Home Sensitivity analysis for optimal control problems governed by nonlinear evolution inclusions
Article Open Access

Sensitivity analysis for optimal control problems governed by nonlinear evolution inclusions

  • Nikolaos S. Papageorgiou , Vicenţiu D. Rădulescu EMAIL logo and Dušan D. Repovš
Published/Copyright: June 1, 2016

Abstract

We consider a nonlinear optimal control problem governed by a nonlinear evolution inclusion and depending on a parameter λ. First we examine the dynamics of the problem and establish the nonemptiness of the solution set and produce continuous selections of the solution multifunction ξS(ξ) (ξ being the initial condition). These results are proved in a very general framework and are of independent interest as results about evolution inclusions. Then we use them to study the sensitivity properties of the optimal control problem. We show that we have Hadamard well-posedness (continuity of the value function), and we establish the continuity properties of the optimal multifunction. Finally, we present an application on a nonlinear parabolic distributed parameter system.

MSC 2010: 49J24; 49K20; 34G20

1 Introduction

One of the important problems in optimal control theory is the study of the variations of the set of optimal state-control pairs and of the value of the problem when we perturb the dynamics, the cost functional and the initial condition of the problem. Such a sensitivity analysis (also known in the literature as “variational analysis”) is important because it gives information about the tolerances which are permitted in the specification of the mathematical models, it suggests ways to solve parametric problems and it can also be useful in the computational analysis of the problem. For infinite dimensional systems (distributed parameter systems), such investigations were conducted for linear [8, 13, 24, 31], semilinear [30, 38] and nonlinear systems [23, 32, 33]. We also mention the books of Buttazzo [7], Dontchev and Zolezzi [17], Ito and Kunisch [25], and Sokolowski and Zolezio [39] (the latter for shape optimization problems). In this paper we conduct such an analysis for a very general class of systems driven by nonmonotone evolution inclusions.

So, let T=[0,b] be the time interval and (X,H,X*) an evolution triple of spaces (see Section 2). We assume that XH compactly. The space of controls is modelled by a separable reflexive Banach space Y, and E is a compact metric space that corresponds to the parameter space. As we have already mentioned, we consider systems monitored by evolution inclusions. These inclusions represent a way to model systems with deterministic uncertainties, see the books of Aubin and Frankowska [2], Fattorini [18] and Roubicek [37].

The problem under consideration is the following:

(1.1){J(x,u,λ)=0bL(t,x(t),λ)𝑑t+0bH(t,u(t),λ)𝑑t+ψ^(ξ,x(b),λ)inf=m(ξ,λ),-x(t)Aλ(t,x(t))+F(t,x(t),λ)+G(t,u(t),λ)for almost all tT,x(0)=ξ,u(t)U(t,λ)for almost all tT,λE.

In this problem,

Aλ:T×X2X*for every λ>0,F:T×H×E2H{},G:T×Y×E2H{},

and the precise conditions on them will be given in Section 4. For every initial state ξH and every parameter λE, we denote the set of admissible state-control pairs (that is, pairs (x,u) which satisfy the dynamics and the constraints of problem (1.1)) by Q(ξ,λ). We investigate the dependence of Q(ξ,λ) on the two variables (ξ,λ)H×E. Also, Σ(ξ,λ) denotes the set of optimal state-control pairs (that is, (x*,u*)Σ(ξ,λ) such that J(x*,u*,ξ,λ)=m(ξ,λ)). So, Σ(ξ,λ)Q(ξ,λ). We establish the nonemptiness of the set Σ(ξ,λ) and examine the continuity properties of the value function (ξ,λ)m(ξ,λ) and of the multifunction (ξ,λ)Σ(ξ,λ).

The nonemptiness and other continuity and structural properties of the set Q(ξ,λ) are consequences of general results about evolution inclusions, which we prove in Section 3 and which are of independent interest. The class of evolution inclusions considered in Section 3 is more general than the classes studied by Chen, Wang and Zu [11], Denkowski, Migorski and Papageorgiou [14], Liu [28], and Papageorgiou and Kyritsi [34].

In the next section, for the convenience of the reader, we review the main mathematical tools which we will need in this paper.

2 Mathematical background

Suppose that V and Z are Banach spaces and assume that V is embedded continuously and densely into Z (denoted by VZ). Then it is easy to check that:

  1. Z* is embedded continuously into V*,

  2. if V is reflexive, then Z*V*.

Having this observation in mind, we can introduce the notion of evolution triple of spaces, which is central in the class of evolution equations considered here.

Definition 2.1

A triple (X,H,X*) of spaces is said to be an “evolution triple” (or “Gelfand triple” or “spaces in normal position”) if the following hold:

  1. X is a separable reflexive Banach space and X* is its topological dual.

  2. H is a separable Hilbert space identified with its dual H*=H (pivot space).

  3. XH.

According to the remark made in the beginning of this section, we also have H*=HX*. In this paper we also assume that the embedding of X into H is compact. Hence, by Schauder’s theorem (see, for example, [20, Theorem 3.1.22]), so is the embedding of H*=H into X*. In what follows, by (resp. ||, *) we denote the norm of the space X (resp. H,X*). By , we denote the duality brackets for the pair (X*,X) and by (,) the inner product of the Hilbert space H. We know that

,|H×X=(,).

Also, let β>0 be such that

(2.1)||β.

We introduce the following space which has a central role in the study of the evolution inclusions. So, let 1<p< and set

Wp(0,b)={xLp(T,X):xLp(T,X*)}(1p+1p=1).

In this definition the derivative of x is understood in the sense of vectorial distributions (weak derivative). In fact, if we view x as an X*-valued function, then x() is absolutely continuous, hence strongly differentiable almost everywhere. Therefore,

Wp(0,b)AC1,p(T,X*)=W1,p((0,b),X*).

The space Wp(0,b), equipped with the norm

xWp=xLp(T,X)+xLp(T,X*)for all xWp(0,b),

becomes a separable reflexive Banach space. We know that

(2.2)Wp(0,b)C(T,H),
(2.3)Wp(0,b)Lp(T,H) compactly.

The following integration by parts formula is very helpful.

Proposition 2.2

If x,yWp(0,b), then t(x(t),y(t)) is absolutely continuous and

ddt(x(t),y(t))=x(t),y(t)+x(t),y(t)for almost all tT.

We know that for all 1p<,

Lp(T,X)*=Lp(T,X*)

with p=+ if p=1 (see [20, Theorem 2.2.9]).

Now, let (Ω,Σ) be a measurable space and V a separable Banach space. We introduce the following hyperspaces:

Pf(c)(V)={CV:C is nonempty, closed, (convex)},
P(w)k(c)(V)={CV:C is nonempty, (weakly-)compact, (convex)}.

Given a multifunction F:Ω2V{}, the “graph” of F is the set

GrF={(ω,v)Ω×V:vF(ω)}.

We say that F() is “graph measurable” if GrFΣ×B(V) with B(V) being the Borel σ-field of V. If μ() is a σ-finite measure on Σ and F:Ω2V{} is graph measurable, then the Yankov–von Neumann–Aumann selection theorem (see [22, Theorem 2.14, p. 158]) implies that F() admits a measurable selection, that is, there exists a Σ-measurable function f:ΩV such that f(ω)F(ω) μ-almost everywhere. In fact, there is a whole sequence {fn}n1 of such measurable selections such that F(ω){fn(ω)}¯ μ-almost everywhere (see [22, Proposition 2.17, p. 159]). Moreover, the above results are valid if V is only a Souslin space. Recall that a Souslin space need not be metrizable (see [21, p. 232]).

A multifunction F:ΩPf(V) is said to be “measurable” if for all yV, the function

ωd(y,F(ω))=infvF(ω)y-vV

is Σ-measurable. A multifunction F:ΩPf(V) which is measurable is also graph measurable. The converse is true if (Ω,Σ) admits a complete σ-finite measure μ. If (Ω,Σ,μ) is a σ-finite measure space and F:Ω2V{} is a multifunction, then for 1p we introduce the set

SFP={fLp(Ω,Y):f(ω)F(ω)μ-a.e.}.

Evidently, SFP if and only if ωinfvF(ω)vV belongs to Lp(Ω). Moreover, the set SFP is “decomposable”, that is, if (A,f1,f2)Σ×SFP×SFP, then

χAf1+χΩAf2SFP.

Here, for CΣ, by χC we denote the characteristic function of the set CΣ.

For every DΣ,D, we define

|D|=supvDvV  and  σ(v*,D)=supvDv*,vVfor all v*V*.

Here, ,V denotes the duality brackets of the pair (V*,V). The function σ(,D):V*¯={+} is known as the “support function” of D.

Let Z,W be Hausdorff topological spaces. We say that a multifunction G:Z2W{} is “upper semicontinuous” (USC for short), respectively, “lower semicontinuous” (LSC for short), if for all UW open, the set

G+(U)={zZ:G(z)U},respectively,G-(U)={zZ:G(z)U}

is open in Z. If G() is both USC and LSC, then we say that G() is continuous. On a Hausdorff topological space (W,τ) (τ being the Hausdorff topology), we can define a new topology τseq whose closed sets are the sequentially τ-closed sets. Then topological properties with respect to this topology have the prefix “sequential”. Note that ττseq and the two are equal if τ is first countable (see [7, p. 9] and [21, p. 808]). We say that G:Z2W{} is “closed” if the graph GrGZ×W is closed.

For any Banach space V, on Pf(V) we can define a generalized metric, known as the “Hausdorff metric”, by setting

h(E,M)=max[supeEd(e,M),supmMd(m,E)].

Recall that (Pf(V),h) is a complete metric space (see [22, p. 6]). If Z is a Hausdorff topological space, a multifunction G:ZPf(V) is said to be “h-continuous”, if it is continuous from Z into (Pf(V),h).

Also, if E,MV are nonempty, bounded, closed and convex subsets, then (Hörmander’s formula)

h(E,M)=supv*V*,v*V*1|σ(v*,E)-σ(v*,M)|.

Let (W,τ) be a Hausdorff topological space with topology τ and let {En}n12W{}. We define

Kseq(τ)-lim infnEn={yW:y=τ-limnyn,ynEn for all n},
Kseq(τ)-lim supnEn={yW:y=τ-limnynk,ynkEnk,n1<n2<<nk<}.

Sometimes we drop the Kseq-symbol and simply write τ-lim supnEn and τ-lim infnEn.

Returning to the setting of an evolution triple, we consider a sequence of multivalued maps

an,a:Lp(T,X)2Lp(T,X*){}(n)

such that for every h*Lp(T,X*) the inclusions

y+an(y)h*(n)  and  y+a(y)h

have unique solutions en(h*),e(h*)Wp(0,b).

We say that ddt+an “PG-converges” to ddt+a (denoted by ddt+anPGddt+a as n) if for every h*Lp(T,X*), we have

en(h*)𝑤e(h*)in Wp(0,b).

In what follows, by Xw (respectively, Hw,Xw*) we denote the space X (respectively, H,X*) furnished with the weak topology. Also, by ||1 we denote the Lebesgue measure on and by ((,)) we denote the duality brackets for the pair (Lp(T,X*),Lp(T,X)). So, we have

((h*,f))=0bh*(t),f(t)𝑑tfor all h*Lp(T,X*) and all fLp(T,X).

Next, let us recall some useful facts from the theory of nonlinear operators of monotone type. So, let V be a reflexive Banach space, L:D(L)VV* a linear maximal monotone operator and a:V2V*. We say that a() is “L-pseudomonotone” if the following conditions hold:

  1. For every vV,a(v)Pwkc(V*).

  2. a() is bounded (that is, maps bounded sets to bounded sets).

  3. If {vn}n1D(L), vn𝑤vD(L) in V, L(vn)𝑤L(v) in V*, vn*a(vn) for all n, vn*𝑤v* in X* and lim supnvn*,vn-vV0, then v*a(v) and vn*,vnVv*,vV.

Such maps have nice surjectivity properties.

The next result is due to Papageorgiou, Papalini and Renzacci [35], and it extends an earlier single-valued result of Lions [27, Theorem 1.2, p. 319].

Proposition 2.3

Assume that V is a reflexive Banach space which is strictly convex, L:D(L)VV* is a linear maximal monotone operator and A:V2V* is L-pseudomonotone and strongly coercive, that is,

infv*A(v)v*,vVvV+as vV+.

Then R(L+V)=V* (that is, L+V is surjective).

In the next section we obtain some results about a general class of evolution inclusions, which will help us study problem (1.1) (see Section 4).

3 Nonlinear evolution inclusions

Let T=[0,b] and let (X,H,X*) be an evolution triple with XH compactly (see Definition 2.1). In this section we deal with the following evolution inclusion:

(3.1)-x(t)A(t,x(t))+E(t,x(t))for almost all tT,x(0)=ξ.

The hypotheses on the data of (3.1) are as follows.

  1. A:T×X2X* is a map such that the following hold:

    1. tA(t,x) is graph measurable for all xX.

    2. GrA(t,) is sequentially closed in Xw×Xw* and xA(t,x) is pseudomonotone for almost all tT.

    3. For almost all tT, all xX and all h*A(t,x), we have

      h*a1(t)+c1xp-1

      with 2p,a1Lp(T) and c1>0.

    4. For almost all tT, all xX and all h*A(t,x), we have

      h*,xc2xp-a2(t)

      with c2>0,a2L1(T).

