• 検索結果がありません。

THE CONSTRAINED CASE

N/A
N/A
Protected

Academic year: 2022

シェア "THE CONSTRAINED CASE"

Copied!
27
0
0

読み込み中.... (全文を見る)

全文

(1)

OF DIFFERENTIAL INCLUSIONS:

THE CONSTRAINED CASE

WOJCIECH KRYSZEWSKI Received 11 March 2002

We survey and announce some current results on the existence, the viability, and the topological structure of the viable solutions of differential equations and inclusion in Banach spaces under set constraints. Some new results concerning semilinear differential inclusions with state variables constrained to the so-called regular and strictly regular sets, together with their applications, are presented and discussed.

1. Introduction

It is our purpose to study solutions of the Cauchy problem for a semilinear dif- ferential inclusion

u(t)Au(t) +ϕt, u(t), tJ, uD, ut0

=x0D, (1.1)

whereϕ:J×DEis an upper-Carath´eodory set-valued map,J is an inter- val (i.e., a connected subset) of the real axis R, t0J, D is a closed subset of a Banach space E, and Ais the infinitesimal generator of aC0-semigroup {U(t)}t0 of bounded linear operators onE(A0 and/orE=RN is not ex- cluded). Problem (1.1) withD=Ewas studied by many authors—see the mono- graphs [43,44,46,51] and the rich bibliography therein—and diversity of results has been obtained. Here, we address the question of the topological character- ization of the set of all solutions (understood in an appropriate sense) of (1.1) under rather weak assumptions concerning the geometry ofD(satisfied, e.g., if Dis convex) and some natural boundary conditions. Ifϕis defined onJ×E, then our problem is intimately related to theviabilityor the invarianceproper- ties ofD(see [5,6]) with respect toϕ, that is, the existence of a solutionu:JE

Copyright©2003 Hindawi Publishing Corporation Abstract and Applied Analysis 2003:6 (2003) 325–351

2000 Mathematics Subject Classification: 49K24, 34C30, 34C25, 47D06 URL:http://dx.doi.org/10.1155/S1085337503204115

(2)

of (1.1) such thatu(t)Don J; however, we also determine the topological structure of the set of such viable solutions. Problems of this kind seem to be of importance not only from the viewpoint of the academic interest. An appro- priate characterization of the set of viable solutions and the use of topological methods involving the fixed-point index theory helps to establish the existence of the constrained periodic dynamics and stationary points (equilibria) of the system governed by the above differential inclusion.

We first state the standing hypotheses (Hi),i=1, . . . ,5, and recall some pre- liminary terminology. Throughout the paper,Estands for a separableBanach space.

(H1)ϕ:J×DEis aset-valued upper-Carath´eodorymap; that is, for each tJ,xD, the valueϕ(t, x) is anonempty, compact, andconvexsubsetE; the mapϕ(t,·) :DEisupper semicontinuousandϕ(·, x) :JEismeasurable.

(See, e.g., [31] or [8] for the concepts of continuity and measurability of set- valued maps; we only remark that, under separability, assumption measurability coincides with strong (Bochner) measurability.) Upper-Carath´eodory maps en- joy theweak superpositional measurabilityproperty: for a continuousu:JD, ϕ(·, u(·)) possesses a measurable selectionw(·). Hence, from the viewpoint of the solvability of (1.1), the regularity requirements seem to be rather minimal.

Simple examples show that we cannot dispense with the convexity assumption.

(H2)ϕhaslinear growth; that is, there is acL1loc(J,R) such that supzF(t,x)z

c(t)(1 +x) onJ×D. Therefore,ϕinduces the set-valued (Nemytskij) oper- atorNϕdefined on the setC(J, D) of continuous functionsJDinto the space L1loc(J, E) of (locally) Bochner integrable functionsJEgiven by

Nϕ(u) :=

wL1loc(J, E)|w(t)ϕt, u(t)a.e. onJ. (1.2)

(H3)ϕtransforms precompact subsets ofJ×Dinto compact ones.

