• Aucun résultat trouvé

Second order analysis for the optimal control of parabolic equations under control and final state constraints

N/A
N/A
Protected

Academic year: 2021

Partager "Second order analysis for the optimal control of parabolic equations under control and final state constraints"

Copied!
23
0
0

Texte intégral

(1)

HAL Id: hal-01096660

https://hal.archives-ouvertes.fr/hal-01096660

Preprint submitted on 18 Dec 2014

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

Second order analysis for the optimal control of parabolic equations under control and final state

constraints

Francisco José Silva

To cite this version:

Francisco José Silva. Second order analysis for the optimal control of parabolic equations under control and final state constraints. 2014. �hal-01096660�

(2)

Second order analysis for the optimal control of parabolic equations under control and final state constraints

F. J. Silva ∗†

December 15, 2014

Abstract

We consider the optimal control of a semilinear parabolic equation with pointwise bounds constraints on the control and finitely many integral constraints on the final state. Using the standard Robinson’s constraint qualification [37], we provide a second order necessary condition over a set of strictly critical directions. The main feature of this result is that the qualification condition needed for the second order analysis is the same as for classical finite-dimensional problems and does not imply the uniqueness of the Lagrange multiplier. We establish also a second order sufficient optimality condition which implies, for problems with a quadratic Hamiltonian, the equivalence between solutions satisfying the quadratic growth property in the L1 and L topologies.

Keywords. Optimal control, parabolic equations, box constraints for the control, finitely many con- straints for the state, Robinson constraint qualification, second order optimality conditions, quadratic Hamiltonian.

MSC. 49J20, 49K20, 35Q93.

1 Introduction

The study of second order optimality conditions for optimal control problems governed by semilinear elliptic or parabolic equations has received a considerable attention over the last two decades (see [26, 21, 6, 22, 13, 5, 17, 15, 18, 39, 19, 9, 20, 3] and the references therein). The main reason is that under the absence of convexity, this theory plays a crucial role for the analysis of stability properties of the problem at hand. As a consequence, second order analysis is fundamental for establishing the local convergence of the solutions of some finite dimensional discretizations as well as the convergence of some optimization methods in function spaces. We refer the reader to the recent monograph [41]

for a rather complete account of the theory and to [27, 31], and the references therein, for the analysis of some associated Newton-based methods.

The study of optimality conditions for optimal control problems has several difficulties depending on the type of constraints considered. A common feature is that, in contrast to finite-dimensional optimization problems where the norms are equivalent, a two-norm approach for second order opti- mality conditions is necessary (see [29, 36]). In fact, the expansion of the cost function in terms of

XLIM - DMI UMR CNRS 7252 Facult´e des Sciences et Techniques, Universit´e de Limoges (fran- cisco.silva@unilim.fr) +33 5 87506787

This research benefited from the support of the “FMJH Program Gaspar Monge in optimization and operation research”, and from the support to this program from EDF.

(3)

control perturbations shows that if a second order sufficient condition has to be supposed, the natural norm for the coercitivity condition of the associated quadratic form is the L2-norm. On the other hand, since the cost function is in generalC2 with respect to theL-norm (and not with respect to theL2-norm), the neighborhood on which local optimality can be expected is aL-neighborhood of the nominal point.

In this article we consider the optimal control problem of a semilinear parabolic equation, where bounds constraints are imposed on the control for almost all (a.a.) (t, x) and a finite number of equality and inequality constraints in integral form are imposed on the final state. Since the bounds constraints on the control are polyhedric, our framework for the constraints is a particular case of the one introduced in [13] and in [18]. The main difference with respect to [13] is that in our case the cost function and the finite dimensional constraints are not C2 in the L2 topology. We could have considered other types offinitely many constraints for the state, but we have decided to focus in the case of final state constraints because of its analogy with the classical setting in the optimal control of ODEs.