Remark 3.1

If A(,) is single-valued, then in hypothesis (H$A$1) (ii) we can drop the condition on the graph of GrA(t,) and only assume that xA(t,x) is pseudomonotone for almost all tT. The same applies if A(t,) is maximal monotone for almost all tT. An example of where the condition on the graph of A(t,) is satisfied is the following (for simplicity we drop the t-dependence):

A(x)=-divφ(Dx)-divξ(Dx),

where φ:Lp(Ω,N) is continuous and convex, and ξ:Lp(Ω,N) is continuous and satisfies |ξ(y)|c^(1+|y|τ-1) for all yN, c^>0 and 1τ<p. Then recalling that W1,p(Ω)W1,τ(Ω) compactly (see Zeidler [41, p. 1026]), we easily see that GrA is sequentially closed in W1,p(Ω)w×W1,p(Ω)w*.

  1. F:T×HPfc(H) is a multifunction such that the following hold:

    1. tF(t,x) is graph measurable for all xH.

    2. GrF(t,) is sequentially closed in H×Hw for almost all tT.

    3. For almost all tT, all xH and all hF(t,x), we have

      |h|a3(t)+c3|x|

      with a3L2(T),c3>0 and if p=2, then β2c3<c2 (see (2.1)).

By a solution of problem (3.1) we understand a function xWp(0,b) such that

-x(t)=h*(t)+f(t)for almost all tT,

where h*Lp(T,X*) and fL2(T,H) are such that

h*(t)A(t,x(t))andf(t)F(t,x(t))for almost all tT.

By S(ξ) we denote the set of solutions of problem (3.1). In the sequel we investigate the structure of S(ξ).

Consider the multivalued map a:Lp(T,X)2Lp(T,X*) defined by

(3.2)a(x)={h*Lp(T,X*):h*(t)A(t,x(t)) for almost all tT}for all xLp(T,X).

Note that a() has values in Pwkc(Lp(T,X*)) (see hypotheses (H$A$1) (i) and (iii), and use the Yankov–von Neumann–Aumann selection theorem, see [22, Theorem 2.14, p. 158]).

Lemma 3.2

If hypotheses (H$A$1) hold, xn𝑤x in Wp(0,b), xn(t)𝑤x(t) in X for almost all tT, hn*𝑤h* in Lp(T,X*) and hn*a(xn) for all nN, then h*a(x).

Proof.

Let vX and consider the function xσ(v,A(t,x)) (see Section 2). We will show that it is sequentially upper semicontinuous. To this end, we need to show that given λ, the superlevel set

Eλ={xX:λσ(v,A(t,x))}

is sequentially closed in Xw. So, we consider a sequence {xn}n1Eλ and assume that

xn𝑤xin X.

Let hn*A(t,xn) (n) be such that

(3.3)hn*,v=σ(v,A(t,xn))for all n

(recall A(t,xn)Pwkc(X*)). Evidently, {hn*}n1X* is bounded (see hypothesis (H$A$1) (iii)) and so, by passing to a subsequence if necessary, we may assume that hn*𝑤h* in X*. Therefore, by hypothesis (H$A$1) (ii),

(3.4)h*A(t,x)

Then, from (3.3) and (3.4), we have

λh*,vσ(v,A(t,x)),

and thus xEλ. This proves the upper semicontinuity of the map xσ(v,A(t,x)).

Now let vLp(T,X). Then we have (see [22, Theorem 3.24, p. 183])

((hn*,v))σ(v,a(xn))=0bσ(v(t),A(t,xn(t)))𝑑tfor all n,

and by Fatou’s lemma,

((h*,v))lim supnσ(v,a(xn))0blim supnσ(v(t),A(t,xn(t)))dt.

By the first part of the proof and since, by hypothesis, xn(t)𝑤x(t) in X,

((h*,v))0bσ(v(t),A(t,x(t)))𝑑t=σ(v,a(x)).

Thus, h*a(x). ∎

Lemma 3.3

If hypotheses (H$A$1) hold, then the multivalued map a:Lp(T,X)2Lp(T,X*), defined by (3.2), is L-pseudomonotone.

Proof.

Suppose xn𝑤x in Wp(0,b),hn*a(xn) for all n, hn*𝑤h* in Lp(T,X*) and

(3.5)lim supn((hn*,xn-x))0.

From (2.2), we infer that

(3.6)xn(t)𝑤x(t)in H for all tT as n.

We set ϑn(t)=hn*(t),xn(t)-x(t). Let N be the Lebesgue-null set in T=[0,b] outside of which hypotheses (H$A$1) (ii), (iii) and (iv) hold. Using hypotheses (H$A$1) (iii) and (iv), we have

(3.7)ϑn(t)c2xn(t)2-a2(t)-(a1(t)+c1xn(t)p-1)x(t)for all tTN.

We introduce the Lebesgue measurable set DT defined by

D={tT:lim infnϑn(t)<0}.

Suppose that |D|1>0. If tD(TN), then from (3.7) we see that {xn(t)}n1X is bounded. Then, from (3.6), it follows that

(3.8)xn(t)𝑤x(t) in X for all tD(TN).

We fix tD(TN) and choose a subsequence {nk} of {n} (in general this subsequence depends on t) such that

limkϑnk(t)=lim infnϑn(t).

By hypothesis (H$A$1) (ii), A(t,) is pseudomonotone, and since tD, we infer that

limkhnk*(t),xnk(t)-x(t)=0,

a contradiction. So, |D|1=0 and we have

(3.9)0lim infnϑn(t)for almost all tT.

Invoking the extended Fatou’s lemma (see [15, Theorem 2.2.33]), we have

00blim infnϑn(t)dt(see (3.9))
lim infn0bϑn(t)𝑑t
lim supn0bϑn(t)𝑑t
=lim supn0bhn*(t),xn(t)-x(t)𝑑t
=lim supn((hn*,xn-x))0(see (3.5)).

Therefore,

(3.10)0bϑn(t)𝑑t0.

We write

(3.11)|ϑn(t)|=ϑn+(t)+ϑn-(t)=ϑn(t)+2ϑn-(t).

Note that, from (3.9),

(3.12)ϑn-(t)0for almost all tT.

Moreover, from (3.7) we have

ϑn(t)ηn(t)for almost all tT and all n

with {ηn}n1L1(T) uniformly integrable. Then

ϑn-(t)ηn-(t)for almost all tT and all n

with {ηn-}n1L1(T) uniformly integrable. Using (3.12) and invoking Vitali’s theorem we infer that ϑn-0 in L1(T). Hence, from (3.10) and (3.11),

(3.13)ϑn0in L1(T).

Then, from (3.13) and the fact that hn*𝑤h* in Lp(T,X*)), we have

|((hn*,xn))-((h*,x))||((hn*,xn-x))|+|((hn*-h*,x))|0,

which implies ((hn*,xn))((h*,x)). In addition, from (3.8) and Lemma 3.2, we have that h*a(x). This proves the L-pseudomonotonicity of a(). ∎

Remark 3.4

From the above proof it is clear why in the case of a single-valued map A(t,x), in hypothesis (H$A$1) (ii) we can drop the condition on the graph of A(t,) and only assume that xA(t,x) is pseudomonotone for almost all tT. Indeed, in this case, from (3.13) we have (at least for a subsequence) that

ϑn(t)0for almost all tT,

which, since A(t,) is pseudomonotone, implies

A(t,xn(t))𝑤A(t,x(t))for almost all tT in X*.

In the multivalued case, there is no canonical way to identify the pointwise limit of the sequence {hn*(t)}n1X*. If for almost all tT,A(t,) is maximal monotone, then again, we do not need the graph hypothesis on A(t,). In this case a() is also maximal monotone and then the lemma is a consequence of (3.5) and [4, Lemma 1.3]. It is worth mentioning that a similar strengthening of the topology in the range space was used by Defranceschi [12], while studying G-convergence of multivalued operators.

Without loss of generality, invoking the Troyanski renorming theorem (see [21, Remark 2.115]), we may assume that both X and X* are locally uniformly convex, hence Lp(T,X) and Lp(T,X*) are strictly convex.

We are now ready for the first result concerning the solution set S(ξ).

Theorem 3.5

If hypotheses (H$A$1), (H$F$1) hold and ξH, then the solution set S(ξ) is nonempty, weakly compact in Wp(0,b) and compact in C(T,H).

Proof.

First suppose that ξX. We define

A1(t,x)=A(t,x+ξ)andF1(t,x)=F(t,x+ξ).

Evidently, A1(t,x) and F1(t,x) have the same measurability, continuity and growth properties as the multivalued maps A(t,x) and F(t,x). So, we may equivalently consider the following Cauchy problem:

(3.14)-x(t)A1(t,x(t))+F1(t,x(t))for almost all tT,x(0)=0.

Note that if xWp(0,b) is a solution of (3.14), then x^=x-ξ is a solution of (3.1) (when ξX, that is, the initial condition is regular). Consider the linear densely defined operator L:D(L)Lp(T,X)Lp(T,X*) defined by

L(x)=xfor all xWp0(0,b)={yWp(0,b):y(0)=0}

(the evaluation y(0)=0 makes sense by virtue of (2.2)).

Consider the multivalued maps a1,G1:Lp(T,X)2Lp(T,X*){} defined by

a1(x)={h*Lp(T,X*):h*(t)A1(t,x(t)) for almost all tT},
G1(x)={fLp(T,X*):f(t)F1(t,x(t)) for almost all tT}.

We set K(x)=a1(x)+G1(x) for all xLp(T,X). Then

K:Lp(T,X)2Lp(T,X*){}.
Claim 1

K is L-pseudomonotone.

Clearly, K has values in Pwkc(Lp(T,X*)) and it is bounded (see hypotheses (H$A$1) (iii), (H$F$1) (iii)).

Next, we consider a sequence {xn}n1D(L) such that

(3.15){xn𝑤xD(L)in Lp(T,X),L(xn)L(x)in Lp(T,X*),kn*K(xn),kn*𝑤k*in Lp(T,X*)  and  lim supn((kn*,xn-x))0.

Then we have

kn*=hn*+fnwithhn*a1(xn),fnG1(xn)for all n.

Hypotheses (H$A$1) (iii) and (H$F$1) (iii) imply that

{hn*}n1Lp(T,X*)and{fn}n1Lp(T,H) are bounded.

So, we may assume (at least for a subsequence) that

hn*𝑤h*in Lp(T,X*)  and  fn𝑤fin Lp(T,H).

By (3.15) we have xn𝑤x in Wp(0,b), and due to (2.3),

(3.16)xnxin Lp(T,H).

Hence,

((fn,xn-x))=0b(fn(t),xn(t)-x(t))𝑑t0,

and in view of (3.15),

lim supn((hn*,xn-x))0.

Therefore, from Lemma 3.3,

h*a1(x)and((hn*,xn))((h*,x)).

Recall that

(3.17)fn(t)F(t,xn(t))for almost all tT and all n.

By (3.15), (3.16), (3.17) and [34, Proposition 6.6.33], we have

f(t)conv¯w-lim supnF(t,xn(t))F(t,x(t))for almost all tT

(see hypothesis (H$F$1) (ii)), and thus fG1(x).

Since ((fn,xn-x))=0b(fn(t),xn(t)-x(t))𝑑t0, we conclude that K is L-pseudomonotone. This proves Claim 1.

Claim 2

K is coercive.

Let xLp(T,X) and k*K(x). Then

k*=h*+fwithh*a1(x),fG1(x).

We have (see hypothesis (H$A$1) (iv))

(3.18)((k*,x))=((h*,x))+0b(f(t),x(t))dtc2xLp(T,X)p-a21-0b|f(t)||x(t)|dt.

Note that from hypothesis (H$F$1) (iii) and using Young’s inequality with ϵ>0, we have

(3.19)0b|f(t)||x(t)|dt0b(a3(t)|x(t)|+c3|x(t)|2)dt0b(c(ϵ)a3(t)2+(c3+ϵ)|x(t)|2)dt

Returning to (3.18) and using (3.19), we see that

(3.20)((k*,x))c2xLp(T,X)p-c4xLp(T,X)2-c5for c4,c5>0

(recall that 2p and in case p=2, choose ϵ>0 small so that c4<c2, see hypothesis (H$F$1) (iii)). From (3.20) it follows that K is coercive. This proves Claim 2.

Now Claims 1 and 2 permit the use of Proposition 2.3 to find xWp(0,b) solving problem (3.1) when ξX.

Next, we remove the restriction ξX. So, suppose ξH. We can find {ξn}n1X such that ξnξ in H (recall that X is dense in H). From the first part of the proof, we know that we can find xnS(ξn)Wp(0,b) for all n. We have

{-xn(t)A(t,xn(t))+F(t,xn(t))for almost all tT,xn(0)=ξn,n.

It follows that

(3.21)-xn=hn*+fnwithhn*(t)A(t,xn(t)),fn(t)F(t,xn(t))

for almost all tT and all n.

We have

((xn,xn))+((hn*,xn))0b|fn(t)||xn(t)|dt,

and thus

(3.22)12|xn(b)|2+c2xnLp(T,X)pc6+c7xnLp(T,X)2for c6,c7>0

(see hypotheses (H$A$1) (iv), (H$F$1) (iii) and if p=2 ,then, as before, we have c7<c2).

From (3.21), (3.22) and hypotheses (H$A$1) (iii), (H$F$1) (iii), it follows that {xn}n1Wp(0,b) is bounded. So, we may assume that as n (see (2.3)),

(3.23)xn𝑤xin Wp(0,b)  and  xnxin Lp(T,H).

