.8) ku − uk ¯

1

+ ky − yk ¯

≤ ε and kuk

≤ R,

a weak minimum i there exists ε > 0 such that

φ(¯ y

0

, y ¯

T

) ≤ φ(y

0

, y

T

), for all (u, y) ∈ F(P ) such that (2.9) ku − uk ¯

+ ky − yk ¯

≤ ε.

Obviously, (2.7) ⇒ (2.8) ⇒ (2.9).

Denition 2.5. We say that the quadratic growth property for bounded strong solutions holds at (¯ u, y) ¯ i for all R > k¯ uk

, there exist ε

R

> 0 and α

R

> 0 such that for all feasible trajectory (u, y) satisfying kuk

≤ R and ky − yk ¯

≤ ε ,

φ(y

0

, y

T

) − φ(¯ y

0

, y ¯

T

) ≥ α

R

ku − uk ¯

22

. (2.10) The goal of the article is to characterize this property. If it holds at (¯ u, y) ¯ , then (¯ u, y) ¯ is a bounded strong solution to the problem.

2.3 Multipliers

We dene the Hamiltonian and the augmented Hamiltonian respectively by

H [p](t, u, y) := pf (t, u, y), H

a

[p, ν](t, u, y) := pf (t, u, y) + νc(t, u, y), (2.11) for (p, ν, t, u, y) ∈ R

n∗

× R

nc

× [0, T ] × R

m

× R

n

. We dene the end points Lagrangian by

Φ[β, Ψ](y

0

, y

T

) := βφ(y

0

, y

T

) + ΨΦ(y

0

, y

T

), (2.12) for (β, Ψ, y

0

, y

T

) ∈ R × R

nΦ

× R

n

× R

n

, where n

Φ

= n

ΦE

+ n

ΦI

and Φ =

Φ

E

Φ

I

. We set

K

c

:= L

(0, T ; R

nc

), K

g

:= C([0, T ]; R

ng

), K

Φ

:= {0}

RnΦE

× R

nΦI

, (2.13) so that the constraints (2.3)-(2.5) can be rewritten as

c(·, u, y) ∈ K

c

, g(·, y) ∈ K

g

, Φ(y

0

, y

T

) ∈ K

Φ

. (2.14) Recall that the dual space of C([0, T ]; R

ng

) is the space M([0, T ]; R

ng

) of nite vector-valued Radon measures. We denote by M([0, T ]; R

ng

)

+

the cone of positive measures in this dual space. Let

E := R × R

nΦ

× L

(0, T ; R

nc

) × M([0, T ]; R

ng

). (2.15)

(10)

Let N

Kc

(c(·, u, ¯ y)) ¯ be the set of elements in the normal cone to K

c

at c(·, u, ¯ y) ¯ that belong to L

(0, T ; R

nc

) , i.e.

N

Kc

(c(·, u, ¯ y)) := ¯

ν ∈ L

(0, T ; R

n+c

) : ν

t

c(t, ¯ u

t

, y ¯

t

) = 0 for a.a. t . (2.16) Let N

Kg

(g(·, y)) ¯ be the normal cone to K

g

at g(·, y) ¯ , i.e.

N

Kg

(g(·, y)) := ¯ (

µ ∈ M([0, T ]; R

ng

)

+

: Z

[0,T]

(dµ

t

g(t, y ¯

t

)) = 0 )

. (2.17)

Let N

KΦ

(Φ(¯ y

0

, y ¯

T

)) be the normal cone to K

Φ

at Φ(¯ y

0

, y ¯

T

) , i.e.

N

KΦ

(Φ(¯ y

0

, y ¯

T

)) :=

Ψ ∈ R

nΦ

: Ψ

i

≥ 0

Ψ

i

Φ

i

(¯ y

0

, y ¯

T

) = 0 for n

ΦE

< i ≤ n

Φ

. (2.18)

Finally, let

N (¯ u, y) := ¯ R

+

× N

KΦ

(Φ(¯ y

0

, y ¯

T

)) × N

Kc

(c(·, u, ¯ y)) ¯ × N

Kg

(g(·, y)) ¯ ⊂ E. (2.19) We dene the costate space

P := BV ([0, T ]; R

n∗

). (2.20)

Given λ = (β, Ψ, ν, µ) ∈ E , we consider the costate equation in P ( −dp

t

= D

y

H

a

[p

t

, ν

t

](t, u ¯

t

, y ¯

t

)dt + dµ

t

Dg(t, y ¯

t

),

p

T+

= D

yT

Φ[β, Ψ](¯ y

0

, y ¯

T

). (2.21) Denition 2.6. Let λ = (β, Ψ, ν, µ) ∈ E . We say that the solution of the costate equation (2.21) p

λ

∈ P is an associated costate i

−p

λ0

= D

y0

Φ[β, Ψ](¯ y

0

, y ¯

T

). (2.22) Let N

π

(¯ u, y) ¯ be the set of nonzero λ ∈ N(¯ u, y) ¯ having an associated costate.

We dene the set-valued mapping U : [0, T ] ⇒ R

m

by

U (t) := cl {u ∈ R

m

: c(t, u, y ¯

t

) < 0} for a.a. t, (2.23) where cl denotes the closure in R

m

. We can now dene two dierent notions of multipliers.