For a weak local solution ¯u, i.e. when the cost is minimized at ¯u locally on a neighborhood in L relative to the feasible set (see Definition 2.4(i)), we prove two types of second order necessary conditions. The first one is stated in Theorem 5.2, which is our main result. It shows that, under the classical Robinson constraint qualification, for every direction v which belongs to thestrict critical cone at ¯u, there exists a Lagrange multiplier associated to the finitely-many integral constraints, such that the associated quadratic form, which depends onλ, is non-negative atv. The strength of this result is that only the standard Robinson condition is assumed (see [37]), in analogy to finite dimensional optimization problems. The drawback is that the strict critical cone is in general smaller than the usual critical cone. However, under a strict complementary assumption associated to the finite dimensional problem appearing in Pontryagin’s principle, both cones coincide. We remark that, at least in the case of pure control constraints, the strict complementary assumption plays an important role in the asymptotic expansion of solutions under data perturbation (see [35, 12]).

The proof of Theorem 5.2 is essentially based on the ideas exposed in [10, 8], a precise second order expansion of the Lagrangian with respect to bounded perturbations of the local solution (see Section 3), some density arguments and Dmitruk’s lemma (see [23]).

The second type of second order necessary optimality condition is proved in Theorem 5.7. The main assumption of this result is a strong regularity property introduced in [40] and revisited in [13], which implies Robinson constraint qualification and also the uniqueness of the Lagrange multiplier associated to the final state constraints. The key point is that this assumption allows to approximate the critical cone by its intersection with the radial cone of the polyhedric constraints. This regularity condition has been applied in [13] in order to develop a second order theory for optimization problems with partially polyhedric constraints andC2data. The main advantage of the result in Theorem 5.7 is that the associated quadratic form is non-negative for all critical directions, but has the disadvantage that requires stronger assumptions than those expected from classical optimization theory. We point out that the same type of result can also be obtained by using the techniques in [15, 18], but under a stronger surjectivity assumption (see Remark 5.3(ii)). The problem of finding a second order necessary condition over the classical critical cone, without requiring additional assumptions besides Robinson condition, remains still open in the field of optimization of PDE systems.

Using classical results about the dependence of solutions of linear parabolic equations on the right hand side (which are recalled in Section 2) and the asymptotic expansions studied in Section 3, the proof of the second order sufficient condition for the weak quadratic growth property, under the assumption that the associated quadratic form is a Legendre form, does not present particular difficulties. We have chosen this approach since in the case of a quadratic Hamiltonian, the Legendre form assumption is automatically verified. As a consequence, under the assumption that the strict critical cone coincides with the usual critical cone, or alternatively that the stronger form of Robinson

(4)

condition is verified, we have the equivalence of the quadratic growth property with respect to the L and L1 topologies (see Section 6). Apparently, this characterization is new in the literature.

The paper is organized as follows. In Section 2 we fix some notations, we recall some basic facts about linear and semilinear parabolic equations and we state the optimal control problem under consideration together with the standard assumptions. Using the asymptotic expansions for the Lagrangian studied in Section 3, we provide in Sections 5 and 6 our main results. Namely, we prove the two types of second order necessary condition described above and also a sufficient second order optimality condition. The latter is crucial in the analysis of optimal control problems with quadratic Hamiltonians for which the equivalence between weak and L1,1 solutions satisfying the quadratic growth condition is established. We end our paper by explaining how our assumptions have to be modified in the case of unbounded controls.

2 Preliminaries

Let us fixT >0 and recall that given a Banach space (X,k·kX) ands [1,∞[, the spaceLs([0, T];X) (respectively, L([0, T];X)) consists in all measurable functions f : [0, T]X satisfying that

kfkLs([0,T];X):=

Z T

0

kf(t)ksXdt

!1s

<

respectively kfkL([0,T];X):= ess supt∈[0,T]kf(t)kX< .

We have that (Ls([0, T];X),k · kLs([0,T];X)) is a Banach space (see e.g. [34]). Let us also recall that W1,s([0, T];X) is the space of allX-valued distributionsf Ls([0, T];X) such that the vector valued distributional derivative tf can be identified with an element in Ls([0, T];X). Endowed with the norm kfkW1,s([0,T];X) := kfkLs([0,T];X)+k∂tfkLs([0,T];X), we have that W1,s([0, T];X) is a Banach space.