This assumption is automatically satisfied ifϕis (jointly) upper semicontin- uous or if dimE <and cconst and seems to be a minimal compactness condition required in an infinite dimensional setting and in the presence of con- straints.

(H4) A closed densely defined linear operatorA:ED(A)Eis theinfini- tesimal generatorof aC0-semigroupᐁ= {U(t)}t0such thatU(t)exp(ωt) whereωRfort0.

It is clear (using an appropriate renorming procedure) that this does not re- strict generality (for details, cf. [50, Chapter VII] and [53,58]).

(H5)Dis aclosedsubsetE, invariant with respect toᐁ, that is,U(t)DD, t0.

This condition may be stated in terms ofAonly (see [50, Proposition VII.5.3, Remark VII.5.2] and [51]) holds if and only if lim inft0+dD(U(t)x)/t=0 for all

(3)

xD, wheredDis the distance function dD(x) :=inf

yDxy, xE. (1.3)

Givenx0E,t0J, and f L1loc(J, E), the function Mx0, t0;f(t)=Utt0

x0+ t

t0

U(ts)f(s)ds, tJ (1.4) is, by definition, themild solutionof the initial value problem

u(t)=Au(t) +f(t), ut0

=x0. (1.5)

Note that even the continuity off does not imply that (1.5) has astrongsolution, that is, an almost everywhere (a.e.) differentiable functionu:JEsuch that uL1loc(J, E), u(t0)=x0, and u(t)=Au(t) +f(t) a.e. on J; however, ifᐁis uniformly continuous(i.e.,IU(t)0 whent0+), or the functionv:t t

t0U(ts)f(s)dsis differentiable a.e. withvL1locandx0D(A), then the mild solution is a (unique) strong solution (see [53]).

A continuous functionu:JD is amild solution to (1.1) if there is w Nϕ(u) such thatu=M(x0, t0;w); hence, the setS(x0, t0) of all mild solutions of (1.1) coincides with the set of fixed points of the set-valued operatorM(x0, t0;·) Nϕdefined onC(J, D).

To state the results, we need to recall some other concepts. By theBouligand and Clarke tangent conestoDatxD, we understand the cones

TD(x) :=

yE: lim inf

h0+

dD(x+hy)

h =0

, CD(x)= uE| lim

h0+, y−→D x

dD(y+hu)

h =0

,

(1.6)

respectively (y−→D xmeans thatyconverges toxremaining inD). Observe that CD(x) is a closed convex cone andCD(x)TD(x). IfDis convex, thenTD(x)= CD(x) (see [8] for details).

Given a locally Lipschitz continuous function f :ER, by f(x;u) we de- note the Clarkegeneralized directional derivativeof f atxEin the direction uE. The Clarke generalized gradient of f at x, ∂ f(x) := {pE| p, u f(x;u) for alluE}, is a nonempty w-compact convex subset of E and f(x;u)=supp∂ f(x)p, u. It is clear thatu∂ f(x)(where∂ f(x):= {uE| p, u0 for allp∂ f(x)}is the negative polar cone) if and only if f(x;u) 0. In particular,CD(x)=∂dD(x)for allxD(see [21] for details).

Given a boundedΩE,

α(Ω) :=infε >0|Ωadmits a finite covering by sets of diameterε, β(Ω) :=infε >0|Ωadmits a finite covering byε-balls (1.7)

(4)

are the Kuratowski and Hausdorffmeasures of noncompactness, respectively.

These measures areregular, monotone, and nonsingular, that is, forγ=αorβ, we haveγ(Ω)=0 if and only ifΩis precompact,γ(Ω)γ(Ω) ifΩ, and γ({x} ∪Ω)=γ(Ω) for eachxE(see [1] for details).