Denition 2.7. (i) We say that λ ∈ N

π

(¯ u, y) ¯ is a generalized Lagrange multiplier i

D

u

H

a

[p

λt

, ν

t

](t, u ¯

t

, y ¯

t

) = 0 for a.a. t. (2.24) We denote by Λ

L

(¯ u, y) ¯ the set of generalized Lagrange multipliers.

(ii) We say that λ ∈ Λ

L

(¯ u, y) ¯ is a generalized Pontryagin multiplier i

H [p

λt

](t, u ¯

t

, y ¯

t

) ≤ H[p

λt

](t, u, y ¯

t

) for all u ∈ U (t), for a.a. t. (2.25) We denote by Λ

P

(¯ u, y) ¯ the set of generalized Pontryagin multipliers.

Note that even if (¯ u, y) ¯ is a Pontryagin minimum, inequality (2.25) may not be satised for

some t ∈ [0, T ] and some u ∈ R

m

for which c(t, u, y ¯

t

) = 0 , as we show in [2, Appendix].

(11)

2.4 Reduction of touch points

Let us rst recall the denition of the order of a state constraint. For 1 ≤ i ≤ n

g

, assuming that g

i

is suciently regular, we dene by induction g

(j)i

: [0, T ] × R

m

× R

n

→ R, j ∈ N, by

g

i(j+1)

(t, u, y) := D

t

g

i(j)

(t, u, y) + D

y

g

i(j)

(t, u, y)f (t, u, y), g

i(0)

:= g

i

. (2.26) Denition 2.8. If g

i

and f are C

qi

, we say that the state constraint g

i

is of order q

i

∈ N i

D

u

g

(j)i

≡ 0 for 0 ≤ j ≤ q

i

− 1, D

u

g

(qi i)

6≡ 0. (2.27) If g

i

is of order q

i

, then for all j < q

i

, g

i(j)

is independent of u and we do not mention this dependence anymore. Moreover, the mapping t 7→ g

i

(t, y ¯

t

) belongs to W

qi,∞

(0, T ) and

d

j

dt

j

g

i

(t, y ¯

t

) = g

(j)i

(t, y ¯

t

) for 0 ≤ j < q

i

, (2.28) d

j

dt

j

g

i

(t, y ¯

t

) = g

(j)i

(t, u ¯

t

, y ¯

t

) for j = q

i

. (2.29) Denition 2.9. We say that τ ∈ [0, T ] is a touch point for the constraint g

i

i it is a contact point for g

i

, i.e. g

i

(τ, y ¯

τ

) = 0 , and τ is isolated in {t : g

i

(t, y ¯

t

) = 0} . We say that a touch point τ for g

i

is reducible i τ ∈ (0, T ) ,

dtd22

g

i

(t, y ¯

t

) is dened for t close to τ , continuous at τ , and

d

2

dt

2

g

i

(t, y ¯

t

)|

t=τ

< 0. (2.30) For 1 ≤ i ≤ n

g

, let us dene

T

g,i

:=

( ∅ if g

i

is of order 1,

{ touch points for g

i

} otherwise . (2.31) Note that for the moment, we only need to distinguish the constraints of order 1 from the other constraints, for which the order may be undened if g

i

or f is not regular enough.

Assumption 3. For 1 ≤ i ≤ n

g

, the set T

g,i

is nite and only contains reducible touch points.

2.5 Tools for the second-order analysis

We dene now the linearizations of the system, the critical cone, and the Hessian of the La- grangian. Let us set

V

2

:= L

2

(0, T ; R

m

), Z

1

:= W

1,1

(0, T ; R

n

), and Z

2

:= W

1,2

(0, T ; R

n

). (2.32) Given v ∈ V

2

, we consider the linearized state equation in Z

2

˙

z

t

= Df(t, u ¯

t

, y ¯

t

)(v

t

, z

t

) for a.a. t ∈ (0, T ). (2.33) We call linerarized trajectory any (v, z) ∈ V

2

×Z

2

such that (2.33) holds. For any (v, z

0

) ∈ V

2

× R

n

, there exists a unique z ∈ Z

2

such that (2.33) holds and z

0

= z

0

; we denote it by z = z[v, z

0

] . We also consider the second-order linearized state equation in Z

1

, dened by

ζ ˙

t

= D

y

f (t, u ¯

t

, y ¯

t

t

+ D

2

f (t, u ¯

t

, y ¯

t

)(v

t

, z

t

[v, z

0

])

2

for a.a. t ∈ (0, T ). (2.34)

(12)

We denote by z

2

[v, z

0

] the unique ζ ∈ Z

1

such that (2.34) holds and such that z

0

= 0 . The critical cone in L

2

is dened by

C

2

(¯ u, y) := ¯

 

 

 

 

(v, z) ∈ V

2

× Z

2

: z = z[v, z

0

] Dφ(¯ y

0

, y ¯

T

)(z

0

, z

T

) ≤ 0

DΦ(¯ y

0

, y ¯

T

)(z

0

, z

T

) ∈ T

KΦ

(Φ(¯ y

0

, y ¯

T

)) Dc(·, u, ¯ y)(v, z) ¯ ∈ T

Kc

(c(·, u, ¯ y)) ¯ Dg(·, y)z ¯ ∈ T

Kg

(g(·, y)) ¯

 

 

 

 

(2.35)