In this article, Ω Rd denotes a non-empty bounded open set with a smooth boundary. Set Q=]0, T[×Ω and Σ =]0, T[×∂Ω. For notational convenience, fors1,s2 [1,∞], we setLs1,s2(Q) :=

Ls1([0, T];Ls2(Ω)) and kfks1,s2 := kfkLs1([0,T];Ls2(Ω)). We also write k · ks1 := k · kLs1(Ω). Given α,β > 0 we recall that Cα(Ω) is the space of α- H¨older continuous functions in Ω and Cα,β(Q) is the space of functions defined on Q which are α-H¨older continuous with respect to t and β-H¨older continuous with respect tox.

We will also need the following notations: for any two Banach spaces X, Y paired in duality, h·,·iX,Y denotes the associated bilinear form. We denote R+ := {r R ; r 0} and R := {r R; r 0}. Given a Banach space (X,k · kX) and a setAX, we letdX(z, A) := inf{kzxkX ; x A}. For a distribution T in Ω we set xiT (i = 1, . . . , d) for the weak first order derivatives, with similar notations for higher order derivatives.

Given z0 L2(Ω),aL∞,∞(Q),v L2([0, T];H−1(Ω)) it is well known (see e.g. [34]) that the linear parabolic equation

tz∆z+a(t, x)z = v, inQ, z = 0, in Σ, z(0) = z0 in Ω,

(2.1) admits a unique solution z W(0, T) := L2([0, T];H01(Ω))W1,2([0, T];H−1(Ω)) in the following weak sense: for a.a. t[0, T] and q H01(Ω) we have that

h∂tz(t), qiH−1(Ω),H1

0(Ω)+h∇z(t),∇qiL2(Ω),L2(Ω)+ha(t)z(t), qiL2(Ω),L2(Ω) = hv(t), qiH−1(Ω),L2(Ω), z(0) = z0.

(2.2)

(5)

Note that the second line in (2.2) is meaningful in view of the embeddingW(0, T)֒C([0, T];L2(Ω)) (see e.g. [34]). Moreover, there exists a constantc >0 such that

kzkL2([0,T];H1(Ω))+k∂tzkL2([0,T];H−1(Ω))c kz0kL2(Ω)+kvkL2([0,T];H−1(Ω))

, (2.3)

and, in view of the integration by parts formula, for allq W(0, T) we have that RT

0

h−h∂tq(t), z(t)iH−1(Ω),H01(Ω)+h∇z(t),∇q(t)iL2(Ω),L2(Ω)+ha(t)z(t), q(t)iL2(Ω),L2(Ω)i dt

=hz(0), q(0)iL2(Ω),L2(Ω)− hz(T), q(T)iL2(Ω),L2(Ω)+RT

0 hv(t), q(t)iH−1(Ω),H01(Ω)dt. (2.4) Now, for s∈]1,∞[ consider the space

Vs:=Ls([0, T];W2,s(Ω)W01,s(Ω))W1,s([0, T];Ls(Ω)). (2.5) Endowed with the natural norm

kzkVs :=kzks,s+k∂tzks,s+

d

X

i=1

k∂xizks,s+

d

X

i,j=1

k∂xixjzks,s

Vs is a Banach space. It is well known that if the data z0, v are more regular than L2(Ω) and L2([0, T];H−1(Ω)), respectively, then more regularity can be obtained for the solutionz. We collect in the following proposition some useful results of this type. For simplicity we will assume that z0 0, but similar results can be established for the more general case z0 W2−

2 s,s

0 (Ω) (see [32, Chapter 4]).

Proposition 2.1. Let s∈]1,∞[. Assume that z0 0 and vLs,s(Q). Then, problem (2.1) admits a unique strong solution z[v]∈ Vs (i.e. the equation is satisfied almost everywhere and z[v](0) = 0).

Moreover, the following assertion hold true:

(i) There exists a constant cs>0 such that

kz[v]kVs cskvks,s. (2.6) (ii) There exists a constant c1 >0 such that

kz[v](·, T)k1+kz[v]k1,1c1kvk1,1. (2.7) (iii) The linear application v L2,2(Q) z[v] L2,2(Q)C([0, T];L2(Ω)) is continuous when L2,2(Q) and L2,2(Q)C([0, T];L2(Ω)) are respectively endowed with the weak and the strong (e.g.

with the normk · k2,2+k · kC([0,T];L2(Ω)) for the latter) topologies.