In the spaceC(J, E) of continuous functionsJE, we consider thecompact- opentopology. Thus, C(J, E) is a Fr´echet (locally convex metrizable and com- plete) space with the metric

ρ(u, v) :=max

k1

2kpk(uv)

1 +pk(uv), u, vC(J, E), (1.8) where pk(u)=suptJku(t)and{Jk}k=1is a family of compact intervals such that JkintJk+1 and J=

k=1Jk. Thus, C(J, E) has theprojective topology in- troduced by restrictions{πk:C(J, E)C(Jk, E)|k1},πk(u)=u|Jk foru C(J, E). A setSC(J, E) is compact if and only if Sk:=πk(S) is compact for eachk1. It is also easy to see that Sis homeomorphic to the inverse limit lim invk→∞Skof the inverse system{Sk, πkl}, whereπklis the restriction of func- tions fromSktoJl(lk).

2. Existence

Among many existence results (see, e.g., [43,44,51,57] or [46] with huge bibli- ography), the one due to Bothe [16, Theorem 7.2, Corollary 7.1] seems to be the most general.

Theorem2.1. Assume that

ϕ(t, x)TD(x)= ∅ for a.e.tJand allxD. (2.1) Then, the set-valued mapS:D×JC(J, D), assigning to(x0, t0)D×Jthe set S(x0, t0)of all mild solutions of (1.1), is upper semicontinuous with nonempty com- pact values provided one of the following conditions holds:

(i)for any boundedD,

hlim0+βϕJ(t, h)×k(t)β(Ω) for eachtJ, (2.2) whereJ(t, h) :=(th, t+h)JandkL1loc(J,R); or

(ii)the semigroupis compact (ᐁis compact ifU(t)is compact for allt >0).

Remark 2.2. (1) Observe that a compactness assumption (2.2) implies (H3).

Moreover, ifD=E, then a weaker condition

βϕ{t} ×k(t)β(Ω) onJ (2.3) for boundedΩD(i.e., (2.2) withJ(t, h)= {t}) is also sufficient for the asser- tion ofTheorem 2.1(see [46]). IfD=E, then (2.3), together with (H3), is also sufficient provided we know more aboutD—see Theorems5.4and6.6.

(5)

(2) In particular, if f :J×DEis single valued, for anyxD,tJ, f(·, x) is measurable and maps compact subsets ofJinto compact sets, f(t,·) is locally Lipschitz (uniformly with respect totfrom compact subsets ofJ) and satisfies (2.1), then (1.1) has a unique mild solution which depends continuously on (x0, t0)D×J. Indeed, under these assumption, f satisfies all hypotheses of Theorem 2.1locally; hence, a local unique mild solution exits. To establish the result, we apply the usual continuation method.

Observe that, for anyxD,yE, andh >0, dD

U(h)x+hyhU(h)yy+dD

U(h)(x+hy); (2.4) thus, condition (2.1) implies

F(t, x)TD(x)= ∅ for a.e.tJand allxD, (2.5) where

TD(x) :=

yE: lim inf

h0+

dD

U(h)x+hy

h =0

(2.6) (in general,TD(x) is not a cone).Theorem 2.1has been proved under assump- tion (2.5) instead of (2.1). Condition (2.5) is strictly weaker than our (2.1). To see this, considerE=R,U(t)=etfort0, and letD=[1,1]. Then,Dis in- variant with respect toU(t) (i.e.,U(t)DD), butTD(1)=(−∞,0](−∞,1]= TD(1). Both conditions (2.1) and (2.5) are natural since we have the following proposition.

Proposition 2.3. Condition (2.5) is necessary for the existence of solutions of (1.1). Precisely, ifϕis upper semicontinuous and, for everyx0Dandt0J, (1.1) has a mild solution, then (2.5) is satisfied.

Next, in view of the inequality dD

x+h(Ax+y)dD

U(h)x+hyx+hAxU(h)x (2.7)

valid for allxD(A), we see that (2.5) implies

Ax+F(t, x)TD(x)= ∅ for everytJ, xDD(A). (2.8)

In caseE=RN, (then,Ais defined everywhere and bounded) condition (2.8) is sufficient and necessary for the existence. Simple, constructive and based on the technique of the so-called proximal aiming, proof of this fact is given in [14].

(6)

Moreover, (2.8) may be relaxed (instead ofTD(·), we can consider the convex envelope convTD(·)).