Note that by [6, Examples 2.63 and 2.64], the tangent cones T

Kg

(g(·, y)) ¯ and T

Kc

(c(·, u, ¯ y)) ¯ are resp. described by

T

Kg

= {ζ ∈ C([0, T ]; R

n

) : ∀t, g(t, y ¯

t

) = 0 = ⇒ ζ

t

≤ 0}, (2.36) T

Kc

= {w ∈ L

2

([0, T ]; R

m

) : for a.a. t, c(t, u ¯

t

, y ¯

t

) = 0 = ⇒ w

t

≤ 0} (2.37) Finally, for any λ = (β, Ψ, ν, µ) ∈ E , we dene a quadratic form, the Hessian of Lagrangian, Ω[λ] : V

2

× Z

2

→ R by

Ω[λ](v, z) :=

Z

T 0

D

2

H

a

[p

λt

, ν

t

](t, u ¯

t

, y ¯

t

)(v

t

, z

t

)

2

dt + D

2

Φ[β, Ψ](¯ y

0

, y ¯

T

)(z

0

, z

T

)

2

+ Z

[0,T]

t

D

2

g(t, y ¯

t

)(z

t

)

2

− X

τ∈Tg,i

1≤i≤ng

µ

i

(τ )

Dg

i(1)

(τ, y ¯

τ

)z

τ

2

g

(2)i

(τ, u ¯

τ

, y ¯

τ

)

. (2.38)

We justify the terms involving the touch points in T

g,i

in the following section.

3 Reduction of touch points

We recall in this section the main idea of the reduction technique used for the touch points of state constraints of order greater or equal than 2. Let us mention that this approach was described in [12, Section 3] and used in [14, Section 4] in the case of optimal control problems.

As shown in [3], the reduction allows to derive no-gap necessary and sucient second-order optimality conditions, i.e., the Hessian of the Lagrangian of the reduced problem corresponds to the quadratic form of the necessary conditions. We also prove a strict dierentiability property for the mapping associated with the reduction, that will be used in the decomposition principle.

Recall that for all 1 ≤ i ≤ n

g

, all touch points of T

g,i

are supposed to be reducible (Assumption 3). Let ε > 0 be suciently small so that for all 1 ≤ i ≤ n

g

, for all τ ∈ T

g,i

, the time function

t ∈ [τ − ε, τ + ε] 7→ g(t, y ¯

t

) (3.1) is C

2

and is such that for some β > 0 ,

ddt22

g

i

(t, y ¯

t

) ≤ −β , for all t in [τ − ε, τ + ε] . From now on, we set for all i and for all τ ∈ T

g,i

ετ

= [τ − ε, τ + ε] and ∆

εi

= [0, T ]\

τ∈Tg,i

ετ

, (3.2) and we consider the mapping Θ

ετ

: U × R

n

→ R dened by

Θ

ετ

(u, y

0

) := max {g

i

(t, y

t

) : y = y[u, y

0

], t ∈ ∆

ετ

}. (3.3)

(13)

We dene the reduced pure constraints as follows:

for all i ∈ {1, ..., n

g

} ,

( g

i

(t, y

t

) ≤ 0, for all t ∈ ∆

εi

, (i)

Θ

ετ

(u, y

0

) ≤ 0, for all τ ∈ T

g,i

. (ii) (3.4) Finally, we consider the following reduced problem, which is an equivalent reformulation of prob- lem ( P ), in which the pure constraints are replaced by constraint (3.4):

min

(u,y)∈U ×Y

φ(y

0

, y

T

) subject to (2.1) , (2.3) , (2.5) , and (3.4) . ( P

0

) Now, for all 1 ≤ i ≤ n

g

, consider the mapping ρ

i

dened by

ρ

i

: µ ∈ M([0, T ]; R

+

) 7→ µ

|∆ε

i

, (µ(τ))

τ∈Tg,i

∈ M(∆

εi

; R

+

) × R

|Tg,i|

. (3.5) Lemma 3.1. The mapping Θ

ετ

is twice Fréchet-dierentiable at (¯ u, y ¯

0

) with derivatives

ετ

(¯ u, y ¯

0

)(v, z

0

) = Dg

i

(τ, y ¯

τ

)z

τ

[v, z

0

], (3.6) D

2

Θ

ετ

(¯ u, y ¯

0

)(v, z

0

)

2

= D

2

g

i

(τ, y ¯

τ

)(z

τ

[v, z

0

])

2

+ Dg

i

(τ, y ¯

τ

)z

2τ

[v, z

0

]

Dg

i(1)

(τ, y ¯

τ

)z

τ

2

g

(2)i

(τ, u ¯

τ

, y ¯

τ

) . (3.7) and the following mappings dene a bijection between Λ

L

(¯ u, y) ¯ and the Lagrange multipliers of problem ( P

0

), resp. between Λ

P

(¯ u, y) ¯ and the Pontryagin multipliers of problem ( P

0

):

λ = β, Ψ, ν, µ

∈ Λ

L

(¯ u, y) ¯ 7→ β, Ψ, ν, (ρ

i

i

))

1≤i≤ng

(3.8) λ = β, Ψ, ν, µ

∈ Λ

P

(¯ u, y) ¯ 7→ β, Ψ, ν, (ρ

i

i

))

1≤i≤ng

. (3.9)

See [3, Lemma 26] for a proof of this result. Note that the restriction of µ

i

to ∆

εi

is associated with constraint (3.4(i)) and (µ

i

(τ))

τ∈Tg,i

with constraint (3.4(ii)). The expression of the Hessian of Θ

ετ

justies the quadratic form Ω dened in (2.38). Note also that in the sequel, we will work with problem P

0

and with the original description of the multipliers, using implicitly the bijections (3.8) and (3.9).