Proof. The existence and uniqueness of a strong solution and the estimates in assertion (i) are proved e.g. in [32, Chapter 4, Theorem 9.1]. Now, writing z=z[v] for notational convenience, consider the equation

−∂tq∆q+αq = sgn(z) inQ,

q = 0 in Σ,

q(T,·) = sgn(z(T,·)) in Ω.

(2.8) This equation has a unique weak solution q W(0, T). Moreover, since the terminal condition (which can be seen as an initial condition after the change of variablet =Tt) and the right hand side are bounded uniformly by 1, by the maximum principle for weak solutions (see [32, Chapter 3,

(6)

Theorem 7.1]) we get thatkqk∞,∞is bounded by a constantcindependent ofz. Therefore, by (2.4),

we obtain Z

Q

|z(t, x)|dtdx+ Z

|z(T, x)|dx= Z

Q

v(t, x)q(t, x)dtdx

which together withkqk∞,∞cimply (ii). Finally, if vnv weakly in L2,2(Q) by (i) we have that zn:=z[vn] is bounded inV2. Therefore, there exists ˆz∈ V2 such that, except for some subsequence, zn zˆ weakly in L2([0, T];H2(Ω)) and tzn tzˆ weakly in L2,2(Q). Therefore, passing to the limit in the first equation of (2.2) we get that ˆzsatisfies it and so ˆz=z[v] by the uniqueness of weak solutions. On the other hand, since H2(Ω) is compactly embedded in L2(Ω), by Aubin’s Theorem (see [2] and [33, Chapter 1, Th´eor`eme 5.1]), we have that zn z[v] strongly in L2,2(Q) and in C([0, T];L2(Ω)). The result follows.

Remark 2.2. Note that by the embeddingsVs֒Ls,s(Q) andVs֒C([0, T];Ls(Ω)) (fors∈]1,∞[), and the proof of Proposition 2.1(ii) (for s= 1)imply that

kzks,scskvks,s, max

t∈[0,T]kz(t,·)kscskvks,s, s[1,∞[. (2.9) Given uL∞,∞(Q) and ϕ:Q×R×RR, we consider the Cauchy problem

ty∆y+ϕ(t, x, y, u) = 0 inQ,

y = 0 in Σ, y(0,·) =y0(·) in Ω. (2.10) (H1)We assume that

(i) The initial conditiony0 belongs to W02−2s,s(Ω), with s >(d+ 2)/2.

(ii) The function ϕ is measurable and the following monotonicity-type property holds true: for all R >0 there exists c=c(R)>0 such that

−c 1 +|y|2

ϕ(t, x, y, u)y for a.a. (t, x, y)Q×R, |u| ≤R.

(iii) For a.a. (t, x) Q the function ϕ(t, x,·,·) is C1 and for all R > 0 there exists c = c(R) such that for a.a (t, x)Q

|ϕ(t, x, y, u)|+y(t, x, y, u)|+u(t, x, y, u)| ≤c if |(y, u)| ≤R.

We recall that y W(0, T) is a weak solution of (2.10) if for a.a. t∈]0, T[ and q H01(Ω) we have that

h∂ty(t), qiH−1(Ω),H1

0(Ω)+h∇y(t),∇qiL2(Ω),L2(Ω)+hϕ(t,·, y(t), u(t)), qiL2(Ω),L2(Ω) = 0,

y(0) = y0. (2.11) We say that y is a strong solution if there exists s ∈]1,∞[ such that y ∈ Vs(Q). The following well-posedness result for (2.10) holds true:

Proposition 2.3. Under assumption (H1) for anyuL∞,∞(Q), equation (2.10) admits a unique strong solutiony[u]. Moreover, there existsβ >0such thaty[u]Vs(Q)∩Cβ/2,β(Q)for alls∈]1,∞[.