Some other tangency conditions leading to the existence in the unconstrained case are considered, for example, in [19,32] or [51].

3. Solution sets of the semilinear system

A classical result of Aronszajn [4] states that the solution set for the Cauchy problem inRNis a compactRδ-set; it is also true for differential inclusions—see for example, [25,40,41] (autonomous systems), [26,38] (nonautonomous), cf.

[7, Corollary 5, page 109] and the surveys [33,38]; asymptotic problems have been studied recently for example in [2,3]. (This paper provides an extensive survey on the characterization of the fixed-point set of set-valued maps.)

Recall that a compact metric spaceX is anRδ-setif there is an ANR (abso- lute neighborhood retract)Y containingXas a closed subspace such thatXis contractible in each of its open neighborhoods (i.e., given an open neighbor- hoodV ofXinY, there is a continuous maph:X×[0,1]V withh(x,0)=x andh(x,1)=x0V, for allxX). (Rδ-sets are sometimes calledcell-like sets.) TheRδ-property is a homotopy invariant: if compactaX1andX2are homotopy equivalent andX1Rδ, then so isX2. Similarly, given an ANRZcontainingXas a closed subspace, ifXRδ, thenXis contractible in each of its neighborhoods inZ. It is clear that ifXis a subset of a metric spaceT and there is a decreas- ing family{Xn}n1 of closed contractible sets such thatγ(Xn)0, whereγis a regular, monotone and nonsingular measure of noncompactness onT, then XRδ. The celebrated result of Hyman [45] states that ifXRδ, then there is a decreasing sequence of contractible compacta{Xn}containingX as a closed subspace such thatX=

n1Xn.Rδ-sets are connected, have trivial shape, and are acyclic with respect to any continuous (co)homology theory, that is, they have the same (co)homology as a one-point space. Recall (see [35]) that ifXis homeomorphic to the inverse limit lim invn1Xnof the countable inverse system {Xn;πnm:XmXn, nm}andXnRδ for all n, thenXRδ. In particular, SC(J, E) isRδif and only if, for alln1, the setSk=πk(S) of restrictions of functions fromStoJkisRδ.

We establish theRδ-structure of the setS(x0, t0) of mild solutions of the con- strained semilinear system (1.1). This result is known whenD=E(see [16,46]);

it is also believed that, in fact, the existence implies theRδ-structure (see [20]).

See also [22,43,44] for other results and many bibliographical comments. The constrained case seems to be more involved. The result holds ifDis invariant with respect toᐁ, convex, and intD= ∅, or ifDis convex proximinal (i.e., each xEadmits a nearest point inD) andᐁis nonexpansive (see [11,17,43,44]) andϕobeys the tangency condition (2.1). For the finite-dimensional situation, see [13,29,30,31,42].

To see that even in the convex case the situation is more complicated, we consider the following example.

(7)

Example 3.1(see [31]). LetD=[0,), f(x)=2|x|sgnx, andx0=0. Then, f(x)TD(x)=CD(x)=RonD, butu(t)= −t2Dis a solution of (1.1) (with ϕ=f,A0, andt0=0).

The example shows that even a stronger tangency condition

ϕ(t, x)TD(x) (3.1)

(orϕ(t, x)CD(x)) onJ×Ddoes not preventsomesolutions to leaveD. Hence, in the caseϕis defined onJ×E, the “unconstrained” results fail to help in the characterization of the set of solutions surviving inD: to get convex-constrained results, we need slightly more involved arguments. In the next section, we pro- vide a result valid for a general convex closedD.

The topological structure ofS(t0, x0) changes dramatically in case the setD is not convex. We study some examples (for simplicity, we consider a finite- dimensional situation withD(automatically) invariant with respect to the semi- group generated byA0).

Example 3.2(see [31]). LetD= {(x, y)R2|x0, y=x2} ∪[0,)× {0}and f(x, y)=(1,2y) onD. Then, f(z)TD(z) forzDand the problemu= f(u),u(0)=0 has only two solutions:u(t)=(t,0) andu(t)=(t, t2) fort0.