By (3.21), for all n, we have

(3.24)((xn,xn-x))+((hn*,xn-x))=-((fn,xn-x))=-0b(fn(t),xn(t)-x(t))𝑑t.

By Proposition 2.2 we know that

((xn-x,xn-x))=12|xn(b)-x(b)|2-12|ξn-ξ|2.

Hence,

((xn,x-xn))12|ξn-ξ|2+((x,x-xn)).

By (3.23) and since ξnξ in H,

(3.25)lim supn((xn,x-xn))0.

Hypothesis (H$F$1) (iii) implies that {fn}n1L2(T,H) is bounded. Hence, from (3.23),

0b(fn(t),x(t)-xn(t))𝑑t0,

and so, by (3.25),

lim supn[((xn,x-xn))+((fn,x-xn))]0.

Therefore, from (3.24),

(3.26)lim supn((hn*,xn-x))0.

By hypothesis (H$A$1) (iii) we see that {hn*}n1Lp(T,X*) is bounded. So, we may assume that

(3.27)hn*𝑤h*in Lp(T,X*) as n.

From (3.23), (3.26) and (3.27), we see that we can use Lemma 3.3 and infer that

(3.28)h*(t)A(t,x(t))for almost all tT.

As we have already mentioned {fn}n1L2(T,H) is bounded and so we may assume that

(3.29)fn𝑤fin L2(T,H).

Using [34, Proposition 6.6.33], we have (see hypothesis (H$F$1) (ii))

(3.30)f(t)conv¯w-lim supnF(t,xn(t))F(t,x(t))for almost all tT.

In (3.21), we pass to the limit as n and use (3.23), (3.27) and (3.29) to obtain

-x=h*+fwithh*a(x) (see (3.28)),fG(x) (see (3.30)),x(0)=ξ.

Hence, xS(ξ). So, we have proved that when ξH, the solution set S(ξ) is a nonempty subset of Wp(0,b).

Next, we will prove the compactness of S(ξ) in Wp(0,b)w and in C(T,H). Let xS(ξ). For every tT we have

0tx(s),x(s)ds+0th*(s),x(s)ds0t|f(s)||x(s)|dswith h*a(x),

which implies

12|x(t)|212c82+c90t|x(s)|2dsfor c8,c9>0,

and hence, by Gronwall’s inequality,

(3.31)|x(t)|Mfor some M>0, all tT and all xS(ξ).

Then let rM:HH be the M-radial retraction defined by

rM(x)={xif |x|M,Mx|x|if |x|>M.

Because of the a priori bound (3.31), we can replace F(t,x) by

F^(t,x)=F(t,rM(x)).

Note that for all xH, tF^(t,x) is graph measurable (hence also measurable, see Section 2) and for almost all tT, xF^(t,x) has a graph which is sequentially closed in H×Hw. Moreover, we see that

|F^(t,x)|a4(t)for almost all tT and all xH with a4L2(T).

We introduce the set

C={fL2(T,H):|f(t)|a4(t) for almost all tT}.

We consider the following Cauchy problem:

(3.32){-x(t)A(t,x(t))+f(t)for almost all tT=[0,b],x(0)=ξH,fC.

Let H:C2C(T,H) be the map (in general, multivalued) that assigns to each fC the set of solutions of problem (3.32). It is a consequence of Proposition 2.3 and Lemma 3.3 that H() has nonempty values.

Claim 3

H(C)C(T,H) is compact.

Let {xn}n1H(C). Then

(3.33)-xn=hn*+fnwithhn*a(xn),fnCfor all n.

Hence, for all tT, we have

12|xn(t)|212c102+0ta4(s)|xn(s)|𝑑sfor some c10>0 and all n,

which implies (see [6, Lemme A.5])

(3.34)|xn(t)|M1for M1>0, all tT and all n.

Also, using hypothesis (H$A$1) (iv), we have (see (3.34))

(3.35)c2xnLp(T,X)pc11+0ba4(t)|xn(t)|𝑑tc12for all n.

From (3.33) and (3.35) it follows that {xn}n1Wp(0,b) is bounded. So, we may assume that

(3.36)xn𝑤xin Wp(0,b),hn*𝑤h*in Lp(T,X*),fn𝑤fin L2(T,H).

Passing to the limit as n in (3.33) and using (3.36), we obtain -x=h*+f. Also, from (3.33) we have

(3.37)((hn*,xn-x))+((xn,xn-x))=-0b(fn(t),xn(t)-x(t))𝑑t.

Note that, by (3.36) and (2.3),

(3.38)0b(fn(t),xn(t)-x(t))𝑑t0.

Also using Proposition 2.2, we have (recall that xn(0)=x(0)=ξ for all n)

((xn-x,xn-x))=12|xn(b)-x(b)|20for all n,

and so

(3.39)((x,xn-x))((xn,xn-x))for all n.

From (3.36), it follows that

(3.40)((x,xn-x))0.

Returning to (3.37), passing to the limit as n and using (3.38), (3.39) and (3.40), we obtain

lim supn((hn*,xn-x))0,

and thus h*a(x) (see Lemma 3.3 and (3.36)). Therefore, H(C) is w-compact in Wp(0,b). From the proof of Lemma 3.3 (see (3.13)), we know that

(3.41)0b|hn*(t),xn(t)-x(t)|dt0as n.

In a similar fashion, we also have

(3.42)0b|h*(t),xn(t)-x(t)|dt0as n.

Also, by (2.3), (3.34), (3.36) and Vitali’s theorem, we have

(3.43)0b|(fn(t)-f(t),xn(t)-x(t))|dt0as n.

For every tT and every n, using Proposition 2.2, we have

12|xn(t)-x(t)|20b|hn*(t)-h*(t),xn(t)-x(t)|dt0b|(fn(t)-f(t),xn(t)-x(t))|dt,

and so, from (3.41), (3.42) and (3.43), xn-xC(T,H)0. Thus, H(C)C(T,H) is compact.

However, from the previous parts of the proof it is clear that S(ξ)H(C) is weakly closed in Wp(0,b) and closed in C(T,H). Therefore, we conclude that S(ξ) is weakly compact in Wp(0,b) and compact in C(T,H). ∎

Next, we want to produce a continuous selection of the multifunction ξS(ξ) (we refer to [36] for more details about continuous selections of multivalued mappings). Note that S() is in general not convex-valued, and so the Michael selection theorem (see [22, Theorem 4.6, p. 92]) cannot be used. To produce a continuous selection of the solution multifunction ξS(ξ), we need to strengthen the conditions on the multimap A(t,) in order to guarantee that certain Cauchy problems admit a unique solution.

The new hypotheses on the map A(t,x) are as follows.

  1. A:T×X2X*{} is a multivalued map such that the following hold:

    1. tA(t,x) is graph measurable for every xX.

    2. xA(t,x) is maximal monotone for almost all tT.

    3. For almost all tT, all xX and all h*A(t,x), we have

      h**a1(t)+c1xp-1

      with a1Lp(T),c1>0,2p.

    4. For almost all tT, all xX and h*A(t,x), we have

      h*,xc2xp-a2(t)

      with c2>0,a2L1(T)+.

Remark 3.6

As we have already mentioned in an earlier remark, since now A(t,) is maximal monotone for almost all tT, we do not need the condition on the graph of A(t,) (see hypothesis (H$A$1) (ii) and [4, Lemma 1.3]).

Also, we strengthen the condition on the multifunction F(t,).

  1. F:T×HPfc(H) is a multifunction such that the following hold:

    1. tF(t,x) is graph measurable for every xH.

    2. For almost all tT and all x,yH, we have

      h(F(t,x),F(t,y))k(t)|x-y|with kL1(T)+.
    3. For almost all tT, all xH and all hF(t,x), we have

      |h|a3(t)+c3|x|

      with a3L2(T)+, c3>0, and if p=2, then β2c3<c2 (see (2.1)).

Remark 3.7

Hypothesis (H$F$2) (ii) is stronger than condition (H$F$1) (ii). Indeed, suppose that (H$F$2) (ii) holds and we have

(3.44)xnx,hn𝑤hin H  and  hnF(t,xn)for all n.

By the definition of the Hausdorff metric (see Section 2), we have

d(hn,F(t,x))d(hn,F(t,xn))+h(F(t,xn),F(t,x))=h(F(t,xn),F(t,x)),

and therefore (see (3.44) and hypothesis (H$F$2) (ii))

d(hn,F(t,x))0as n.

The function yd(y,F(t,x)) is continuous and convex, hence weakly lower semicontinuous. Therefore, by (3.44) we have

d(h,F(t,x))lim infnd(hn,F(t,x))=0,

which implies hF(t,x). This proves that condition (H$F$1) (ii) holds.

So, we can use Theorem 3.5 and establish that given any ξH, the solution set S(ξ) is nonempty, weakly compact in Wp(0,b) and compact in C(T,H). The next result extends an earlier result of Cellina and Ornelas [9] for differential inclusions in N with A0.

Proposition 3.8

If hypotheses (H$A$2),(H$F$2) hold, then there exists a continuous map ϑ:HC(T,H) such that

ϑ(ξ)S(ξ)for all ξH.

Proof.

Consider the following auxiliary Cauchy problem:

-x(t)A(t,x(t))for almost all tT=[0,b],x(0)=ξ.

This problem has a unique solution x0(ξ)Wp(0,b) (see Proposition 2.3 and use the monotonicity of A(t,) and Proposition 2.2 to check the uniqueness of this solution).

If ξ1,ξ2H, then

-x0(ξ1)=h1*and-x0(ξ2)=h2*with hk*a(x0(ξk)) for k=1,2.

So, using Proposition 2.2, we have

12|x0(ξ1)(t)-x0(ξ2)(t)|2+0th1*(s)-h2*(s),x0(ξ1)(s)-x0(ξ2)(s)𝑑s=12|ξ1-ξ2|2for all tT,

and hence (recall that A(t,) is monotone)

(3.45)x0(ξ1)-x0(ξ2)C(T,H)|ξ1-ξ2|.

We consider the multifunction Γ0:HPwkc(L1(T,H)) defined by

Γ0(ξ)=SF(,x0(ξ)())1for all ξH.

We have

h(Γ0(ξ),Γ0(ξ2))=sup[|σ(g,Γ0(ξ1))-σ(g,Γ0(ξ2))|:gL(T,H)=L1(T,H)*,gL(T,H)1]
0bsup|v|1[|σ(v,F(t,x0(ξ1)(t)))-σ(v,F(t,x0(ξ2)(z)))|]dt(see [34, Theorem 6.4.16])
=0bh(F(t,x0(ξ1)(t)),F(t,x0(ξ2)(t)))𝑑t
0bk(t)|x0(ξ1)(t)-x0(ξ2)(t)|𝑑t
0bk(t)|ξ1-ξ2|𝑑t(see (3.45))
=k1|ξ1-ξ2|.

Therefore, ξΓ0(ξ) is h-Lipschitz.

Also, Γ0() has decomposable values. So, we can apply the selection theorem of Bressan and Colombo [5] (see also [22, Theorem 8.7, p. 245]) and find a continuous map γ0:HL1(T,H) such that γ0(ξ)Γ0(ξ) for all ξH. Evidently, γ0(ξ)L2(T,H) for all ξH.

We consider the following auxiliary Cauchy problem:

-x(t)A(t,x(t))+γ0(ξ)(t)for almost all tT,x(0)=ξ.

This problem has a unique solution x1(ξ)Wp(0,b). By induction we will produce two sequences

{xn(ξ)}n1Wp(0,b)and{γn(ξ)}n1L2(T,H),

which satisfy the following:

  1. xn(ξ)Wp(0,b) is the unique solution of the Cauchy problem

    (3.46)-x(t)A(t,x(t))+γn-1(ξ)(t)for almost all tT,x(0)=ξ.
  2. ξγn(ξ) is continuous from H into C(T,H).

  3. γn(ξ)(t)F(t,xn(ξ)(t)) for almost all tT and all ξH.

  4. |γn(ξ)(t)-γn-1(ξ)(t)|k(t)βn(ξ)(t) for almost all tT and all ξH, where

    βn(ξ)(t)=20tη(ξ)(s)(τ(t)-τ(s))n-1(n-1)!𝑑s+2b(k=1nϵ2k+1)τ(t)n-1(n-1)!

    with ϵ>0, η(ξ)(t)=a2(t)+c2|x0(ξ)(t)| and τ(t)=0tk(s)𝑑s.

Note that the maps ξη(ξ) and ξβn(ξ) are continuous from H into L1(T). So, suppose we have produced {xk(ξ)}k=1n and {γk(ξ)}k=1n (induction hypothesis). Let xn+1(ξ)Wp(0,b) be the unique solution of the Cauchy problem

(3.47)-x(t)A(t,x(t))+γn(ξ)(t)for almost all tT,x(0)=ξ.

By (3.46) and (3.47) we have

(3.48)-xn(ξ)=hn*+γn-1(ξ)and-xn+1(ξ)=hn+1*+γn(ξ)in Lp(T,X*)

with

(3.49)hn*a(xn(ξ)),hn+1*a(xn+1(ξ)).