Proof. The existence and uniqueness of a weak solution y[u] W(0, T) is based on standard argu- ments using the Garlekin’s approximation, for the existence, and Gronwall Lemma, for the uniqueness (see [28, Proposition 2.1], in our framework, and [14, Theorem 5.1] for the case of homogeneous Neu- mann boundary conditions). On the other hand, by (H1)(i) and the Sobolev embeddings (see e.g.

[1, 24, 25]) we have the existence of α > 0 such that y0 Cα(Ω), which, by the results in [4], implies the existence of β > 0 such that y[u] Cβ/2,β(Q). Therefore, by (H1)(iii) we have that φ(·,·, y[u](·,·), u(·,·))L∞,∞(Q) and so the Vs(Q) regularity is a direct consequence of (H1)(i) and Proposition 2.1(i).

(7)

Now, for nE, nI N let : Q×R×R R, Φ : Ω×R R, ΦE : Ω×R RnE and ΦI : Ω×RRnI be such that:

(H2)The function is measurable and satisfies the same assumption stated for ϕin(H1)(iii). We suppose thatψ= Φ, ΦE, ΦI is measurable and for a.a. x Ω we have that ψ(x,·) isC1 and there exists a constant c=c(R) such that if|y| ≤R then |ψ(x, y)| ≤cand y(x, y)| ≤c.

In order to perform a second order analysis we will also need the following assumption

(H3) For a.a. (t, x) Q the functions ϕ(t, x,·,·) and ℓ(t, x,·,·) are C2. Moreover, there exists c=c(R) such that for all (y, u), (y, u), satisfying that |(y, u)| ≤R and |(y, u)| ≤R, we have that

(y,u)2(t, x,0,0)|+|ℓ(y,u)2(t, x,0,0)| ≤c,

(y,u)2(t, x, y, u)ϕ(y,u)2(t, x, y, u)| ≤c(|yy|+|uu|),

|ℓ(y,u)2(t, x, y, u)(y,u)2(t, x, y, u)| ≤c(|yy|+|uu|).

Analogously, for ψ = Φ, ΦiE, ΦjI (i= 1, . . . , nE, j = 1, . . . , nI), we suppose that for all y, y with

|y| ≤R and|y| ≤R,

yy(t, x,0)| ≤c, yy(t, x, y)ψyy(t, x, y)| ≤c|yy|.

Let us define GE :L∞,∞(Q)RnE,GI :L∞,∞(Q)RnI as GiE(u) :=

Z

ΦiE(x, y[u](T, x))dx i= 1, . . . , nE, GjI(u) :=

Z

ΦjI(x, y[u](T, x))dx j = 1, . . . , nI, and letJ :L∞,∞(Q)Rand G:L∞,∞(Q)RnE×RnI be defined as

J(u) :=

Z

Q

ℓ(t, x, y[u](t, x), u(t, x))dxdt+ Z

Φ(x, y[u](T, x))dx, G(u) := (GE(u), GI(u)).

Note that under (H2)the above functions are well defined. Now, given a,b L∞,∞(Q) satisfying thataba.e. in Q and essinf(b(t, x)a(t, x))>0, let us define the sets

K1 = {uL∞,∞(Q)|a(t, x)u(t, x)b(t, x), for a.a.(t, x)Q},

K2 = {0}nE×RnI, K=K1G−1(K2). (2.12) We consider the problem

u∈L∞,∞inf (Q)J(u) subject to u∈ K. (P) Definition 2.4. (i) We say thatu¯∈ K is a weak local solution of(P) if there exists ε >0 such that J(u)J(¯u) for all u∈ K such thatkuuk¯ ∞,∞ε.

(ii) We say that J satisfies a local quadratic growth condition in the weak sense at u¯ ∈ K if there exists α, ε >0 such that

J(u)Ju) + α

2kuuk¯ 22,2 for all u∈ K such that kuuk¯ ∞,∞ε. (2.13) (iii) We say that J satisfies a local quadratic growth condition in the Ls-weak sense (s[1,∞[) at

¯

u∈ K if there existsα, ε >0 such that J(u)J(¯u) +α

2kuuk¯ 22,2 for all u∈ K such that kuuk¯ s,sε. (2.14) Remark 2.5. Since K is bounded in L∞,∞(Q), for any s[1,∞[the topology of Ls,s(Q) restricted toK is equivalent to the topology ofL1,1(Q) restricted toK (see[3, Remark 2.7]). This implies thatu¯ satisfies theLs,s-weak quadratic growth condition (2.14)iffu¯ satisfies theL1,1-weak quadratic growth condition.