Example 3.3. LetE=R3,D:={x=(x1, x2, x3)E| |x| ≤

2 andx12+x22x3}, S:= {xD|x21+x22=1 andx3=1}, andZ= {x E|x21+x221 andx3=1}. Next, forxD, let

ϕ(x)=

Z forxD\S ,

convZ

x2, x1,0 forxS. (3.2) Clearly,ϕ:DEis upper semicontinuous andϕ(x)TD(x)= ∅onD. But it is easy to see thatS(x0, t0) (witht0=0,x0=0) is homeomorphic to the unit sphere S1:= {xR2| x =1}; hence, it is not anRδ-set. Notice that, for allxD, x=0, the Bouligand and the Clarke tangent conesTD(x) andCD(x) coincide;

however,TD(0)=CD(0) andF(0)CD(0)= ∅.

In the above example, if we considerϕsatisfying (2.1) with the Clarke cone replacing the Bouligand one, then the situation becomes clear. However, it is not true that such a procedure would be the general remedy.

Example 3.4(see [14]). LetD=S1S1, whereSi= {x=(x1, x2)R2|(x1i)2 +x22=1}, and, forxD, let

ϕ(x)=

x2,1x1

forxS1, x2,1 +x1

forxS1. (3.3)

(8)

Then,ϕ(x)TD(x)=CD(x) onD, butS(x0, t0) (witht0=0 andx0=0) is not connected.

It seems, therefore, that in order to state the correct tangency condition which implies the expected topological structure of solution sets to (1.1), we should replace in (2.1) the Bouligand cones by the Clarke ones, that is, assume that

(t, x)J×D, F(t, x)CD(x)= ∅ (3.4) and take care of the geometry of the involved set D. The first attempt in this direction (forE=RN andA0) was done by Plaskacz [54], where he studies the classρof sets (calledproximate retractsin [37]) and assumes (3.1); Plaskacz’s result was extended to the Hilbert space context (see [39]). Up to now, the most general results for the finite dimensional case were given in [13]. Below, we will generalize them to the present infinite dimensional situation.

4. Convex case

The general strategy to obtain theRδ-structure of solution sets is to approximate ϕin an appropriate way by a sequence of auxiliary set-valued maps{ϕn}possess- ing locally Lipschitz selections and, then, to show thatS(x0, t0) is an intersection of solution sets corresponding toϕn. In the constrained case, the main difficulty is to assure that mapsϕnand their locally Lipschitz selections obey the necessary tangency condition implying existence.

We first deal with the general convex case. We state the result with a sketch of the proof in order to show how the above described procedure works.

For a convex closed subsetXof a normed spaceYandxX,

CX(x)=TX(x)=SX(x), (4.1) where

SX(x) :=

h>0

Xx

h . (4.2)

Our improvement of the mentioned results on the structure of solutions living in convex sets (cf. [11,18]) is based upon the following lemma.

Lemma4.1. IfΦ:XY is upper semicontinuous with closed convex values and, for eachxX,

Φ(x)TX(x)= ∅, (4.3)

then, for anyε >0, there is a locally Lipschitz mapF:XYsuch that

xX, F(x)ΦBX(x, ε)+BY(0, ε) (4.4)

(9)

(whereBX(x, r) := {yX|d(x, y)< r}is the ball inX; usually the subindexXis suppressed from the notation) and

xX, F(x)TX(x). (4.5)

Remark 4.2. IfY is complete, then the existence of acontinuous(single-valued) mapF:XY satisfying conditions (4.4) and (4.5) follows from a general result from [14]. Here, we need no completeness and improve this result obtaining a locally Lipschitzianε-selection.

Proof. Takeε >0 andxX. There isv(x)Ysuch that v(x)Φ(x) +BY

0,ε

4

SX(x) (4.6)

in view of (4.3) and (4.1). Hence, by (4.2), there isα(x)>0 such that

x+α(x)v(x)X. (4.7)

By upper semicontinuity, chooseγ(x), 0< γ(t, x)< ε/4 such that ΦBXx,2γ(x)Φ(x) +BY

0,ε

2

(4.8) andδ(x), 0< δ(x)<min{γ(x), γ(x)/α(x)}.