Using (3.48) we can write

xn+1(ξ)(t)-xn(ξ)(t),xn+1(ξ)(t)-xn(ξ)(t)
=hn*(ξ)(t)-hn+1*(ξ)(t),xn+1(ξ)(t)-xn(ξ)(t)+(γn-1(ξ)(t)-γn(ξ)(t),xn+1(ξ)(t)-xn(ξ)(t))
(γn-1(ξ)(t)-γn(ξ)(t),xn+1(ξ)(t)-xn(ξ)(t))for almost all tT

(see hypothesis (H$A$2) (ii) and (3.49)). Therefore, from Proposition 2.2,

12ddt|xn+1(ξ)(t)-xn(ξ)(t)|2|γn-1(ξ)(t)-γn(ξ)(t)||xn+1(ξ)(t)-xn(ξ)(t)|for almost all tT,

and thus

(3.50)12|xn+1(ξ)(t)-xn(ξ)(t)|20t|γn-1(ξ)(s)-γn(ξ)(s)||xn+1(ξ)(s)-xn(ξ)(s)|𝑑sfor all tT.

By (3.50) and [6, Lemma A.5], we infer that

|xn+1(ξ)(t)-xn(ξ)(t)|0t|γn-1(ξ)(s)-γn(ξ)(s)|ds
0tk(s)βn(ξ)(s)𝑑s(by the induction hypothesis, see (d))
=20tk(s)0sη(ξ)(r)(τ(s)-τ(r))n-1(n-1)!𝑑r𝑑s+2b(k=0nϵ2k+1)0tk(s)τ(s)n-1(n-1)!𝑑s
=20tη(ξ)(s)stk(r)(τ(r)-τ(s))n-1(n-1)!𝑑r𝑑s+2b(k=0nϵ2k+1)0tddsτ(s)nn!𝑑s
=20tη(ξ)(s)stddr(τ(r)-τ(s))nn!𝑑r𝑑s+2b(k=0nϵ2k+1)τ(t)nn!
=20tη(ξ)(s)(τ(t)-τ(s))nn!𝑑s+2b(k=0nϵ2k+1)τ(t)nn!
(3.51)<βn+1(ξ)(t)for almost all tT

(see (d)). Using the induction hypothesis (see (c)) and hypothesis (H$F$2) (ii), we have

d(γn(ξ)(t),F(t,xn+1(ξ)(t)))h(F(t,xn(ξ)(t)),F(t,xn+1(ξ)(t)))
k(t)|xn(ξ)(t)-xn+1(ξ)(t)|
(3.52)<k(t)βn+1(ξ)(t)for almost all tT

(see (3.51)).

Consider the multifunction Γn+1:H2L1(T,H) defined by

Γn+1(ξ)={fSF(,xn+1(ξ)())1:|γn(ξ)(t)-f(t)|<k(t)βn+1(ξ)(t) for almost all tT}.

By (3.52) and [22, Lemma 8.3, p 239], we have that ξΓn+1(ξ) has nonempty decomposable values and it is LSC. Thus, ξΓn+1(ξ)¯ is LSC with decomposable values.

We can apply the selection theorem of Bressan and Colombo [5, Theorem 3] to find a continuous map γn+1:HL1(T,H) such that γn+1(ξ)Γn+1(ξ)¯ for all ξH. This completes the induction process and we have produced two sequences {xn(ξ)}n1, {γn(ξ)}n1 which satisfy properties (a)–(d) stated earlier.

From (3.51) we have

0b|γn(ξ)(t)-γn-1(ξ)(t)|dt<0bβn+1(ξ)(t)dt
<0bη(ξ)(t)(τ(b)-τ(t))nn!𝑑t+2bϵτ(b)nn!
(3.53)τ(b)nn![η(ξ)1+2bϵ].

Recall that ξη(ξ) is continuous from H into L1(H) and maps bounded sets to bounded sets. So, from (3.53) it follows that {γn(ξ)}n1L1(T) is Cauchy, uniformly on bounded sets of H. Moreover, from (3.51) and (3.53), we have

xn+1(ξ)-xn(ξ)C(T,H)γn(ξ)-γn-1(ξ)L1(T,H)τ(b)nn![η(ξ)1+2bϵ].

Thus, {xn(ξ)}n1C(T,H) is Cauchy, uniformly on bounded sets. Therefore, we have

(3.54)xn(ξ)x(ξ)in C(T,H)  and  γn(ξ)γ(ξ)in L1(T,H).

Evidently, ξx(ξ) is continuous from H into C(T,H), while because of hypothesis (H$F$2) (iii), we have γn(ξ)γ(ξ) in L2(T,H). Let x^(ξ)Wp(0,b) be the unique solution of

-y(t)A(t,y(t))+γ(ξ)(t)for almost all tT,y(0)=0.

As before, exploiting the monotonicity of A(t,) (see hypothesis (H$A$2) (ii)), we have

12|xn+1(ξ)(t)-x^(ξ)(t)|20t|γn(ξ)(s)-γ(ξ)(s)||xn+1(ξ)(s)-x^(ξ)(s)|𝑑sfor all tT,

which implies (see [6, Lemma A.5])

xn+1(ξ)-x^(ξ)C(T,H)γn(ξ)-γ(ξ)L1(T,H).

Therefore, from (3.54), x(ξ)=x^(ξ).

So, x(ξ)S(ξ) and the map ϑ:HC(T,H) defined by ϑ(ξ)=x(ξ) is a continuous selection of the solution multifunction ξS(ξ). ∎

An easy but useful consequence of Proposition 3.8 and its proof is a parametric version of the Filippov–Gronwall inequality (see [1, Theorem 1, pp. 120–121] and [19]) for differential inclusions. So, we consider the following parametric version of problem (3.1):

-x(t)A(t,x(t))+F(t,x(t),λ)for almost all tT,x(0)=ξ(λ).

The parameter space D is a complete metric space. The hypotheses on the parametric vector field F(t,x,λ) and the initial condition ξ(λ) are as follows.

  1. F:T×H×DPfc(H) is a multifunction such that the following hold:

    1. tF(t,x,λ) is graph measurable for all (x,λ)H×D.

    2. For almost all tT, all x,yH and all λD, we have

      h(F(t,x,λ),F(t,y,λ))k(t)|x-y|

      with kL1(T)+.

    3. For almost all tT, all xH, all λD and all hF(t,x), we have

      |h|a3(t)+c3|x|

      with a3L2(T)+, c3>0, and if p=2, then β2c3<c2 (see (2.1)).

    4. For almost all tT and all xH, the multifunction λF(t,x,λ) is LSC.

  1. The mapping λξ(λ) is continuous from D into H.

Assume that λ(u(λ),h(λ)) is a continuous map from D into C(T,H)×L2(T,H). We can find a continuous map p:DL2(T) such that

d(h(λ)(t),F(t,u(λ)(t),λ))p(λ)(t)for almost all tT,

see hypothesis (H$F$2)’ (iii).

In what follows, by e(h,λ)Wp(0,b) we denote the unique solution of the Cauchy problem

-u(t)A(t,u(t))+h(t)for almost all tT,u(0)=ξ(λ)

with hL2(T,H).

We have the following approximation result.

Proposition 3.9

Assume that hypotheses (H$A$2), (H$F$2)’, (H0) hold, λ(u(λ),h(λ)) is a continuous map from D into C(T,H)×L2(T,H) with u(λ)=e(h(λ),λ), ϵ>0 and p:DL2(T)+ is a continuous map such that

d(h(λ)(t),F(t,u(λ)(t),λ))p(λ)(t)for almost all tT.

Then there exists a continuous map λ(x(λ),f(λ)) from D into C(T,H)×L2(T,H) such that

x(λ)=e(f(λ),λ)with f(λ)SF(,x(λ)(),λ)2

and

|x(λ)(t)-u(λ)(t)|bϵeτ(t)+0tp(λ)(s)eτ(t)-τ(s)𝑑sfor all tT

with τ(t)=0tk(s)𝑑s.

Proof.

Consider the multifunction Rϵ:D2L1(T,H) defined by

Rϵ(λ)={vSF(,u(λ)(),λ)1:|v(t)-h(λ)(t)|<p(λ)(t)+ϵ for almost all tT}.

This multifunction has nonempty, decomposable values and it is LSC (see [22, Lemma 8.3, p 239]). Hence, λRϵ(λ)¯ has the same properties. So, we can find a continuous map γ0:DL1(T,H) such that

γ0(λ)Rϵ(λ)¯for all λD.

Let x1(λ)Wp(0,b) be the unique solution of the following Cauchy problem:

-x(t)A(t,x(t))+γ0(λ)(t)for almost all tT,x(0)=ξ(λ).

Then as in the proof of Proposition 3.8, we can generate by induction two sequences

{xn(λ)}n1Wp(0,b)and{γn(λ)}n1L2(T,H),

which satisfy properties (a)–(d) listed in the proof of Proposition 3.8. As before (see the proof of Proposition 3.8), we have

|xn+1(λ)(t)-xn(λ)(t)|γn(λ)-γn-1(λ)L1(T,H)for all (λ,t)D×T.

From this inequality and property (d) of the sequences (see the proof of Proposition 3.8), we infer that

{xn(λ)}n1C(T,H)and{γn(λ)}n1L1(T,H)

are both Cauchy uniformly in λKD compact (recall that λp(λ) is continuous, hence locally bounded). So, we have

xn(λ)x^(λ)in C(T,H)  and  γn(λ)γ^(λ)in L1(T,H),

and both maps Dλx^(λ)C(T,H) and Dλγ^(λ)L1(T,H) are continuous. Moreover, we have γ^(λ)SF(,x^(λ)(),λ)2 (see the proof of Theorem 3.5) and that λγ^(λ) is continuous from D into L2(T,H). If x(λ)=e(γ^(λ),λ), then

|xn(λ)(t)-x(λ)(t)|0b|γn-1(λ)(s)-γ^(λ)(s)|ds0for all tT,

which implies

x^(λ)=x(λ)=e(γ^(λ),λ)for all λD.

From the triangle inequality, we have

|u(λ)(t)-xn(λ)(t)||u(λ)(t)-x1(λ)(t)|+k=1n-1|xk(λ)(t)-xk+1(λ)(t)|for all tT.

Using property (d) (see the proof of Proposition 3.8), we have

|xk(λ)(t)-xk+1(λ)(t)|1k!0tp(λ)(s)(τ(t)-τ(s))k𝑑s+bϵk!τ(t)kfor all tT.

So, finally we can write that

|u(λ)(t)-x(λ)(t)|bϵeτ(b)+0tp(λ)(s)eτ(t)-τ(s)𝑑sfor all tT and all λD.

We want to strengthen Proposition 3.8, and require that the selection ϑ() passes through a preassigned solution. We mention that an analogous result for differential inclusions in N with A0, was proved by Cellina and Staicu [10].

We start with a simple technical lemma.

Lemma 3.10

If {uk}k=0NL1(T,H) and {Tk(ξ)}k=0N is a partition of T=[0,b] with endpoints which depend continuously on ξH, then there exists d^L1(T)+ for which the following holds: “Given ϵ>0, we can find δ>0 such that for |ξ-ξ|δ,

|k=0NχTk(ξ)(t)uk(t)-k=0NχTk(ξ)(t)uk(t)|d^(t)χC(t)

with CT measurable and |C|1ϵ”.

Proof.

We have

|k=0NχTk(ξ)(t)uk(t)-k=0NχTk(ξ)(t)(t)uk(t)|k=0N|χTk(ξ)(t)-χTk(ξ)||uk(t)|
(3.55)=kχTk(ξ)ΔTk(ξ)(t)|uk(t)|.

We set d^(t)=k=0N|uk(t)|L1(T)+. From the hypothesis concerning the partition {Tk(ξ)}k=0N of T, we see that given ϵ>0, we can find δ>0 such that for |ξ-ξ|δ,

(3.56)χTk(ξ)ΔTk(ξ)(t)χC(t)for almost all tT and all k{0,,N}

with CT measurable, |C|1ϵ. Then, from (3.55) and (3.56),

|k=0NχTk(ξ)(t)uk(t)-k=0NχTk(ξ)(t)uk(t)|χC(t)k=0N|uk(t)|=d^(t)χC(t)for almost all tT.

The proof is now complete. ∎

With this lemma, we can produce a continuous selection of the solution multifunction ξS(ξ), which passes through a preassigned point.

Proposition 3.11

If hypotheses (H$A$2), (H$F$2) hold, KH is compact, ξ0K and vS(ξ0), then there exists a continuous map ψ:KC(T,H) such that

ψ(ξ)S(ξ)for all ξK  𝑎𝑛𝑑  ψ(ξ0)=v.

Proof.

Since vS(ξ0), we have

(3.57)-v(t)A(t,v(t))+f(t)for almost all tT,v(0)=ξ0

with fSF(,v())2. Given gL2(T,H), we consider the unique solution of the Cauchy problem

(3.58)-y(t)A(t,y(t))+g(t)for almost all tT,y(0)=ξH.

In what follows, by e(g,ξ)Wp(0,b) we denote the unique solution of problem (3.58) and we set μ0(ξ)=e(f,ξ). An easy application of the Yankov–von Neumann–Aumann selection theorem (see [22, Theorem 2.14, p. 158]) gives γ0(ξ)L2(T,H) such that

γ0(ξ)(t)F(t,μ0(ξ)(t))for almost all tT

and

|f(t)-γ0(ξ)(t)|=d(f(t),F(t,μ0(ξ)(t)))
k(t)|v(t)-μ0(ξ)(t)| (see hypothesis (HF2) (ii))
=k(t)|e(f,ξ0)(t)-e(f,ξ)(t)|
k(t)|ξ0-ξ|for almost all tT,

see (3.45).