(8)

3 Second order expansions for the Lagrangian associated to (P)

We first prove some Taylor type expansion for the function u L∞,∞(Q) y[u] ∈ VsC(Q) (s ∈]1,∞[). Let us fix ¯u L∞,∞(Q) and set ¯y := y[¯u], ¯ϕ(t, x) := ϕ(t, x,y(t, x),¯ u(t, x)). For¯ notational convenience we also write

¯

ϕy(t, x) :=ϕy(t, x,y(t, x),¯ u(t, x)),¯ ϕ¯u(t, x) :=ϕu(t, x,y(t, x),¯ u(t, x)),¯

¯

ϕ(y,u)(t, x)(z, v) = ¯ϕy(t, x)z+ ¯ϕu(t, x)v (z, v)R2, (3.1) with a similar notation for the second order derivatives evaluated at (t, x,y(t, x),¯ u(t, x)). Given¯ vL∞,∞(Q) definez1[v]∈ Vs(Q) as the unique solution of

tz∆z+ ¯ϕ(y,u)(t, x)(z, v) = 0 in Q, z = 0 in Σ, z(0,·) = 0 in Ω.

(3.2)

Note that under(H1)(iii), z1[v] is well-defined. We now prove some technical results:

Lemma 3.1. Let us set δy=y[¯u+v]y[¯u]. Then,

kz1[v]kVs+kδykVs = O(kvks,s), for all s[1,∞[,

kδyz1[v]ks,s = O(k(δy)2ks,s+kv2ks,s), for all s[1,∞[, kδy(T,·)z1[v](T,·)ks = O(k(δy)2ks,s+kv2ks,s) for all s[1,∞[.

(3.3)

Proof. The estimate forkz1[v]kVs is an immediate consequence of (3.2), Proposition 2.1(i) and (2.9).

Let us set δϕ(t, x) :=ϕ(t, x, y[¯u+v](t, x),u(t, x) +¯ v(t, x))ϕ(t, x). By definition and Proposition¯ 2.3 we have that δy satisfies the following equation in the strong sense

tδy∆δy+δϕ(t, x) = 0 inQ, δy = 0 in Σ, δy(0,·) = 0 in Ω.

(3.4)

Expanding δϕ(t, x), the first equation in (3.4) becomes

tδy∆δy+ Z 1

0

ϕ(y,u)(t, x,y¯+τ δy,u¯+τ v)(δy, v)dτ = 0, for a.a. (t, x)Q.

Therefore, by Proposition 2.1(i) and(H1)(iii), we obtain the estimates forkδykVsin the first equation of (3.3). Now, set d1 :=y[¯u+v]y[¯u]z1[v]. Then, d1 satisfies the equation

td1∆d1+ ¯ϕy(t, x)d1+δϕ(t, x)ϕ¯(y,u)(t, x)(δy, v) = 0 inQ, d1 = 0 in Σ d1(0,·) = 0 in Ω.

(3.5)

Expanding δϕ(t, x), as before, the first equation in (3.5) becomes

td1∆d1+ ¯ϕy(t, x)d1+ Z 1

0

ϕ(y,u)(t, x,y¯+τ δy,u¯+τ v)ϕ¯(y,u)(t, x)

(δy, v)dτ = 0. (3.6) and the second and third equation (3.5) follow from(H1)(iii), Proposition 2.1(i) and (2.9).

(9)

Now, we study second order expansions. Let us set z2[v]∈ Vs for the unique solution of

tz∆z+ ¯ϕy(t, x)z+12ϕ¯(y,u)2(t, x)(z1[v], v)2 = 0 in Q, z = 0 in Σ, z(0,·) = 0 in Ω.

(3.7)

Note that under(H3), the estimates forz1[v] in Lemma 3.1 imply thatz2[v] is well defined.