Let{λs:X[0,1]}sSbe a locally finite locally Lipschitz partition of unity refining the open cover{BX(x, δ(x)α(x))}xX. For anysS, there isxsXsuch that suppλsBX(xs, δsαs) where we have putδs:=δ(xs) andαs:=α(xs). Addi- tionally, we setvs:=v(xs) andγs:=γ(xs).

For anysS, consider a mapFs:XY given by Fs(x) := 1

αs

xsx+vs, xX. (4.9) ForsS,xX,

x+αsFs(x)=xs+αsvsX (4.10) in view of (4.7); hence,

Fs(x)SX(x)TX(x). (4.11) Clearly,Fs,sS, is Lipschitz continuous (with the Lipschitz constantαs1).

A mapF:XY defined by the formula F(x) :=

sS

λs(x)Fs(x), xX, (4.12)

satisfies the requirements ofLemma 4.1.

(10)

Remark 4.3. In the course of the proof, we have not used the lower semiconti- nuity ofTX(·). Instead, the following astonishingly simple observation (already employed in a different situation in [24]) has been used.

IfX is a convex closed subset of a normed space Y, then, for every x0 X,v0SX(x0) andα0>0 such thatx0+α0v0X, an affine mappingg(x)= (1/α0)(x0x) +v0,xX, provides a selection ofSX(x). This proves the lower semicontinuity of bothSX(·) andTX(·).

Theorem4.4. If Dis convex and (2.1) holds, thenS(x0, t0)is anRδ-set inC(J, E) provided that (2.2) is satisfied oris compact.

Proof. To illustrate the setting, we consider an upper-semicontinuousϕ. Take (x0, t0)D×J and choose a family{Jk=[ak, bk]}k=1of compact subintervals inJ such thatJkintJk+1,Jk=J, andt0Jk for allk1. Fixk1 and let Φ:Jk×DY:=R×E,Φ(t, x)= {0} ×ϕ(t, x) (inY, we consider the norm (t, x) =max{|t|,x}). ByLemma 4.1, there is a locally LipschitzFn:Jk×D Y such thatFn(t, x)TJk×D(t, x) andFn(t, x)Φ(B((t, x), n1)) +BY(0, n1) on Jk×D(n1). Define fn:=pFnwherep:YEis the projectionp(t, x)=x.

Then, fn(t, x)ϕn(t, x) :=convϕ(B((t, x), n1)) +B(0, n1) andfn(t, x)TD(x) onJk×D. Clearly, forn1,ϕ(t, x)ϕn(t, x)ϕn+1(t, x) onJk×D; thus,

∅ =Sknx0, t0

Skn+1, Skx0, t0

n1

Sknx0, t0

, (4.13)

whereSk(x0, t0) (resp.,Skn(x0, t0)) stands for the set of all mild solutions onJkof (1.1) (resp., of (1.1) withϕreplaced byϕn).

LetunSkn(x0, t0) for alln1. We then show that there exists a subsequence (unm)m1such thatunmu0Sk(x0, t0) inC(Jk, E) asm→ ∞. It follows that

Skx0, t0

=

n=1

Sknx0, t0

, β0

Sknx0, t0

−→0 asn−→ ∞,

(4.14)

whereβ0stands for the Hausdorffmeasure of noncompactness inC(Jk, E).

For eachn1,zJk, andyD, the problem

vAv+fn(t, v), v(z)=y (4.15) admits a unique solutionvn(·;z, y) : [ak, bk]D, which depends continuously on (z, y) (seeRemark 2.2). Define a homotopyh: [0,1]×Skn(x0, t0)C(J, E) by

(11)

the formula h(λ, u)(s)

:=

u(s) ifs

ak+λt0ak, bk+λt0bk

, vn

s;ak+λt0ak

, uak+λt0ak

ifs

ak, ak+λt0ak , vns;bk+λt0bk, ubk+λt0bk ifsbk+λt0bk, bk

(4.16) for uSkn(x0, t0) andλ[0,1]. It is easy to see that his continuous. More- over, we haveh([0,1]×Skn(x0, t0))Skn(x0, t0). From the continuity ofh, we in- fer thath([0,1]×Skn(x0, t0))Skn(x0). Finally, observe thath(1, u)=vn(·;t0, x0) and h(0, u)=u for every uSkn(x0, t0), that is, Skn(x0) is contractible. Hence,

Sk(x0, t0)Rδand so isS(x0, t0).