Now, let us x i and τ ∈ T

g,i

. The following lemma is a dierentiability property for the map- ping Θ

ετ

, related to the one of strict dierentiability, that will be used to prove the decomposition theorem.

Lemma 3.2. There exists ε > 0 such that for all u

1

and u

2

in U , for all y

0

in R

n

, if

ku

1

− uk ¯

1

≤ ε, ku

2

− uk ¯

1

≤ ε, and |y

0

− y ¯

0

| ≤ ε, (3.10) then

Θ

ετ

(u

2

, y

0

) − Θ

ετ

(u

1

, y

0

) = g(τ, y

τ

[u

2

, y

0

]) − g(τ, y

τ

[u

1

, y

0

]) + O ku

2

− u

1

k

1

(ku

1

− uk ¯

1

+ ku

2

− uk ¯

1

+ |y

0

− y ¯

0

|)

. (3.11)

An intermediate lemma is needed to prove this result. Consider the mapping χ dened as follows:

χ : x ∈ W

2,∞

(∆

ετ

) 7→ sup

t∈[τ−ε,τ+ε]

x

t

∈ R . (3.12)

Let us set x

0

= g

i

(·, y) ¯

|∆ε

τ

. Note that x ˙

0τ

= 0 .

(14)

Lemma 3.3. There exists α

0

> 0 such that for all x

1

and x

2

in W

2,∞

(∆

τ

) , if k x ˙

1

− x ˙

0

k

≤ α

0

and k x ˙

2

− x ˙

0

k

≤ α

0

, then

χ(x

2

) − χ(x

1

) = x

2

(τ ) − x

1

(τ)

+ O k x ˙

2

− x ˙

1

k

(k x ˙

1

− x ˙

0

k

+ k x ˙

2

− x ˙

0

k

)

. (3.13)

Proof. Let 0 < α

0

< βε and x

1

, x

2

in W

2,∞

(∆

τ

) satisfy the assumption of the lemma. Denote by τ

1

(resp. τ

2

) a (possibly non-unique) maximizer of χ(x

1

) (resp. χ(x

2

) ). Since

˙

x

1τ−ε

≥ x ˙

0τ−ε

− α

0

≥ βε − α

0

> 0 and x ˙

1τ+ε

≤ x ˙

0τ+ε

+ α ≤ −βε + α < 0, (3.14) we obtain that τ

1

∈ (τ − ε, τ + ε) and therefore that x ˙

1τ1

= 0 . Therefore,

β |τ

1

− τ | ≤ | x ˙

0τ

1

− x ˙

0τ

| = | x ˙

1τ

1

− x ˙

0τ

1

| ≤ k x ˙

1

− x ˙

0

k

(3.15) and then, |τ

1

− τ| ≤ k x ˙

1

− x ˙

0

k

/β . Similarly, |τ

2

− τ | ≤ k x ˙

2

− x ˙

0

k

/β . Then, by (3.15),

χ(x

2

) ≥ x

1

1

) + (x

2

1

) − x

1

1

))

= χ(x

1

) + (x

2

(τ) − x

1

(τ)) + O(k x ˙

2

− x ˙

1

k

1

− τ |) (3.16) and therefore, the l.h.s. of (3.13) is greater than the r.h.s. and by symmetry, the converse inequality holds. The lemma is proved.

Proof of Lemma 3.2. Consider the mapping

G

τ

: (u, y

0

) ∈ (U × R

n

) 7→ t ∈ ∆

τ

7→ g

i

(t, y

t

[u, y

0

])

∈ W

2,∞

(∆

τ

). (3.17) Since g

i

is not of order 1 and by Assumption 1, the mapping G

τ

is Lipschitz in the following sense : there exists K > 0 such that for all (u

1

, y

0,1

) and (u

2

, y

0,2

) ,

kG

τ

(u

1

, y

0,1

) − G

τ

(u

2

, y

0,2

)k

1,∞

≤ K(ku

2

− u

1

k

1

+ |y

0,2

− y

0,1

|). (3.18) Set α = α

0

/(2K) . Let u

1

and u

2

in U , let y

0

in R

n

be such that (3.10) holds. Then by Lemma 3.3 and by (3.18),

Θ

ετ

(u

2

, y

0

) − Θ

ετ

(u

1

, y

0

)

= χ(G

τ

(u

2

, y

0

)) − χ(G

τ

(u

1

, y

0

))

= g(y

τ

[u

2

, y

0

]) − g(y

τ

[u

1

, y

0

])

+ O ku

2

− u

1

k

1

(ku

2

− uk ¯

1

+ ku

1

− uk ¯

1

+ |y

0

− y ¯

0

|)

, (3.19)

as was to be proved.

4 A decomposition principle

We follow a classical approach by contradiction to prove the quadratic growth property for bounded strong solutions. We assume the existence of a sequence of feasible trajectories (u

k

, y

k

)

k

which is such that u

k

is bounded and such that ky

k

− yk ¯

→ 0 and for which the quadratic growth property does not hold. The Lagrangian function rst provides a lower estimate of the cost function φ(y

0k

, y

Tk

) . The diculty here is to linearize the Lagrangian, since we must consider large perturbations of the control in L

norm. To that purpose, we extend the decomposition principle of [5, Section 2.4] to our more general framework with pure and mixed constraints.