Lemma 3.2. For all s∈]1,∞[, we have the following estimates:

kz2[v]ks,s = O(k|z1(v)|2ks,s+k|v|2ks,s), ky[¯u+v]y[¯u]z1[v]z2[v]]k1,1 = O(kvk∞,∞kvk22,2),

ky[¯u+v](T,·)y[¯u](T,·)z1[v](T,·)z2[v]](T,·)k1 = O(kvk∞,∞kvk22,2).

(3.8)

Proof. As before, we set δy =y[¯u+v]y[¯u]. The first estimate follows directly from Proposition 2.1(i) and (H3). For notational convenience we setd2 := δyz1[v]z2[v] =d1z2[v]. By (3.5) and (3.7),d2 satisfies

td2∆d2+ ¯ϕy(t, x)d2+δϕ(t, x)ϕ¯(y,u)(t, x)(δy, v)12ϕ¯(y,u)2(t, x)(z1[v], v)2 = 0 in Q, (3.9) with zero boundary conditions. By a Taylor expansion we easily check, omitting the (t, x) argument,

δϕϕ¯(y,u)(δy, v)12ϕ¯(y,u)2(z1[v], v)2 = R1

0(1τ)

ϕ(y,u)2(t, x,y¯+τ δy,u¯+τ v)ϕ¯(y,u)2

(δy, v)2 +O(|d1||δy|+|d1||z1[v]|+|d1||v|),

the result follows from(H3), Proposition 2.1(ii), the estimates forδyand theL1-estimates ford1 in Lemma 3.1 (takes= 1 in the second and third equations of (3.1) and use thatkδyk∞,∞ckvk∞,∞, kz1[v]k∞,∞ckvk∞,∞, for somec >0 by the maximum principle).

Let us define the Lagrangian L:L∞,∞(Q)×RnE×RnI Rassociated to (P) as

L(u, λ) :=J(u) +λG(u) whereλ= (λE, λI)RnE ×RnI. (3.10) Using the estimates in Lemma 3.1 and Lemma 3.2 we study now a second order expansion forL(·, λ).

As forϕ, for notational convenience we set ¯ℓ(t, x) :=ℓ(t, x,y(t, x),¯ u(t, x)) and¯ ¯y(t, x) :=y(t, x,y(t, x),¯ u(t, x)),¯ ¯u(t, x) :=u(t, x,y(t, x),¯ u(t, x)),¯

with a similar notation for the second order derivatives. Givenλ= (λE, λI)RnE ×RnI we set Φ[λ](x, y) := Φ(x, y) +λEΦE(x, y) +λIΦI(x, y) for all yR, x

As before, we denote ¯Φ[λ](x) := Φ[λ](x,y(T, x)), with analogous notations for the derivatives. We¯ have the following second order expansion of the Lagrangian.

Proposition 3.3. LetCRnE×RnI be a nonempty compact set. Then, for everyλ= (λE, λI)C and vL∞,∞(Q) we have that

L(¯u+v, λ) = L(¯u, λ) +R

Q

¯y(t, x)[z1[v] +z2[v]] + ¯u(t, x)v+12¯(y,u)2(t, x)(z1[v], v)2 dtdx +R

Φ¯y[λ](x) [z1[v](T, x) +z2[v](T, x)] +12Φ¯yy[λ](x) (z1[v](T, x)])2 dx +O kvk∞,∞kvk22,2

,

(3.11)

where O kvk∞,∞kvk22,2

is uniform for λC.

Références

Documents relatifs

For exponentially stable normal semigroups (not necessarily ana- lytic), [JZ09, theorem 1.3] proves that the resolvent condition (3) implies final- observability in infinite time

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

[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

In sections 4 and 5 we prove our main results: the sufficient condition for quadratic growth in the strong sense, the characterization of this property in the case of pure

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

For second-order state constraints, a shooting approach can be used for the stability analysis if all the touch points are reducible, see [20, 8], but not in presence of

Abstract: This paper provides an analysis of Pontryagine mimina satisfying a quadratic growth condition, for optimal control problems of ordinary differential equations with

The contribution of this paper are first and second order necessary opti- mality conditions for an optimal control problem for a semilinear parabolic equation with cubic