Let ϑ>0. We define

δ(ξ)={min{2-3ϑ,|ξ-ξ0|2}if ξξ0,2-3ϑif ξ=ξ0.

The family {Bδ(ξ)(ξ)}ξK is an open cover of the compact set K. So, we can find {ξk}k=0NK such that {Bδ(ξk)(ξk)}k=0N is a finite subcover of K. Let {ηk}k=0N be a locally Lipschitz partition of unity subordinated to the finite subcover. We define

T0(ξ)=[0,η0(ξ)b]andTk(ξ)=[(i=0k-1ηi(ξ))b,(i=0kηi(ξ))(b)]for all k{1,,N}.

The endpoints in these intervals are continuous functions of ξ. We consider the following Cauchy problem:

(3.59)-y(t)A(t,y(t))+k=0NχTk(ξ)(t)γ0(ξk)(t)for almost all tT,y(0)=ξK.

Problem (3.59) has a unique solution μ1(ξ)Wp(0,b). Let

λ0(ξ)()=k=0NχTk(ξ)()γ0(ξk)()L2(T,H).

Using Lemma 3.10, we can find d^L1(T)+ such that, for any given ϵ>0, we can find δ>0 for which we have

(3.60)ξ,ξK,|ξ-ξ|1δ|λ0(ξ)(t)-λ0(ξ)(t)|d^(t)χC(t)for almost all tT

with CT measurable, |C|1ϵ. We have μ1(ξ)=e(λ0(ξ),ξ). As before, exploiting the monotonicity of A(t,) (see hypothesis (H$A$2) (ii)) and using [6, Lemma A.5], we have

(3.61)|μ1(ξ)(t)-μ1(ξ)(t)||ξ-ξ|+0t|λ0(ξ)(s)-λ0(ξ)(s)|𝑑sfor all tT.

Let ϵ>0 be given. By the absolute continuity of the Lebesgue integral, we can find δ1>0 such that

(3.62)Cd^(s)𝑑sϵ2for all CT measurable with |C|1δ1.

Also, using (3.60), we can find δ(0,ϵ/2) such that

(3.63)ξ,ξK,|ξ-ξ|1δ|λ0(ξ)(t)-λ0(ξ)(t)|d^(t)χC1(t)for almost all tT

with C1T measurable, |C1|1δ1. So, returning to (3.61) and using (3.62) and (3.63), we see that

ξ,ξK,|ξ-ξ|δ|μ1(ξ)(t)-μ1(ξ)(t)|ϵ2+0td^(s)χC1(s)𝑑sϵ2+ϵ2=ϵfor all tT.

Therefore, ξμ1(ξ) is continuous from H into C(T,H). Again, with an application of the Yankov–von Neumann–Aumann selection theorem, we obtain γ1(ξ)L2(T,H) such that

γ1(ξ)(t)F(t,μ1(ξ)(t))for almost all tT

and

|γ0(ξ)(t)-γ1(ξ)(t)|=d(γ0(ξ)(t),F(t,μ1(ξ)(t)))for almost all tT and all ξK.

As in the proof of Proposition 3.8, we produce inductively two sequences

{μn(ξ)}n0Wp(0,b)and{γn(ξ)}n0L2(T,H),ξK,

which satisfy the following properties:

  1. μn(ξ)=e(λn-1(ξ),ξ) with λn-1(ξ)=k=0NχTk(ξ)γn-1(ξk)(t),γ-1(ξ)=f,

  2. ξμn(ξ) is continuous from K into C(T,H),

  3. |μn(ξ)(t)-μn-1(ξ)(t)|ϑ2n+2n!(0tk(s)𝑑s)n for all ξK,

  4. γn(ξ)(t)F(t,μn(ξ)(t)) for almost all tT and

    |γn-1(ξ)(t)-γn(ξ)(t)|=d(γn-1(ξ)(t),F(t,μn(ξ)(t)))for almost all tT.

So, by the induction hypothesis, suppose that we have produced

{μk(ξ)}k=0nWp(0,b)and{γk(ξ)}k=0nL2(T,H),

which satisfy properties (a)–(d) stated above. We set

μn+1(ξ)=e(λn(ξ),ξ)withλn(ξ)(t)=k=0nχTk(ξ)(t)γn(ξk)(t).

As above (see in the first part of the proof the argument concerning the map ξμ1(ξ)), we can show that ξμn+1(ξ) is continuous from K into C(T,H). Also, by the monotonicity of A(t,) (see hypothesis (H$A$2) (ii) and [6, Lemma A.5]), we have (see hypothesis (H$F$2) (ii) and property (d) of the induction hypothesis)

|μn+1(ξ)(t)-μn(ξ)(t)|k=0n0tχTk(ξ)(s)k(s)|μn(ξ)(s)-μn-1(ξ)(s)|𝑑s
k=0n0tχTk(ξ)(s)k(s)ϑ2n+2n!(0sk(s)𝑑r)n𝑑s
=0tϑ2n+2(n+2)!dds(0sk(r)𝑑r)n+1𝑑s
=ϑ2n+2(n+1)!(0tk(s)𝑑s)n+1.

Note that for the second inequality we used property (c) of the induction hypothesis. Moreover, a standard measurable selection argument produces a measurable map γn+1(ξ):TH, ξK, such that

γn+1(ξ)(t)F(t,μn+1(ξ)(t))and|γn(ξ)(t)-γn+1(ξ)(t)|=d(γn(ξ)(t),F(t,μn+1(ξ)(t)))

for almost all tT. This completes the induction process.

Note that from property (c),

γn+1(ξ)-μn(ξ)C(T,H)ϑ2n+3ek1.

Therefore, we can say that

(3.64)μn(ξ)ψ(ξ)in C(T,H) as n, uniformly in ξK.

It follows that ξψ(ξ) is continuous from K into C(T,H).

Note that T0(ξ0)=T=[0,b] and so μ0(ξ0)=e(f,ξ0)=v (see (3.57)). Hence, ψ(ξ0)=v. It remains to show that ψ is a selection of the solution multifunction ξS(ξ). By property (d) and hypothesis (H$F$2) (ii), we have

|γn(ξ)(t)-γn+1(ξ)(t)|k(t)|μn(ξ)(t)-μn+1(ξ)(t)|for almost all tT,

and thus

(3.65)γn+1(ξ)γ^(ξ)in L2(T,H).

Let

μ^(ξ)=e(k=0N0tχTk(ξ)(s)γ^(ξk)(s)𝑑s,ξ).

Since, from (3.65),

k=0NχTk(ξ)γn(ξk)k=0NχTk(ξ)γ^(ξk)in L2(T,H),

we have μn(ξ)μ^(ξ) in C(T,H). Therefore, from (3.64),

ψ(ξ)=μ^(ξ)S(ξ)for all ξK.

4 Optimal control problems

In this section we deal with the sensitivity analysis of the optimal control problem (1.1).

Let Q(ξ,λ)Wp(0,b)×L2(T,Y) be the admissible “state-control” pairs. First we investigate the dependence of this set on the initial condition ξH and the parameter λE. Recall that the control space Y is a separable reflexive Banach space and the parameter space E is a compact metric space. To have a useful result on the dependence of Q(ξ,λ) on (ξ,λ)H×E, we introduce the following conditions on the data of the evolution inclusion in problem (1.1) (the dynamical constraint of the problem).

  1. A:T×X×E2X*{} is a multifunction such that

    1. tAλ(t,x) is graph measurable for every (x,λ)X×E.

    2. xAλ(t,x) is maximal monotone for almost all tT, all λE.

    3. For almost all tT, all xX, all λE and all h*Aλ(t,x), we have

      h**aλ(t)+cλxp-1

      with {aλ}λELp(T) bounded, {cλ}λE(0,+) bounded and 2p<.

    4. For almost all tT, all xX, all λE and all h*Aλ(t,x), we have

      h*,xc^xp-a^(t)

      with c^>0,a^L1(T)+.

    5. If λnλ in E, then ddt+aλnPGddt+aλ as n.

Hypotheses (H$A$3) (i)–(iv) are the same as hypotheses (H$A$2) (i)–(iv) for every map Aλ, λE. The new condition is hypothesis (H$A$3) (v), which requires elaboration. In the examples that follow, we present characteristic situations where this hypothesis is satisfied.

Example 4.1

(a) First, we present a situation which will be used in Section 5. So, let ΩN be a bounded domain with Lipschitz boundary Ω. Let X=W01,p(Ω) (2p<), H=L2(Ω) and X*=W-1,p(Ω). Evidently, (X,H,X*) is an evolution triple (see Definition 2.1) with compact embeddings. We consider a map a(t,z,ξ) satisfying the following conditions:

  1. a:T×Ω×NN is a map such that the following hold:

    1. |a(t,z,0)|c0 for almost all (t,z)T×Ω.

    2. (t,z)a(t,z,ξ) is measurable for every ξN.

    3. For almost all (t,z)T×Ω and all ξ1,ξ2N, we have

      |a(t,z,ξ1)-a(t,z,ξ2)|c^(1+|ξ1|+|ξ2|)p-1-α|ξ1-ξ2|α

      with c^1>0, α(0,1].

    4. For almost all (t,z)T×Ω and all ξ1,ξ2N, ξ1ξ2, we have

      (a(t,z,ξ1)-a(t,z,ξ2),ξ1-ξ2)Nc^2|ξ1-ξ2|p

      with c^2>0.

We consider the operator A:T×XX* defined by

A(t,x),h=Ω(a(t,z,Dx),Dh)N𝑑zfor all (t,x,h)T×X×X.

Using the nonlinear Green’s identity (see [20, p. 210]), we have

A(t,x)=-div(B(t,x)),

with B(t,x)()=a(t,,Dx())Lp(Ω,N) for all (t,x)T×X.

Now consider a sequence {an(t,z,ξ)}n1 of such maps satisfying

|an(t,z,ξ)-an(s,z,ξ)|ϑ(t-s)(1+|ξ|p-1)

for almost all zΩ, all t,sT, all ξN and all n, with ϑ:++ being an increasing function which is continuous at r=0 and ϑ(0)=0. We assume that for almost all tT, an(t,,)Ga(t,,) in the sense of Defranceschi [12]. By [40] we have

ddt+anPGddt+a.

(b) We can allow multivalued maps, provided that we drop the t-dependence. So, we consider multivalued maps a(z,ξ) which satisfy the following conditions:

  1. a:Ω×N2N{} is a measurable map such that the following hold:

    1. a(,) is measurable.

    2. ξa(z,ξ) is maximal monotone for almost all zΩ.

    3. For almost all zΩ, all ξN and all ya(z,ξ), we have

      |y|pm1(z)+c~1(y,ξ)with m1L1(Ω),c~1(y,ξ)>0,
      |ξ|pm2(z)+c~2(y,ξ)with m2L1(Ω),c~2(y,ξ)>0(2p<).

We again consider the evolution triple

X=W01,p(Ω),H=L2(Ω),X*=W-1,p(Ω)(2p<),

and consider the multivalued map A:X2X*{} defined by

A(x)={-divg:gSa(,Dx())p}.

We consider a sequence {an(z,ξ)}n1 of such maps and assume that anGa in the sense of Defranceschi [12]. Then by [16] we have

ddt+anPGddt+a.

(c) A third situation leading to hypothesis (H$A$3) (v) is the following one. We consider maps Aλ(t,x) satisfying the following conditions:

  1. A:T×X×EX* is a map such that the following hold:

    1. For all t,t+τT, all xX and all λE, we have

      Aλ(t+τ,x)-Aλ(t,x)O(τ)(1+xp-1).
    2. xAλ(t,x) is semicontinous for all (t,λ)T×E.

    3. For all tT, all x,uX and all λE, we have

      Aλ(t,x)-Aλ(t,u),x-uc~x-up

      with c~>0.

    4. If λnλ in E, then Aλn(t,)GAλ(t,) for all tT. (This means that Aλn-1(t,x*)𝑤Aλ-1(t,x) for all x*X*, see [14, Definition 3.8.20].)

Under these conditions, by [26], we have

ddt+aλnPGddt+aλ.

Next, we introduce the conditions on the multifunctions F and G involved in the dynamics of (1.1).

  1. F:T×H×EPfc(H) is a multifunction such that the following hold:

    1. tF(t,x,λ) is graph measurable for all (x,λ)H×E.

    2. for almost all tT, all x,yH and all λE, we have

      h(F(t,x,λ),F(t,y,λ))k(t)|x-y|

      with kL1(T)+.

    3. For almost all tT, all xH and all λE, we have

      |F(t,x,λ)|aλ(t)+cλ|x|

      with {aλ}λEL2(T) and {cλ}λE(0,+) bounded.

    4. For almost all tT, all xH and all λ,λE, we have

      h(F(t,x,λ),F(t,x,λ))β(d(λ,λ))w(t,|x|)

      with β(r)0+ as r0+ and w(t,) bounded on bounded sets.

  1. G:T×Y×EPfc(H) is a multifunction such that the following hold:

    1. tG(t,u,λ) is graph measurable for all (u,λ)Y×E.

    2. For almost all tT, all λE, uG(t,u,λ) is concave (that is, GrG(t,,λ)Y×H is concave, see [22, Definition 1.1 and Remark 1.2, p. 585]) and (u,λ)G(t,u,λ) is h-continuous.

    3. For almost all tT, all uU(t,λ) and all λE, we have

      |G(t,u,λ)|a^λ(t)

      with {a^λ}λEL2(T) bounded.