This principle is a partial expansion of the Lagrangian, which is decomposed into two terms:

Ω[λ](v

A,k

, z[v

A,k

, y

0k

− y ¯

0

]) , where v

A,k

stands for the small perturbations of the optimal control,

and a dierence of Hamiltonians where the large perturbations occur.

(15)

4.1 Notations and rst estimates

Let R > k¯ uk

, let (u

k

, y

k

)

k

be a sequence a feasible trajectories such that

∀k, ku

k

k

≤ R and ku

k

− uk ¯

2

→ 0. (4.1) This sequence will appear in the proof of the quadratic growth property. Note that the conver- gence of controls implies that ky

k

− yk ¯

→ 0 . We need to build two auxiliary controls u

A,k

and

˜

u

k

. The rst one, ˜ u

k

, is such that

( c(t, u ˜

kt

, y

kt

) ≤ 0, for a.a. t ∈ [0, T ],

k u ˜

k

− uk ¯

= O(ky

k

− yk ¯

). (4.2) The following lemma proves the existence of such a control.

Lemma 4.1. There exist ε > 0 and α ≥ 0 such that for all y ∈ Y with ky − yk ¯

≤ ε , there exists u ∈ U satisfying

ku − uk ¯

≤ αky − yk ¯

and c(t, u

t

, y

t

) ≤ 0, for a.a. t. (4.3) Proof. For all y ∈ Y , consider the mapping C

y

dened by

u ∈ U 7→ C

y

(u) = t 7→ c(t, u

t

, y

t

)

∈ L

(0, T ; R

ng

). (4.4) The inward condition (Assumption 2) corresponds to Robinson's constraint qualication for C

¯y

at u ¯ with respect to L

(0, T ; R

ng

) . Thus, by the Robinson-Ursescu stability theorem [6, Theorem 2.87], there exists ε > 0 such that for all y ∈ Y with ky − yk ¯

≤ ε , C

y

is metric regular at u ¯ with respect to L

(0, T ; R

ng

) . Therefore, for all y ∈ Y with ky − yk ¯

≤ ε , there exists a control u such that, for almost all t , c(t, u

t

, y

t

) ≤ 0 and

ku − uk ¯

= O dist (C

y

(¯ u), L

(0, T ; R

ng

))

= O(ky − yk ¯

).

This proves the lemma.

Now, let us introduce the second auxiliary control u

A,k

. We say that a partition (A, B) of the interval [0, T ] is measurable i A and B are measurable subset of [0, T ] . Let us consider a sequence of measurable partitions (A

k

, B

k

)

k

of [0, T ] . We dene u

A,k

as follows:

u

A,kt

= ¯ u

t

1

{t∈Bk}

+ u

kt

1

{t∈Ak}

. (4.5) The idea is to separate, in the perturbation u

k

− u ¯ , the small and large perturbations in uniform norm. In the sequel, the letter A will refer to the small perturbations and the letter B to the large ones. The large perturbations will occur on the subset B

k

.

For the sake of clarity, we suppose from now that the following holds:

 

 

(A

k

, B

k

)

k

is a sequence of measurable partitions of [0, T ] ,

|y

0k

− y ¯

0

| + ku

A,k

− uk ¯

→ 0,

|B

k

| → 0,

(4.6)

where |B

k

| is the Lebesgue measure of B

k

. We set

v

A,k

:= u

A,k

− u ¯ and v

B,k

:= u

k

− u

A,k

(4.7)

(16)

and we dene

δy

k

:= y

k

− y, ¯ y

A,k

:= y[u

A,k

, y

0k

], and z

A,k

:= z[v

A,k

, δy

k0

]. (4.8) Let us introduce some useful notations for the future estimates:

R

1,k

:= ku

k

− uk ¯

1

+ |δy

0k

|, R

2,k

:= ku

k

− uk ¯

2

+ |δy

k0

|, R

1,A,k

:= kv

A,k

k

1

+ |δy

0k

|, R

2,A,k

:= kv

A,k

k

2

+ |δy

0k

|, R

1,B,k

:= kv

B,k

k

1

, R

2,B,k

:= kv

B,k

k

2

.

(4.9)

Combining the Cauchy-Schwarz inequality and assumption (4.6), we obtain that

R

1,B,k

≤ R

2,B,k

|B

k

|

1/2

= o(R

2,B,k

). (4.10) Note that by Gronwall's lemma,

kδy

k

k

= O(R

1,k

) = O(R

2,k

) and kz

A,k

k

= O(R

1,A,k

) = O(R

2,k

). (4.11) Note also that

kδy

k

− (y

A,k

− y)k ¯

= O(R

1,B,k

) = o(R

2,k

) (4.12) and since ky

A,k

− (¯ y + z

A,k

)k

= O(R

22,k

) ,

kδy

k

− z

A,k

k

= o(R

2,k

). (4.13)

4.2 Result

We can now state the decomposition principle.