5. Epi-Lipschitz case

An important role in optimization is played by the so-called epi-Lipschitz sets.

This notion (in the finite-dimensional context) has been introduced by Rock- afellar [56]. The corresponding notion for subsets of a Banach space has been studied in [14].

Definition 5.1. A closed set DE isepi-Lipschitzif, for allx0D, there are a neighborhoodU ofx0 (inE), a Banach spaceZ, a topological isomorphism L:Z×REwithL(z0, λ0)=x0, and a locally Lipschitz functiong:ZRsuch that

DU=UL(Epig), (5.1)

where Epig:= {(z, λ)|g(z)λ}is theepigraphofg.

Proposition5.2 (see [21,23,56]). IfDE=Rn is closed, then the following conditions are equivalent:

(i)Dis epi-Lipschitz;

(ii)for any xD,intCD(x)= ∅ (or, equivalently, the Clarke normal cone ND(x) :=CD(x)is pointed, that is,ND(x)(ND(x))= {0});

(iii)for anyx∂D,CD(x)=D(x), where, for yR,D(y) :=dD(y) dE\intD(y)and0∂∆D(x).

Implication (i)(ii) is obvious. Implication (iii)(ii) follows since if 0

D(x) for x∂D, then∂D(x)CD(x) (see [21, Theorem 2.4.7]); hence, ND(x)(∂∆(x))=

λ0λ∂D(x) andND(x) is pointed. Both these facts hold if dimE= ∞. In order to prove implications (ii)(i) and (ii)(iii), we need typ- ically finite-dimensional arguments. The author does not know whether they hold when dimE= ∞. The partial answer is given in the following result.

(12)

Proposition5.3. IfdimE≤ ∞, then(i)(iii).

Theorem5.4. IfDis an epi-Lipschitz set and (3.4) is satisfied, then, for eacht0J andx0D, the setS(x0, t0)of all mild solutions of the initial value problem (1.1) is anRδ-set provided the compactness condition (2.3) is satisfied or the semigroupis compact. One may replace (3.4) by (2.1) in caseϕis single-valued.

Proof. The last statement is easy. ForxintD,CD(x)=TD(x)=E. Takex∂D and a sequence (yn) inDsuch thatynx. Then, for alltJ,ϕ(t, yn)TD(yn) and, in view of [8, Theorem 4.1.9],

ϕ(t, x)=lim

n→∞ϕt, yn

Lim inf

y−→D x TD(y)=CD(x) (5.2) and the first part of the theorem applies.

As concerns the first part, we again construct, for eachn1, a map fn:J× DEsuch that fn(·, x) is measurable, fn(t,·) is locally Lipschitz (uniformly with respect tot), each pointxD has a neighborhood W with fn(Jk×W) lying in a compact subset ofE,fn(t, x)CD(x), and

fn(t, x)ϕn(t, x) :=convϕ{t} ×

Bx, n1D+B0, n1 (5.3) for alltJandxD. The construction recalls that fromLemma 4.1, but it also makes a strong use of the facts thatCD(x)=∂∆D(x) and 0∂∆D(x) on∂D.

Namely,

fn(t, x)=

sS

λs(x)ws(t), (5.4)

where{λs}sSis an appropriate locally Lipschitz partition of unity and, forS, ws:JEis a measurable finite-valued function such that∆D(y;ws(t))<0 for all tJandysuppλs. Having this, the proof concludes similarly as above.

6. Regular case

It is clear that epi-Lipschitz sets have nonempty interiors, and, therefore, neither convex sets nor Plaskacz’s proximate retracts are epi-Lipschitz in general. We introduce a class of sets that encompasses epi-Lipschitz or convex sets as well as proximate retracts. Namely, we will deal with the so-called regular domains.