Remark 4.2

A typical situation resulting to a concave multifunction uG(t,u,λ) is when

G(t,u,λ)=Bλ(t)u+C(t,λ)for all (t,u,λ)T×Y×E

with Bλ(t)(Y,H) and C(t,λ)Pfc(H) for all (t,λ)T×E.

Another situation, leading to the concavity of G(t,,λ), is when H is an ordered Hilbert space and gλ,g~λ:T×YH are two Carathéodory maps such that for almost all tT

gλ(t,) is order convex and g~λ(t,) is order concave.

We set G(t,u,λ)={hH:gλ(t,u)hg~λ(t,u)}. Then G(t,,λ) is concave.

Finally, we impose conditions on the control constraint U(t,λ).

  1. U:T×EPfc(Y) is a multifunction such that the following hold:

    1. tU(t,λ) is graph measurable for all λE.

    2. λU(t,λ) is h-continuous for almost all tT.

    3. |U(t,λ)|a~λ(t) for almost all tT and all λE with {a~λ}λEL2(T) bounded.

Proposition 4.3

If hypotheses (H$A$3), (H$F$3), (H$G$), (H$U$) hold and (ξn,λn)(ξ,λ) in H×E, then

Kseq(s×w)-lim supnQ(ξn,λn)Q(ξ,λ)in Lp(T,H)×L2(T,Y),
K(s×s)-lim infnQ(ξn,λn)Q(ξ,λ)in C(T,H)×L2(T,Y).

Proof.

Let (x,u)Kseq(s×w)-lim supnQ(ξn,λn). By definition (see Section 2), we can find a subsequence {m} of {n} and (xm,um)Q(ξm,λm), m such that

(4.1)xmxin Lp(T,H)  and  um𝑤uin L2(T,Y)  as m.

For every m, we have

(4.2)-xm(t)Aλm(t,xm(t))+fm(t)+gm(t)for almost all tT,xm(0)=ξm,

with fm,gmL2(T,H) such that

(4.3)fm(t)F(t,xm(t),λm)andgm(t)G(t,um(t),λm)for almost all tT.

We deduce by hypotheses (H$F$3) (iii), (H$G$) (iii) and Theorem 3.5 and its proof that {xm}mWp(0,b) is bounded and {xm}mC(T,H) is relatively compact. So, from (4.1) we obtain

(4.4)xm𝑤xin Wp(0,b)  and  xmxin C(T,H)  as m.

By (4.3) and hypotheses (H$F$3) (iii), (H$G$) (iii), it is clear that {fm}m,{gm}mL2(T,H) are bounded. Hence, we may assume (at least for a subsequence), that

(4.5)fm𝑤f  and  gm𝑤gin L2(T,H)  as m.

Proposition 6.6.33 of [34], implies that

(4.6)f(t)conv¯w-lim supmF(t,xm(t),λm)for all tTN,|N|1=0.

Fix tTN and let yw-lim supmF(t,xm(t),λm). By definition, we know that there exists a subsequence {k} of {m}, and ykF(t,xk(t),λk) for all k such that yk𝑤y in H as k. The function vd(v,F(t,x(t),λ)) is continuous and convex, hence weakly lower semicontinuous. Therefore,

(4.7)d(y,F(t,x(t),λ))lim infkd(yk,F(t,x(t),λ)).

On the other hand, we have

(4.8)d(yk,F(t,x(t),λ))h(F(t,xk(t),λk),F(t,x(t),λ)).

Using hypotheses (H$F$3) (ii) and (iv), we have

h(F(t,xk(t),λk),F(t,x(t),λ))h(F(t,xk(t),λk),F(t,x(t),λk))+h(F(t,x(t),λk),F(t,x(t),λ))
k(t)|xk(t)-x(t)|+β(d(λk,λ))w(t,|x(t)|),

and so, from (4.4),

h(F(t,xk(t),λk),F(t,x(t),λ))0as k.

Then, from (4.7) and (4.8), we obtain d(y,F(t,x(t),λ))=0, hence yF(t,x(t),λ). Therefore,

w-lim supmF(t,xm(t),λm)F(t,x(t),λ)for all tTN,|N|1=0,

which implies (see (4.6) and recall that F is convex-valued)

f(t)F(t,x(t),λ)for all tTN,|N|1=0.

Next, for each m, we have

gmSG(,um(),λm)2.

Let hL2(T,H) and let (,)L2(T,H) denote the inner product of L2(T,H) (recall that L2(T,H)*=L2(T,H)). Then (see [34, Theorem 6.4.16])

(4.9)(h,gm)L2(T,H)σ(h,SG(,um(),λm)2)=0bσ(h(t),G(t,um(t),λm))𝑑t.

The concavity of G(t,,λ) (see hypothesis (H$G$) (ii)), implies that the function uσ(h(t),G(t,u,λ)) is concave. Since E is a complete metric space, it can be isometrically embedded, by the Arens–Eells theorem (see [21, Theorem 4.143]), as a closed subset of a separable Banach space (recall that E is compact). So, by [3], we have (see hypothesis (H$G$) (ii))

lim supm0bσ(h(t),F(t,um(t),λm))𝑑t0bσ(h(t),F(t,u(t),λ))𝑑t,

and thus

lim supmσ(h,SG(,um(),λm)2)σ(h,SG(,u(),λ)2).

Therefore, from (4.5) and (4.9),

(h,g)L2(T,H)σ(h,SG(,u(),λ)2).

Since hL2(T,H) is arbitrary, it follows that gSG(,u(),λ)2, hence

g(t)G(t,u(t),λ)for almost all tT.

Let ymWp(0,b) be the unique solution of the Cauchy problem

(4.10)-ym(t)Aλm(t,ym(t))+f(t)+g(t)for almost all tT,ym(0)=ξ.

Hypothesis (H$A$3) (v) implies that

(4.11)ym𝑤yin Wp(0,b)

with yWp(0,b) being the unique solution of the Cauchy problem

(4.12)-y(t)Aλ(t,y(t))+f(t)+g(t)for almost all tT,y(0)=ξ,

see Section 2. From (4.2), (4.10) and the monotonicity of Aλm(t,) (see hypothesis (H$A$3) (ii)), we have

xm(t)-ym(t),xm(t)-ym(t)(f(t)+g(t)-fm(t)-gm(t),xm(t)-ym(t))for almost all tT.

Therefore, by Proposition 2.2,

12|xm(t)-ym(t)|212|ξm-ξ|2+0t(f(s)+g(s)-fm(s)-gm(s),xm(s)-ym(s))𝑑sfor all tT,

which yields xm-ymC(T,H)0 as m, and hence, by (4.4) and (4.11), x=y.

Recalling that

f(t)F(t,x(t),λ)andg(t)G(t,u(t),λ)for almost all tT,

it follows from (4.12) that (x,u)Q(ξ,λ), which implies

Kseq(s×w)-lim supnQ(ξn,λn)Q(ξ,λ)in Lp(T,H)×L2(T,Y).

Next, we will prove the second convergence of the proposition. So, let (x,u)Q(ξ,λ). By definition we have

-x(t)Aλ(t,x(t))+F(t,x(t),λ)+g(t)for almost all tT,x(0)=ξ

with gL2(T,H) satisfying

g(t)G(t,u(t),λ)for almost all tT.

For every vL2(T,Y), we have (see [34, Theorem 6.4.16])

d(v,SU(,λn)2)=0bd(v(t),U(t,λn))𝑑t

Hypothesis (H$U$) (ii) and the dominated convergence theorem imply that

0bd(v(t),U(t,λn))𝑑t0bd(v(t),U(t,λ))𝑑t,

and so

d(v,SU(,λn)2)d(v,SU(,λ)2).

Hence, [34, Proposition 6.6.22] implies that we can find unSU(,λn)2 (n) such that

unuin L2(T,Y) as n.

Then hypothesis (H$G$) (ii) guarantees that we can find

gnL2(T,H),gn(t)G(t,un(t),λn)for almost all tT and all n

such that

gngin L2(T,H) as n.

Given ξH, let S(ξ)Wp(0,b) be the set of solutions of the Cauchy problem

-y(t)Aλn(t,y(t))+F(t,y(t),λ)+g(t)for almost all tT,y(0)=ξ.

Let K={ξn,ξ}n1H. This is a compact set in H. Invoking Proposition 3.11 (with ξ0=ξ), we produce a continuous map ψ:KC(T,H) such that

(4.13)y^=ψ(ξ^)S(ξ^)for all ξ^H,ψ(ξ)=x.

Let yn=ψ(ξn) (n) and use Proposition 3.9 to find xnWp(0,b) solution of the Cauchy problem

-xn(t)Aλn(t,xn(t))+F(t,xn(t),λn)+gn(t)for almost all tT,xn(0)=ϵn,

for which, we have

(4.14)|xn(t)-yn(t)|bϵeτ(t)+0tηn(s)eτ(t)-τ(s)𝑑sfor all tT

with ϵ>0, τ(t)=0tk(s)𝑑s, ηnL1(T), ηn0 in L1(T). So, we obtain (see (4.14))

lim supnxn-ynC(T,H)bϵeτ(b).

Since ϵ>0 is arbitrary, it follows that

(4.15)xn-ynC(T,H)0as n.

Finally, from (4.13) we have

xn-xC(T,H)xn-ynC(T,H)+yn-xC(T,H)=xn-ynC(T,H)+ψ(ξn)-ψ(ξ)C(T,H).

Therefore, from (4.15) and the fact that ψ() is continuous, xn-xC(T,H)0. Since (xn,un)Q(ξn,λn) (n) and unu in L2(T,Y), we conclude that

Q(ξ,λ)K(s×s)-lim infnQ(ξn,λn) in C(T,H)×L2(T,Y).

An immediate consequence of the above proposition is the following corollary concerning the multifunction (ξ,λ)Q(ξ,λ) of admissible state-control pairs.

Corollary 4.4

If hypotheses (H$A$3), (H$F$3), (H$G$), (H$U$) hold, then the multifunction

Q:H×E2C(T,H)×L2(T,Y){}

is LSC and sequentially closed in C(T,H)×L2(T,Y)w (that is, GrQH×E×C(T,H)×L2(T,Y)w is sequentially closed).

Now we bring the cost functional into the picture. The hypotheses on the integrands L(t,x,λ) and H(t,u,λ) are as follows.

  1. L:T×H×E is an integrand such that the following hold:

    1. tL(t,x,λ) is measurable for every (x,λ)H×E.

    2. If λnλ in E, then for all xH, we have L(,x,λn)𝑤L(,x,λ) in L1(T).

    3. For almost all tT, all x,yH and all λE, we have

      |L(t,x,λ)-L(t,y,λ)|=(1+|x||y|)ρ(t,|x-y|),

      where |x||y|=max{|x|,|y|} and ρ(t,r) is a Carathéodory function on T×+ with values in (0,+) such that

      ρ(t,0)=0for almost all tT
      andsup0rϑ[ρ(t,r)]βϑ(t)for almost all tT

      with βϑL1(T)+,ϑ>0.

  1. H:T×Y×E is an integrand such that the following hold:

    1. tH(t,u,λ) is measurable for all (u,λ)Y×E.

    2. uH(t,u,λ) is convex for almost all tT and all vE, and λH(t,u,λ) is continuous for almost all tT and all uY.

    3. For almost all tT and all (u,λ)Y×E, we have

      H(t,u,λ)a(t)(1+uY2)with aL(T).

  1. ψ^:H×E is a continuous function.

Using the direct method of the calculus of variations, we can produce optimal admissible state-control pairs for problem (1.1).

Proposition 4.5

If hypotheses (H$A$3), (H$F$3), (H$G$), (H$U$), (H$L$), (H$H$) and (H${\hat{\psi}}$) hold, then for every (ξ,λ)H×E we can find (x*,u*)Q(ξ,λ) such that J(x*,u*,ξ,λ)=m(ξ,λ).

Proof.

Let {(xn,un)}n1Q(ξ,λ) be a minimizing sequence for problem (1.1). So, we have

J(xn,un,ξ,λ)m(ξ,λ)as n.

Theorem 3.5 and hypothesis (H$U$) imply that

{(xn,un)}n1Wp(0,b)×L2(T,Y)(respectively, C(T,H)×L2(T,Y))

is relatively w×w-compact (respectively, s×w-compact). So, by the Eberlein–Smulian theorem and by passing to a suitable subsequence if necessary, we can say that

(4.16)xn𝑤x*in Wp(0,b),xnx*in C(T,H),un𝑤u*in L2(T,Y).

Then (4.16) and Proposition 4.3 imply that

(4.17)(x*,u*)Q(ξ,λ).

Also, (4.16), hypothesis (H$L$) (iii) and the dominated convergence theorem, imply that

(4.18)0bL(t,xn(t),λ)𝑑t0bL(t,x*(t),λ)𝑑t.

In addition, as before (see the proof of Proposition 4.3), using [3, Theorem 2.1], we obtain

(4.19)0bH(t,u*(t),λ)𝑑tlim infn0bH(t,un(t),λ)𝑑t.

Finally, (4.16) and hypothesis H(ψ^) imply that

(4.20)ψ^(ξ,xn(b),λ)ψ^(ξ,x*(b),λ).

We deduce from (4.17), (4.18), (4.19) and (4.20) that

J(x*,u*,ξ,λ)=m(ξ,λ)with (x*,u*)Q(ξ,λ).