Theorem 4.2. Suppose that Assumptions 1, 2, and 3 hold. Let R > k¯ uk

, let (u

k

, y

k

)

k

be a sequence of feasible controls satisfying (4.1) and (A

k

, B

k

)

k

satisfy (4.6). Then, for all λ = (β, Ψ, ν, µ) ∈ Λ

L

(¯ u, y) ¯ ,

β(φ(y

k0

, y

kT

) − φ(¯ y

0

, y ¯

T

)) ≥

12

Ω[λ](v

A,k

, z

A,k

) +

Z

Bk

H [p

λt

](t, u

kt

, y ¯

t

) − H [p

λt

](t, u ˜

kt

, y ¯

t

)

dt + o(R

22,k

), (4.14) where Ω is dened by (2.38).

The proof is given at the end of the section, page 15. The basic idea to obtain a lower estimate of β(φ(y

0

, y

T

) − φ(¯ y

0

, y ¯

T

)) is classical: we dualize the constraints and expand up to the second order the obtained Lagrangian. However, the dualization of the mixed constraint is particular here, in so far as the nonpositive added term is the following:

Z

Ak

ν

t

(c(t, u

A,kt

, y

kt

) − c(t, u ¯

t

, y ¯

t

)) dt + Z

Bk

ν

t

(c(t, u ˜

kt

, y

kt

) − c(t, u ¯

t

, y ¯

t

)) dt, (4.15) where u ˜

k

and u

A,k

are dened by (4.2) and (4.5). In some sense, we do not dualize the mixed constraint when there are large perturbations of the control. By doing so, we prove that the contribution of the large perturbations is of the same order as the dierence of Hamiltonians appearing in (4.14). If we dualized the mixed constraint with the following term:

Z

T 0

ν

t

(c(t, u

kt

, y

kt

) − c(t, u ¯

t

, y ¯

t

)) dt, (4.16)

(17)

we would obtain for the contribution of large perturbations a dierence of augmented Hamilto- nians.

Let us x λ ∈ Λ

L

(¯ u, y) ¯ and let us consider the following two terms:

I

1k

= Z

T

0

−H

ya

[p

λt

](t, u ¯

t

, y ¯

t

)δy

tk

dt +

Z

Ak

(H

a

[p

λt

](t, u

A,kt

, y

kt

) − H

a

[p

λt

](t, u ¯

t

, y ¯

t

)) dt (4.17a) +

Z

Bk

(H

a

[p

λt

](t, u ˜

kt

, y

tk

) − H

a

[p

λt

](t, u ¯

t

, y ¯

t

)) dt (4.17b) +

Z

Bk

(H [p

λt

](t, u

kt

, y

tk

) − H [p

λt

](t, u ˜

kt

, y

tk

)) dt (4.17c) and

I

2k

= − Z

[0,T]

(dµ

t

Dg(t, y ¯

t

)δy

tk

) +

ng

X

i=1

Z

εi

(g

i

(t, y

tk

) − g

i

(t, y ¯

t

)) dµ

t,i

(4.18a)

+ X

τ∈Tg,i

1≤i≤ng

µ

i

(τ )(Θ

ετ

(u

k

, y

k0

) − Θ

ετ

(¯ u, y ¯

0

)). (4.18b)

Lemma 4.3. Let R > k¯ uk

, let (u

k

, y

k

)

k

be a sequence of feasible trajectories satisfying (4.1), and let (A

k

, B

k

)

k

satisfy (4.6). Then, for all λ ∈ Λ

L

(¯ u, y) ¯ , the following lower estimate holds:

β (φ(y

0k

, y

Tk

)−φ(¯ y

0

, y ¯

T

))

12

D

2

Φ[λ](¯ y

0

, y ¯

T

)(z

A,k0

, z

TA,k

)

2

+ I

1k

+ I

2k

+ o(R

22,k

). (4.19) Proof. Let λ ∈ Λ

L

(¯ u, y) ¯ . In view of sign conditions for constraints and multipliers, we rst obtain that

βφ(y

0k

, y

Tk

) − φ(¯ y

0

, y ¯

T

) ≥ Φ[β, Ψ](y

0

, y

T

) − Φ[β, Ψ](¯ y

0

, y ¯

T

) +

ng

X

i=1

Z

εi

(g

i

(t, y

tk

) − g

i

(t, y ¯

t

)) dµ

i,t

+ X

τ∈Tg,i

1≤i≤ng

µ

i

(τ)(Θ

ετ

(u

k

, y

k0

) − Θ

ετ

(¯ u, y ¯

0

))

+ Z

Ak

ν

t

(c(t, u

A,kt

, y

kt

) − c(t, u ¯

t

, y ¯

t

)) dt + Z

Bk

ν

t

(c(t, u ˜

kt

, y

kt

) − c(t, u ¯

t

, y ¯

t

)) dt. (4.20) Expanding the end-point Lagrangian up to the second order, and using (4.13), we obtain that

Φ[β, Ψ](y

k0

, y

kT

) − Φ[β, Ψ](¯ y

0

, y ¯

T

)

= DΦ[β, Ψ](¯ y

0

, y ¯

T

)(δy

0k

, δy

Tk

) +

12

D

2

Φ[β, Ψ](¯ y

0

, y ¯

T

)(δy

k0

, δy

kT

)

2

+ o(R

22,k

)

= p

λT

δy

kT

− p

λ0

δy

0k

+

12

D

2

Φ[λ](¯ y

0

, y ¯

T

)(z

A,k0

, z

TA,k

)

2

+ o(R

22,k

). (4.21)

(18)

Integrating by parts (see [3, Lemma 32]), we obtain that p

λT

δy

Tk

− p

λ0

δy

k0

=

Z

[0,T]

d p

λt

δy

kt

+ p

λt

δy ˙

kt

dt

= Z

T

0

− H

ya

(t, u ¯

t

, y ¯

t

)δy

tk

+ H (t, u

kt

, y

kt

) − H(t, u ¯

t

, y ¯

t

) dt

− Z

[0,T]

d µ

t

Dg(t, y ¯

t

)δy

kt

. (4.22)

The lemma follows from (4.20), (4.21), and (4.22).