Definition 6.1. We say that a closed setDEisregularif, for anyx∂D, lim inf

y−−→E\D x

∂dD(y)>0, (6.1)

where

∂dD(x):= inf

p∂dD(x)p. (6.2)

(13)

It is clear that

∂dD(x)= sup

u1

inf

p∂dD(x)p, u = − inf

u1dD(x;u). (6.3) Observe that regularity ofDmeans that the distance functiondDhas no criti- cal points in a neighborhood ofDintersected with the complement ofD. If there is a neighborhoodUofDsuch that infyU\D|∂dD(y)|>0, thenDis regular;

ifDis regular compact (or∂Dis compact), then such a neighborhood exists. It appears that regular sets are well designed to study solutions of (1.1) in case of a nonexpansive semigroupᐁ. In order to study a general situation we will also deal with the so-called strictly regular sets.

Definition 6.2. We say that a closed setDEisstrictly regularif there is anr >0 such that

yB(D,r)inf \D

∂dD(y)>0. (6.4)

Clearly, strictly regular sets are regular and compact regular sets are strictly regular.

The class of (strictly) regular sets has been introduced in [24] in a different (and a bit more general) setting and studied in the context of equilibria. This class is rich: for instance, the setDinExample 3.3is strictly regular and the set DfromExample 3.4is not regular.

Example 6.3. (i) Anyconvexclosed setDEis strictly regular: in fact we easily show that|∂dD(y)| ≥1 for allyE\D(see [24]).

(ii) Suppose that a closedDEisproximinal, that is, there is a neighborhood U ofDsuch that, for allyU, the setπD(y) := {zD| yz =dD(y)} =

. If, for any yU,πD(y)Lim infzyπD(z)= ∅(Lim inf denotes the lower limit in the sense of Painlev´e-Kuratowski (see, e.g., [8, Definition 1.4.6]), thenD is regular. Indeed, take yU\D, letxπD(y)Lim infzyπD(z), putu:=x y, and take sequencesyny,hn0+. There isxnπD(yn) such thatxnx.

Hence, dD

yn+hnudD yn

hnxnynu+dDyn+hnxnyndDyn. (6.5) We easily check thatdD(yn+hn(xnyn))=(1hn)dD(yn); therefore,dD(y;u)

uand|∂dD(y)| ≥1. IfUis a ball aroundD, thenDis strictly regular.

(iii) In particular, all proximate retracts (i.e., proximinal closed sets for which πD(x) is a singleton for allxU) are regular,πD, in this case, is continuous.

(iv) IfDis epi-Lipschitz, then it is regular.

(v) A smooth (i.e.,C1) Banach submanifoldMEof codimension 1 is reg- ular.

参照

関連したドキュメント

So far as the large time behaviour of solutions is concerned, we have noticed a few papers (e.g. [5, 9, 10, 14]) including some results about the ω-limit set of each single solution

σ(L, O) is a continuous function on the space of compact convex bodies with specified interior point, and it is also invariant under affine transformations.. The set R of regular

[37] , Multiple solutions of nonlinear equations via Nielsen fixed-point theory: a survey, Non- linear Analysis in Geometry and Topology (T. G ´orniewicz, Topological Fixed Point

This degree theory possesses all the usual properties of the deterministic degree theory such as existence of solutions, excision and Borsuk’s odd mapping theorem. Applications

The aim of this paper is to present general existence principles for solving regular and singular nonlocal BVPs for second-order functional-di ff erential equations with φ- Laplacian

Gatsori, “Controllability results for nondensely defined evolution differential inclusions with nonlocal conditions,” Journal of Mathematical Analysis and Applications, vol..

We show existence of the first non-trivial curve C of the Fuˇ cik spectrum which is used to obtain the variational characterization of a second eigenvalue of the problem defined

In this paper, we study a problem for second order ordinary differential equation with mixed nonlocal boundary conditions combined weighting integral boundary conditions with