This concludes the proof. ∎

We are now ready for the main sensitivity results concerning problem (1.1). The first one establishes the Hadamard well-posedness of the problem.

Theorem 4.6

If hypotheses (H$A$3), (H$F$3), (H$G$), (H$L$), (H$H$) and (H${\hat{\psi}}$) hold, then the value function m:H×ER of problem (1.1) is continuous.

Proof.

Let (ξn,λn)(ξ,λ) in H×E. Let (x,u)Q(ξ,λ) such that (see Proposition 4.5)

J(x,u,ξ,λ)=m(ξ,λ).

Invoking Proposition 4.3, we can find (xn,un)Q(ξn,λn) for all n such that

(4.21)xnxin C(T,H)  and  unuin L2(T,Y).

We claim that

(4.22)|0bL(t,xn(t),λn)dt-0bL(t,x(t),λ)dt|0as n.

To this end, note that

|0bL(t,xn(t),λn)dt-0bL(t,x(t),λ)dt|
(4.23)|0bL(t,xn(t),λn)dt-0bL(t,x(t),λn)dt|+|0bL(t,x(t),λn)dt-0bL(t,x(t),λ)dt|

for all n.

First, we estimate the first summand in the right-hand side of (4.23). Using hypothesis (H$L$) (iii), we have

|0bL(t,xn(t),λn)-L(t,x(t),λn)dt|b0(1+|xn(t)||x(t)|)ρ(t,|xn(t)-x(t)|)dt.

Let M=supn1xnC(T,H)<+ (see (4.21)). Then, from (4.21) and hypothesis (H$L$) (iii), we have

(4.24)|0bL(t,xn(t),λn)dt-0bL(t,x(t),λn)dt|(1+M)b0ρ(t,|xn(t)-x(t)|)dt0as n.

Next, we estimate the second term on the right-hand side of (4.23). Let ϑ>2xC(T,H) and let βϑL1(T)+ as postulated by hypothesis (H$L$) (iii). Given ϵ>0, we can find δ>0 such that

(4.25)“if CT is measurable with |C|1δ, then Cβϑ(t)𝑑tϵ2(1+ϑ).”

Here, we use the absolute continuity of the Lebesgue integral. Invoking the Scorza–Dragoni theorem (see [34, Theorem 6.2.9]), we can find T1T closed with |TT1|δ2 and such that ρ|T1×+ is continuous. Since ρ(t,0)=0, we can find δ1>0 such that

(4.26)“if r[0,δ1], then |ρ(t,r)|ϵ2b(1+ϑ) for all tT1.”

Recall that simple functions are dense in Lp(T,H). Using this fact, the property that Lp(T,H)-convergence implies pointwise convergence for almost all tT for at least a subsequence, and invoking Egorov’s theorem, we can find T2T closed and a simple function s:TH such that

(4.27)sxC(T,H),|TT2|1δ2  and  |x(t)-s(t)|δ1for all tT2.

We set T3=T1T2. This is a closed subset of T with |TT3|1δ. We have (see hypothesis (H$L$) (iii) and (4.27))

|0bL(t,x(t),λn)dt-0bL(t,s(t),λn)dt|(1+xC(T,H))b0ρ(t,|x(t)-s(t)|)dt
(1+ϑ)[T3ρ(t,|x(t)-s(t)|)𝑑t+TT3ρ(t,|x(t)-s(t)|)𝑑t]
(4.28)ϵ2+ϵ2=ϵ,

see (4.25), (4.26) and (4.27).

Similarly, we show that

(4.29)|0bL(t,s(t),λ)dt-0bL(t,x(t),λ)dt|ϵ.

Let s(t)=k=1NvkχCk(t) with vkH,CkT measurable. Using hypothesis (H$L$) (ii), we can find n0 such that

(4.30)|0b(L(t,s(t),λn)-L(t,s(t),λ))dt|k=1N|Ck(L(t,vk,λn)-L(t,vk,λ))dt|ϵfor all nn0.

From (4.28), (4.29) and (4.30) it follows that

0bL(t,x(t),λn)𝑑t0bL(t,x(t),λ)𝑑tas n.

This convergence and (4.24) imply that (4.22) (our claim) is true.

Next, we consider the integral functional

Φ(u,λ)=0bH(t,u(t),λ)𝑑tfor all (u,λ)L2(T,Y)×E.

For every λE,uΦ(u,λ) is convex (see hypothesis (H$H$) (ii)). Also, hypothesis (H$H$) (iii) implies that in a neighborhood of every uL2(T,Y), {Φ(,λ)}λE is equibounded above, hence {Φ(,λ)}λE is equi-locally Lipschitz (see [34, Theorem 1.2.3]). Therefore, it follows that (see (4.21))

(4.31)Φ(un,λn)Φ(u,λ)as n.

Finally, (4.21) and hypothesis (H${\hat{\psi}}$) imply that

(4.32)ψ^(ξn,xn(b),λn)ψ^(ξ,x(b),λ).

By (4.22), (4.31), (4.32), we have

0bL(t,xn(t),λn)𝑑t+0bH(t,un(t),λn)𝑑t+ψ^(ξn,xn(b),λn)J(x,u,ξ,λ)=m(ξ,λ),

which implies

(4.33)lim supnm(ξn,λn)m(ξ,λ).

From Proposition 4.5 we know that for every n, we can find (xn,un)Q(ξn,λn) such that

(4.34)J(xn,un,ξn,λn)=m(ξn,λn).

As in the proof of Theorem 3.5, we can show that {xn}n1Wp(0,b) is bounded. In addition, hypothesis (H$U$) implies that {un}n1L2(T,Y) is bounded. So, by passing to a suitable subsequence if necessary, we may assume that

(4.35)xn𝑤xin Wp(0,b)  and  un𝑤uin L2(T,Y)  as n.

By (4.35) and (2.3), we also have

(4.36)xnxin Lp(T,H) as n.

Then (4.35), (4.36) and Proposition 4.3 imply that (x,u)Q(ξ,λ). Moreover, reasoning as in the proof of Theorem 3.5, we can show that {xn}n1C(T,H) is relatively compact, hence (see (4.36))

(4.37)xnxin C(T,H).

By (4.37) and the first part of the proof, we have

0bL(t,xn(t),λn)𝑑t0bL(t,x(t),λ)𝑑t.

In addition, (4.35) and hypotheses (H$H$) (ii), (H${\hat{\psi}}$) imply (see[3])

0bH(t,u(t),λ)𝑑tlim infn0bH(t,un(t),λn)𝑑t,

and thus

ψ^(ξn,xn(b),λn)ψ^(ξ,x(b),λ).

Therefore, from (4.34) we see that

0bL(t,x(t),λ)𝑑t+0bH(t,u(t),λ)𝑑t+ψ^(ξ,x(b),λ)lim infnm(ξn,λn),

and thus

(4.38)m(ξ,λ)lim infnm(ξn,λn).

We infer from (4.33) and (4.38) that m(ξn,λn)m(ξ,λ), hence m:H×E is continuous. ∎

For every (ξ,λ)H×E, we introduce the set Σ(ξ,λ) of optimal state-control pairs, that is,

Σ(ξ,λ)={(x,u)Q(ξ,λ):J(x,u,ξ,λ)=m(ξ,λ)}.

By Proposition 4.5, we know that Σ(ξ,λ) for every (ξ,λ)H×E. For this multifunction we can prove the following useful continuity property.

Theorem 4.7

If hypotheses (H$A$3), (H$F$3), (H$U$), (H$L$), (H$H$) and (H${\hat{\psi}}$) hold, then the multifunction

Σ:H×E2C(T,H)×L2(T,Y){}

is sequentially USC into C(T,H)×L2(T,Y)w.

Proof.

Let CC(T,H)×L2(T,Y)w be sequentially closed. We need to show that

Σ-(C)={(ξ,λ)H×E:V(ξ,λ)C}

is closed in H×E (see Section 2). To this end, let {(ξn,λn)}n1Σ-(C), and assume that

(ξn,λn)(ξ,λ)in H×E.

Let (xn,un)Σ(ξn,λn)C, n. We know from the proof of Theorem 4.6 that at least for a subsequence, we have

(4.39)xn𝑤xin Wp(0,b),xnxin C(T,H),un𝑤uin L2(T,Y)  as n.

By (4.39) and Proposition 4.3, we have

(4.40)(x,u)Q(ξ,λ).

Also, we know from the proof of Theorem 4.6 that

J(x,u,ξ,λ)lim infnJ(xn,un,ξn,λn)=lim infnm(ξn,λ)=m(ξ,λ).

Therefore, from (4.40), J(x,u,ξ,λ)=m(ξ,λ), and thus (x,u)Σ(ξ,λ). Moreover, from (4.39) and since CC(T,H)×L2(T,Y)w is sequentially closed, we deduce that (x,u)Σ(ξ,λ)C. Therefore, Σ-(C)H×E is closed and this proves the desired sequential upper semicontinuity of the multifunction (ξ,λ)Σ(ξ,λ). ∎

5 Application to distributed parameter systems

In this section we present an application to a class of multivalued parabolic optimal control problems.

So, let T=[0,b] and let ΩN be a bounded domain with a Lipschitz boundary Ω. We examine the following nonlinear, multivalued parabolic optimal control problem:

(5.1){J(x,u,ξ,λ)=0bΩL1(t,z,x(t,z))𝑑z𝑑t+0bΩH1(t,z,u(t,z))𝑑z𝑑tinf=m(ξ,λ),-xt-divaλ(z,Du)+F1(t,z,x(t,z),λ)+g(t,z,λ)u(t,z)on T×Ω,x|T×Ω=0,x(0,z)=ξ(z)for almost all zΩ,u(t,)L2(Ω)r(t,λ)for almost all tT.

Here, aλ:Ω×N2N (λE) is a family of multifunctions as in Example 4.1 (b). For the other data of problem (5.1), we introduce the following conditions:

  1. F1:T×Ω××EPfc() is a multifunction such that the following hold:

    1. (t,z)F1(t,z,x,λ) is measurable for all (x,λ)×E.

    2. For almost all (t,z)T×Ω, all x,y and all λE, we have

      h(F1(t,z,x,λ),F1(t,z,y,λ))k1(t,z)|x-y|

      with k1L1(T,L(Ω)).

    3. For almost all (t,z)T×Ω, all x and all λE, we have

      |F1(t,z,x,λ)|a^1(t,z)+c^|x|

      with a^1L2(T×Ω),c^1>0.

    4. For almost all (t,z)T×Ω, all x and all λ,λE, we have

      h(F1(t,z,x,λ),F1(t,z,x,λ))β(d(λ,λ))w(z,|x|)

      with β(r)0 as r0+ and wLloc(Ω×+).

Remark 5.1

Consider the multifunction F(t,z,x,λ) defined by

F(t,z,x,λ)=[f(t,z,x,λ),f^(t,z,x,λ)],

where f,f^:T×Ω××E are two functions with the following properties:

  1. (t,z)f(t,z,x,λ),f^(t,z,x,λ) are both measurable for all (x,λ)×E.

  2. For almost all (t,z)T×Ω, all x,x and all λ,λE, we have

    |f(t,z,x,λ)-f(t,z,x,λ)|k(t,z)[|x-x|+d(λ,λ)]
    |f^(t,z,x,λ)-f^(t,z,x,λ)|k^(t,z)[|x-x|+d(λ,λ)],

    with k,k^L1(T,L(Ω)).

Then this multifunction satisfies hypotheses (H${F_{1}}$).

  1. g:T×Ω×E is a Carathéodory function (that is, (t,z)g(t,z,λ) is measurable for all λE and λg(t,z,λ) is continuous for almost all (t,z)T×Ω) and for almost all (t,z)T×Ω and all λE, we have |g(t,z,λ)|M with M>0.

  1. r:T×E+ is a Carathéodory function (that is, tr(t,λ) is measurable for all λE and λr(t,λ) is continuous for almost all tT) and for almost all tT and all λE, we have

    0r(t,λ)a(t)with aL2(T).

Now, we introduce the conditions on the two integrands involved in the cost functional problem (5.1).

  1. L:T×Ω××E is an integrand such that the following hold:

    1. (t,z)L1(t,z,x,λ) is measurable for all (x,λ)×E.

    2. If λnλ in E, then for all xL2(Ω) we have L1(,,x(),λn)𝑤L1(,,x(),λ) in L1(T×Ω).

    3. For almost all (t,z)T×Ω, all x,y and all λE,

      |L1(t,z,x,λ)-L1(t,z,y,λ)|c(1+|x||y|)ρ(t,z,|x-y|),

      with ρ(t,z,r) Carathéodory, ρ(t,z,0)=0 for almost all (t,z)T×Ω and for almost all (t,z), all r[0,ϑ] we have

      0ρ(t,z,r)βϑ(t,z)

      with βϑL1(T×Ω).

  1. H1:T×Ω××E is an integrand such that the following hold:

    1. (t,z)H1(t,z,x,λ) is measurable for all (x,λ)×E.

    2. For almost all (t,z)T×Ω, uH1(t,z,u,λ) is convex for all λE, while λH1(t,z,u,λ) is continuous for all u.

    3. For almost all (t,z)T×Ω, all |u|rλ(t,z) and all λE, we have

      |H1(t,z,u,λ)|a^λ(t,z)

      with {a^λ}λEL2(T×Ω) bounded.

We consider the following evolution triple:

X=W01,p(Ω),H=L2(Ω),X*=W-1,p(Ω).