A corollary of Lemma 4.3 is the following estimate, obtained with (4.2):

β(φ(y

k0

, y

Tk

) − φ(¯ y

0

, y ¯

T

)) (4.23)

≥ Z

T

0

H [p

λt

](t, u

kt

, y

tk

) − H [p

λt

](t, u ˜

kt

, y

tk

)

dt + O(kδy

k

k

)

= Z

T

0

H [p

λt

](t, u

kt

, y ¯

t

) − H [p

λt

](t, u ¯

t

, y ¯

t

)

dt + O(kδy

k

k

). (4.24) Proof of the decomposition principle. We prove Theorem 4.2 by estimating the two terms I

1k

and I

2k

obtained in Lemma 4.3.

B Estimation of I

1k

. Let show that

I

1k

= 1 2

Z

T 0

D

2

H

a

[p

λt

](t, u ¯

t

, y ¯

t

)(v

A,kt

, z

A,kt

)

2

dt +

Z

Bk

(H [p

λt

](t, u

kt

, y ¯

t

) − H [p

λt

](t, u ˜

kt

, y ¯

t

)) dt + o(R

22,k

). (4.25) Using (4.13) and the stationarity of the augmented Hamiltonian, we obtain that term (4.17a) is equal to

Z

Ak

H

ya

[p

λt

](t, u ¯

t

, y ¯

t

)δy

kt

dt + 1

2 Z

Ak

D

2

H

a

[p

λt

](t, u ¯

t

, y ¯

t

)(v

A,kt

, z

A,kt

)

2

dt + o(R

22,k

). (4.26) Term (4.17b) is negligible compared to R

22,k

. Since

Z

Bk

(H [p

λt

](t, u

kt

, y

tk

) − H [p

λt

](t, u ˜

kt

, y

tk

)) dt

− Z

Bk

(H [p

λt

](t, u

kt

, y ¯

t

) − H [p

λt

](t, u ˜

kt

, y ¯

t

)) dt = O(|B

k

|R

21,k

) = o(R

22,k

), (4.27) term (4.17c) is equal to

Z

Bk

(H [p

λt

](t, u

kt

, y ¯

t

) − H [p

λt

](t, u ˜

kt

, y ¯

t

)) dt + o(R

22,k

). (4.28)

(19)

The following term is also negligible:

Z

Bk

D

2

H

a

[p

λt

](t, u ¯

t

, y ¯

t

)(v

A,kt

, z

A,kt

)

2

dt = o(R

22,k

). (4.29) Finally, combining (4.17), (4.26), (4.28), and (4.29), we obtain (4.25).

B Estimation of I

2k

. Let us show that

I

2k

= 1 2

Z

[0,T]

t

D

2

g(t, y ¯

t

)(z

tA,k

)

2

− 1 2

X

τ∈Tg,i

1≤i≤ng

µ

i

(τ) (Dg

(1)i

(τ, y ¯

τ

)z

A,kτ

)

2

g

i(2)

(τ, u ¯

τ

, y ¯

τ

)

. (4.30)

Using (4.13), we obtain the following estimate of term (4.18a):

− X

τ∈Tg,i

1≤i≤ng

Z

ετ

Dg

i

(t, y ¯

t

)δy

tk

i,t

+ 1 2

ng

X

i=1

Z

εi

D

2

g

i

(t, y ¯

t

)(z

tA,k

)

2

t

+ o(R

22,k

). (4.31)

Remember that z

2

[v

A,k

, δy

0k

] denotes the second-order linearization (2.34) and that the following holds:

ky

A,k

− (¯ y + z[v

A,k

, δy

k0

] + z

2

[v

A,k

, δy

0k

])k

= o(R

22,k

). (4.32) Using Lemma 3.2 and estimate (4.13), we obtain that for all i , for all τ ∈ T

g,i

,

Θ

ετ

(u

k

, y

k0

) − Θ

ετ

(u

A,k

, y

0k

)

= g

i

(τ, y

τk

) − g

i

(τ, y

τA,k

) + O(R

1,B,k

(R

1,B,k

+ R

1,k

))

= Dg

i

(τ, y ¯

τ

)(y

τk

− y

A,kτ

) + o(R

22,k

)

= Dg

i

(τ, y ¯

τ

)(δy

kτ

− z

A,kτ

− z

2τ

[v

A,k

, δy

0k

]) + o(R

22,k

). (4.33) By Lemma 3.1,

Θ

ετ

(u

A,k

, y

0k

) − Θ

ετ

(¯ u, y ¯

0

)

= Dg

i

(τ, y ¯

τ

)(z

τA,k

+ z

τ2

[v

A,k

, δy

0k

]) + 1

2 D

2

g

i

(τ, y ¯

τ

)(z

τA,k

)

2

− 1 2

(D

y

g

(1)i

(τ, y ¯

τ

)z

A,kτ

)

2

) g

(2)i

(τ, u ¯

τ

, y ¯

τ

)

+ o(R

22,k

). (4.34) Recall that the restriction of µ

i

to ∆

ετ

is a Dirac measure at τ . Summing (4.33) and (4.34), we obtain the following estimate for (4.18b):

X

τ∈Tg,i

1≤i≤ng

h Z

ετ

Dg

i

(t, y ¯

t

)δy

kt

+ 1

2 D

2

g

i

(t, y ¯

t

)(z

tA,k

)

2

i,t

− 1 2

(Dg

i(1)

(τ, y ¯

τ

)z

τA,k

)

2

) g

(2)i

(τ, u ¯

τ

, y ¯

τ

)

i

+ o(R

22,k

). (4.35)

Combining (4.31) and (4.35), we obtain (4.30). Combining (4.25) and (4.30), we obtain the

result.

(20)

5 Quadratic growth for bounded strong solutions

We give in this section sucient second-order optimality conditions in Pontryagin form ensuring the quadratic growth property for bounded strong solutions. Our main result, Theorem 5.3, is proved with a classical approach by contradiction.

Assumption 4. There exists ε > 0 such that for all feasible trajectory (u, y) in (U × Y) with ky − yk ≤ ¯ ε , if (u, y) satises the mixed constraints, then there exists u ˆ such that

c(t, u ˆ

t

, y ¯

t

) ≤ 0, for a.a. t and ku − uk ˆ

= O(ky − yk ¯

). (5.1) This assumption is a metric regularity property, global in u and local in y . Note that the required property is dierent from (4.2).

Denition 5.1. A quadratic form Q on a Hilbert space X is said to be a Legendre form i it is weakly lower semi-continuous and if it satises the following property: if x

k

* x weakly in X and Q(x

k

) → Q(x) , then x

k

→ x strongly in X .

Assumption 5. For all λ ∈ Λ

P

(¯ u, y) ¯ , Ω[λ] is a Legendre form.

Remark 5.2. By [3, Lemma 21], this assumption is satised if for all λ ∈ Λ

P

(¯ u, y) ¯ , there exists γ > 0 such that for almost all t ,

γ ≤ D

uu2

H

a

[p

λt

, ν

t

](t, u ¯

t

, y ¯

t

). (5.2) In particular, in the absence of mixed and control constraints, the quadratic growth of the Hamiltonian (5.4) implies (5.2).

For all R > k¯ uk

, we dene Λ

RP

(¯ u, y) = ¯

λ ∈ Λ

L

(¯ u, y) : ¯ for a.a. t , for all u ∈ U (t) with |u| ≤ R,

H[p

λt

](t, u, y ¯

t

) − H[p

λt

](t, u ¯

t

, y ¯

t

) ≥ 0 . (5.3) Note that Λ

P

(¯ u, y) = ¯ ∩

R>k¯uk

Λ

RP

(¯ u, y) ¯ . Remember that C

2

(¯ u, y) ¯ is the critical cone in L

2

, dened by (2.35).

Theorem 5.3. Suppose that Assumptions 1-5 hold. If the following second-order sucient conditions hold: for all R > k uk ¯

,

1. there exist α > 0 and λ

∈ Λ

RP

(¯ u, y) ¯ such that

( for a.a. t , for all u ∈ U (t) with |u| ≤ R,

H [p

λt

](t, u

t

, y ¯

t

) − H[p

λt

](t, u ¯

t

, y ¯

t

) ≥ α|u − u ¯

t

|

22

, (5.4) 2. for all (v, z) ∈ C

2

\{0} , there exists λ ∈ Λ

RP

(¯ u, y) ¯ such that Ω[λ](v, z) > 0 ,

then the quadratic growth property for bounded strong solutions holds at (¯ u, y) ¯ .

Proof. We prove this theorem by contradiction. Let R > k¯ uk

, let us suppose that there exists a sequence (u

k

, y

k

)

k

of feasible trajectories such that ku

k

k

≤ R , ky

k

− yk ¯

→ 0 and

φ(y

0k

, y

kT

) − φ(¯ y

0

, y ¯

T

) ≤ o(ku

k

− uk ¯

22

+ |y

0k

− y ¯

0

|

2

). (5.5)

We use the notations introduced in (4.9). Let λ

= (β

, Ψ

, ν

, µ

) ∈ Λ

RP

(¯ u, y) ¯ be such that (5.4)

holds.

Références

Documents relatifs

Necessary conditions for optimal terminal time control problems governed by a Volterra integral equation. Jacobi type conditions for

Key-words: Optimal control; pure state and mixed control-state constraints; Pontryagin's principle; Pontryagin multipliers; second-order necessary conditions; partial relaxation..

Key-words: Stochastic optimal control, variational approach, first and second order optimality conditions, polyhedric constraints, final state constraints.. ∗ INRIA-Saclay and CMAP,

[6] et Aronna [5], proved an integral and a pointwise forms of second order necessary and sufficient optimality conditions for bang-singular optimal controls with endpoint

Using the latter gen- eralized theorem, we derive second-order necessary conditions for a feasible solution to be an efficient solution to a multiobjective optimization problem..

Orbital transfer, time optimal control, Riemannian systems with drift, conjugate points.. 1 Available

We now apply the previous results in the framework of optimal control theory, in order to derive necessary and/or sufficient second order optimality conditions.. L ∞ local

state constraints, second order analysis, Goh transform, control-affine problems 11.. AMS