Since 2p<, the Sobolev embedding theorem implies that in this triple the embeddings are compact.

For every λE, let Aλ:X2X*{} be the multivalued map defined by

Aλ(x)={-divg:gLp(Ω,N),g(z)aλ(z,Dx(z)) for almost all zΩ}.

This map is maximal monotone and if λnλ in E, then (see Example 4.1 (b))

ddt+aλnPGddt+aλ.

So, hypotheses (H$A$3) hold. In fact, we can have t-dependence at the expense of assuming that aλ is single-valued. So, we assume that aλ(t,z,ξ) satisfies the conditions of Example 4.1 (a). Then the map Aλ:T×XX* is defined by

Aλ(t,x)()=-divaλ(t,,Dx()).

In fact, by the nonlinear Green’s identity (see [20, p. 210]), we have

Aλ(t,x),h=Ω(aλ(t,z,Dx),Dh)N𝑑zfor all x,hW01,p(Ω).

As we have already mentioned in Example 4.1 (a), we know from [40] that if λnλ in E, then

ddt+aλnPGddt+aλ,

and so hypotheses (H$A$3) hold.

As a special case of interest, we consider the situation where the elliptic differential operator is a weighted p-Laplacian, that is,

div(aλ(t,z)|Dx|p-2Dx) for all xW01,p(Ω).

Here, for every λE,aλ:T×Ω is a measurable function with the following properties:

  1. 0<c^1aλ(t,z)c^2 for almost all (t,z)T×Ω and all λE.

  2. If λnλ in E, then for almost all tT,

    1aλn(t,)p-1𝑤1aλ(t,)p-1inL1(Ω).

For this case we consider the following parametric (with parameter λE) family of convex (in ξN) integrands:

φλ(t,z,ξ)=aλ(t,z)p|ξ|p.

Then the convex conjugate of φλ(t,z,) is given by

φλ*(t,z,ξ*)=1paλ(t,z)p-1|ξ*|p.

By hypothesis we have that λnλ in E, hence

(5.2)φλn*(t,,ξ*)φλ*(t,,ξ*)in L1(Ω) for almost all tT and all ξ*N.

We introduce the integral functional Φλ defined by

Φλ(t,x)=Ωφλ(t,z,Dx)𝑑zfor all (t,x)T×W01,p(Ω).

By [29], we know that (5.2) implies

Φλ(t,x)=Γseq(w)-Φλn(t,x)

with Γseq(w) denoting the sequential Γ-convergence of Φλn(t,) on W01,p(Ω)w (see [7]). Then, from [12, Theorem 3.3], it follows that

aλn(t,,)Gaλ(t,,)for almost all tT,

and so from [40], we conclude that

ddt+aλnPGddt+aλ.

Also, let Y=H=L2(Ω) and

F(t,x,λ)=SF1(t,,x(),λ)2,G(t,u,λ)={g(t,,λ)u():uL2(Ω)r(t,λ)}
U(t,λ)={uL2(Ω):uL2(Ω)r(t,λ)}.

Then hypotheses (H$F$1), (H${g}$), (H${r}$) imply that conditions (H$F$3), (H$G$), (H$U$) hold. So, the dynamics of (5.1) are described by an evolution inclusion similar to the one in problem (1.1).

Finally, let

L(t,x,λ)=ΩL1(t,z,x(z),λ)𝑑zfor all xL2(Ω),
H(t,u,λ)=ΩH1(t,z,u(z),λ)𝑑zfor all uL2(Ω).

Hypotheses (H${L_{1}}$), (H${H_{1}}$) imply that conditions (H$L$),(H$H$), respectively, hold. So, we can apply Theorems 4.6 and 4.7 and obtain the following result concerning the variational stability of problem (5.1).

Proposition 5.2

If the maps aλ are as above and hypotheses (H${F_{1}}$), (H${g}$), (H${r}$), (H${L_{1}}$), (H${H_{1}}$) hold, then for every (ξ,λ)L2(Ω)×E, problem (5.1) admits optimal pairs (that is, Σ(ξ,λ)) and

(ξ,λ)m(ξ,λ) is continuous on L2(Ω)×E,
(ξ,λ)Σ(ξ,λ) is sequentially USC from L2(Ω)×E into C(T,L2(Ω))×L2(T×Ω)w.

Award Identifier / Grant number: P1-0292-0101

Award Identifier / Grant number: J1-6721-0101

Award Identifier / Grant number: J1-7025-0101

Funding statement: V. Rădulescu was partially supported by the Romanian Research Agency in the framework of the Partnership program in priority areas – PN II, MEN – UEFISCDI, project number PN-II-PT-PCCA-2013-4-0614. The research of D. Repovš was supported in part by the Slovenian Research Agency (ARRS) grants P1-0292-0101, J1-6721-0101 and J1-7025-0101.

References

[1] Aubin J.-P. and Cellina A., Differential Inclusions, Springer, Berlin, 1984. 10.1007/978-3-642-69512-4Search in Google Scholar

[2] Aubin J.-P. and Frankowska H., Set Valued Analysis, Birkhäuser, Boston, 1990. Search in Google Scholar

[3] Balder E., Necessary and sufficient conditions for L1-strong-weak lower semicontinuity of integral functionals, Nonlinear Anal. 11 (1987), 1399–1404. 10.1016/0362-546X(87)90092-7Search in Google Scholar

[4] Barbu V., Nonlinear Semigroups and Differential Equations in Banach Spaces, Noordhoff International, Leyden, 1976. 10.1007/978-94-010-1537-0Search in Google Scholar

[5] Bressan A. and Colombo G., Extensions and selections of maps with decomposable values, Studia Math. 90 (1988), 69–85. 10.4064/sm-90-1-69-86Search in Google Scholar

[6] Brezis H., Opérateurs Maximaux Monotones et Semi-Groupes de Contractions dans les Espaces de Hilbert, North Holland, Amsterdam, 1973. Search in Google Scholar

[7] Buttazzo G., Semicontinuity, Relaxation and Integral Representation in the Calculus of Variations, Pitman Res. Not. in Math. Ser. 207, Longman Scientific & Technical, Harlow, 1989. Search in Google Scholar

[8] Buttazzo G. and Dal Maso G., Γ-convergence and optimal control problems, J. Optim. Theory Appl. 38 (1982), 385–407. 10.1007/BF00935345Search in Google Scholar

[9] Cellina A. and Ornelas A., Representation of the attainable set for Lipschitzian differential inclusions, Rocky Mountain J. Math. 22 (1992), 117–124. 10.1216/rmjm/1181072798Search in Google Scholar

[10] Cellina A. and Staicu V., On evolution equations having monotonicities of opposite sign, J. Differential Equations 90 (1991), 71–80. 10.1016/0022-0396(91)90162-3Search in Google Scholar

[11] Chen D. H., Wang R. N. and Zhan Y., Nonlinear evolution inclusions: Topological characterizations of solution sets and applications, J. Funct. Anal. 265 (2013), 2039–2075. 10.1016/j.jfa.2013.05.033Search in Google Scholar

[12] Defranceschi A., Asymptotic analysis of boundary value problems for quasilinear monotone operators, Asymptot. Anal. 3 (1990), 221–247. 10.3233/ASY-1990-3303Search in Google Scholar

[13] Denkowski Z. and Migorski S., Control problems for parabolic and hyperbolic equations via the theory of G and Γ convergence, Ann. Mat. Pura Appl. (4) 149 (1987), 23–39. 10.1007/BF01773923Search in Google Scholar

[14] Denkowski Z., Migorski S. and Papageorgiou N. S., An Introduction to Nonlinear Analysis: Applications, Kluwer, Boston, 2003. 10.1007/978-1-4419-9158-4Search in Google Scholar

[15] Denkowski Z., Migorski S. and Papageorgiou N. S., An Introduction to Nonlinear Analysis: Theory, Kluwer, Boston, 2003. 10.1007/978-1-4419-9158-4Search in Google Scholar

[16] Denkowski Z., Migorski S. and Papageorgiou N. S., On the convergence of solutions of multivalued parabolic equations and applications, Nonlinear Anal. 54 (2003), 667–682. 10.1016/S0362-546X(03)00093-2Search in Google Scholar

[17] Dontchev A. and Zolezzi T., Well-Posed Optimization Problems, Lecture Notes in Math. 1404, Springer, Berlin, 1994. 10.1007/BFb0084195Search in Google Scholar

[18] Fattorini H., Infinite Dimensional Optimization and Control Theory, Cambridge University Press, Cambridge, 1999. 10.1017/CBO9780511574795Search in Google Scholar

[19] Frankowska H., A priori estimates for operational differential inclusions, J. Differential Equations 84 (1990), 100–128. 10.1016/0022-0396(90)90129-DSearch in Google Scholar

[20] Gasinski L. and Papageorgiou N. S., Nonlinear Analysis, Chapman & Hall/CRC, Boca Raton, 2006. 10.1201/9781420035049Search in Google Scholar

[21] Gasinski L. and Papageorgiou N. S., Exercises in Analysis. Part 2: Nonlinear Analysis, Springer, New York, 2016. 10.1007/978-3-319-27817-9Search in Google Scholar

[22] Hu S. and Papageorgiou N. S., Handbook of Multivalued Analysis. Volume I: Theory, Kluwer, Dordrecht, 1997. 10.1007/978-1-4615-6359-4Search in Google Scholar

[23] Hu S. and Papageorgiou N. S., Time-dependent subdifferential evolution inclusions and optimal control, Mem. Amer. Math. Soc. 632 (1998), 1–81. 10.1090/memo/0632Search in Google Scholar

[24] Ito K. and Kunisch K., Sensitivity analysis of solutions to optimization problems in Hilbert spaces with applications to optimal control and estimation, J. Differential Equations 99 (1992), 1–40. 10.1016/0022-0396(92)90133-8Search in Google Scholar

[25] Ito K. and Kunisch K., Lagrange Multiplier Approach to Variational Problems and Applications, SIAM, Philadelphia, 2008. 10.1137/1.9780898718614Search in Google Scholar

[26] Kolpakov A. G., G-convergence of a class of evolution operators, Sib. Math. J. 29 (1988), 233–244. 10.1007/BF00969735Search in Google Scholar

[27] Lions J.-L., Quelques Méthodes de Résolution des Problémes aux Limites Non Linéaires, Dunod, Paris, 1969. Search in Google Scholar

[28] Liu Z., Existence results for evolution noncoercive hemivariational inequalities, J. Optim. Theory Appl. 120 (2004), 417–427. 10.1023/B:JOTA.0000015929.70665.87Search in Google Scholar

[29] Marcellini P. and Sbordone C., Dualita e perturbazione di funzionali integrali, Ric. Mat. 26 (1977), 383–421. Search in Google Scholar

[30] Papageorgiou N. S., Sensitivity analysis of evolution inclusions and its applications to the variational stability of optimal control problems, Houston J. Math. 16 (1990), 509–522. Search in Google Scholar

[31] Papageorgiou N. S., A convergence result for a sequence of distributed parameter optimal control problems, J. Optim. Theory Appl. 68 (1991), 305–320. 10.1007/BF00941570Search in Google Scholar

[32] Papageorgiou N. S., On parametric evolution inclusions of the subdifferential type with applications to optimal control problems, Trans. Amer. Math. Soc. 347 (1995), 203–231. 10.1090/S0002-9947-1995-1282896-XSearch in Google Scholar

[33] Papageorgiou N. S., On the variational stability of a class of nonlinear parabolic optimal control problems, Z. Anal. Anwend. 15 (1996), 245–262. 10.4171/ZAA/697Search in Google Scholar

[34] Papageorgiou N. S. and Kyritsi S., Handbook of Applied Analysis, Springer, New York, 2009. 10.1007/b120946Search in Google Scholar

[35] Papageorgiou N. S., Papalini F. and Renzacci F., Existence of solutions and periodic solutions for nonlinear evolution inclusions, Rend. Circ. Mat. Palermo (2) 48 (1999), 341–364. 10.1007/BF02857308Search in Google Scholar

[36] Repovš D. D. and Semenov P. V., Continuous Selections of Multivalued Mappings, Kluwer, Dordrecht, 1998. 10.1007/978-94-017-1162-3Search in Google Scholar

[37] Roubicek T., Relaxation in Optimization Theory and Variational Calculus, Walter de Gruyter, Berlin, 1997. 10.1515/9783110811919Search in Google Scholar

[38] Sokolowski J., Optimal control in coefficients for weak variational problems in Hilbert space, Appl. Math. Optim. 7 (1981), 283–293. 10.1007/BF01442121Search in Google Scholar

[39] Sokolowski J. and Zolesio J.-P., Introduction to Shape Optimization: Shape Sensitivity Analysis, Springer, Berlin, 1992. 10.1007/978-3-642-58106-9Search in Google Scholar

[40] Svanstedt N., G-convergence of parabolic operators, Nonlinear Anal. 36 (1999), 807–842. 10.1016/S0362-546X(97)00532-4Search in Google Scholar

[41] Zeidler E., Nonlinear Functional Analysis and its Applications II/B: Nonlinear Monotone Operators, Springer, New York, 1990. 10.1007/978-1-4612-0981-2Search in Google Scholar

Received: 2016-4-24
Accepted: 2016-4-27
Published Online: 2016-6-1
Published in Print: 2017-5-1

© 2017 by De Gruyter

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 26.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2016-0096/html
Scroll to top button