• 検索結果がありません。

RESTRICTIVE METRIC REGULARITY AND GENERALIZED DIFFERENTIAL CALCULUS IN BANACH SPACES

N/A
N/A
Protected

Academic year: 2022

シェア "RESTRICTIVE METRIC REGULARITY AND GENERALIZED DIFFERENTIAL CALCULUS IN BANACH SPACES"

Copied!
28
0
0

読み込み中.... (全文を見る)

全文

(1)

PII. S0161171204405183 http://ijmms.hindawi.com

© Hindawi Publishing Corp.

RESTRICTIVE METRIC REGULARITY AND GENERALIZED DIFFERENTIAL CALCULUS IN BANACH SPACES

BORIS S. MORDUKHOVICH and BINGWU WANG Received 20 May 2004

We consider nonlinear mappingsf:X→Ybetween Banach spaces and study the notion of restrictive metric regularityoffaround some point ¯x, that is, metric regularity offfromX into the metric spaceE=f (X). Some sufficient as well as necessary and sufficient conditions for restrictive metric regularity are obtained, which particularly include an extension of the classical Lyusternik-Graves theorem in the case whenfis strictly differentiable at ¯xbut its strict derivative∇f (¯x)is not surjective. We develop applications of the results obtained and some other techniques in variational analysis to generalized differential calculus involving normal cones to nonsmooth and nonconvex sets, coderivatives of set-valued mappings, as well as first-order and second-order subdifferentials of extended real-valued functions.

2000 Mathematics Subject Classification: 49J52, 49J53, 58C20, 90C31.

1. Introduction and preliminaries. This paper is devoted to metric regularity and the generalized differentiation theory in variational analysis which has been well recog- nized as a fruitful area in mathematics mostly oriented in optimization-related prob- lems and their applications. On the other hand, variational principles and methods of variational analysis are widely applied to the study of a broad range of problems that may not be of a variational nature. We refer the reader to the book by Rockafellar and Wets [33] for a systematic exposition of the key features of variational analysis in finite dimensions.

Since nonsmooth objects (sets with nonsmooth boundaries, set-valued mappings, and extended real-valued functions) appear naturally and frequently in constrained optimization and related areas, generalized differentiation is one of the major parts of variational analysis. In this paper, we focus on Fréchet-like generalized differential constructions and their sequential limits that play an important role in nonsmooth variational analysis and its applications; see, respectively, [4,18,20,25,33] with the references and commentaries therein for developments and applications in finite- dimensional and infinite-dimensional spaces.

Most of the results previously obtained for the mentioned constructions require that the Banach spaces in question be Asplund, that is, every separable subspace of them have a separable dual. This includes all spaces with Fréchet differentiable renorms or bump functions, in particular, every reflexive Banach space. On the other hand, there are important Banach spaces that are not Asplund (e.g., the classical functional spaces L1,L, andC). In what follows, we are not going to impose the Asplund property and we will consider thegeneral Banach space setting.

(2)

The results obtained in this paper concernfirst-orderandsecond-ordercalculus; the latter deals with chain rules for second-order subdifferentials. An important property of mappings used in the development of generalized differential calculus is the so-called metric regularity. Recall that a mappingf:X→Y between Banach spaces ismetrically regulararound ¯xif there areµ >0 and neighborhoodsUof ¯xandVof ¯y:=f (x)¯ such that

dist

x;f−1(y)

≤µf (x)−y ∀x∈U , y∈V . (1.1) The celebrated Lyusternik-Graves theorem ensures this property for a mapping f strictly differentiable at ¯xwith thesurjective derivative∇f (x)¯ :X→Y. Moreover, the surjectivity of∇f (x)¯ is alsonecessary for the metric regularity offaround ¯x. We re- fer the reader to the original papers of Lyusternik [17] and Graves [9], as well as to the recent discussions in [7,11] and the works therein.

It is valuable for the theory and applications to relax the surjectivity assumption and to extend a class of mappings for which one can use metric regularity techniques. Since the surjectivity of∇f (x)¯ is necessary for the metric regularity of strict differentiable mappings, one has to modify the above concept to cover mappings with nonsurjective derivatives. In what follows, we consider the metric regularity for mappingsffromX into the image spacef (X), which is a metric space, but in general is far from being Banach, and call this notion therestrictive metric regularity(RMR) offaround ¯x.

Section 2is devoted to the study of RMR for mappingsf:X→Y between Banach spaces X and Y. Although the RMR property concerns in fact the metric regularity of f :X→E with the metric spaceE:=f (X)⊂Y and hence can be treated by the metric space regularity theory (cf. [6,11]), we take an advantage of using the Banach space structure onXandY as well as the strict differentiability of mappings when it applies. In this way, we establish relationships between RMR of nonlinear (generally nonsmooth) mappings and their linear approximations, derive necessary and sufficient conditions for the RMR property via approximations, and obtain efficient criteria for RMR of strictly differentiable mappings with nonsurjective derivatives that extend the Lyusternik-Graves theorem to such mappings important for applications.

InSection 3, we give applications of the RMR property and related results to first- order calculus rules for generalized normals, coderivatives, and subgradients of sets, set-valued mappings, and extended real-valued functions in Banach spaces. Most of the calculus rules obtained are new even under surjectivity assumptions, ensuring the classical metric regularity of mappings involved in compositions. The principal first- order results concern computing generalized normals to inverse images of sets under strictly differentiable mappings with possibly nonsurjective derivatives; related chain rules for coderivatives and subgradients follow from them via a geometric approach.

Section 4concerns chain rules for two kinds of second-order subdifferentials (“nor- mal” and “mixed”) generated by the corresponding coderivatives of first-order subgra- dient mappings. We derive an exact formula for computing mixed second-order subd- ifferentials and obtain an efficient upper estimate for normal ones, which becomes an equality under natural assumptions discussed below (in particular, when the domain spaceXis reflexive).

(3)

Finally, Section 5 contains applications of RMR and calculus rules developed in Section 3to the so-calledsequential normal compactness(SNC) properties of sets, set- valued mappings, and extended real-valued functions in infinite dimensions. The latter properties are automatic in finite-dimensional spaces while playing a crucial role in infinite-dimensional variational analysis; see more discussions below. We obtain effi- cient conditions ensuring the preservation of the SNC properties under some compo- sitions involving RMR mappings.

Our notation is basically standard; compare [25,33]. Unless otherwise stated, all the spaces considered are Banach, with the norm·and the canonical dual pairing ·,·; BXstands for the closed unit ball inX(we use the simplified notationBandBfor the dual balls inXandXif no confusion arises). Given spacesXandY,Ꮾ(X, Y )denotes the collection of all bounded linear operators fromXtoY.

LetL⊂Xbe a closed subspace ofX. AprojectionfromXtoLis an operatorπL(X, X)such that the image ofXunderπL isLand the restriction ofπLonLis the identity operator. We will drop the subindexLif there is no confusion. Recall thatL iscomplemented inX if there is a closed subspaceM ofX withX=L⊕M. It is well known thatLis complemented inXif and only if there is a projection fromXtoL, which happens, in particular, whenLis of finite dimension or finite codimension. Note also that every closed subspace ofXis complemented inXif and only if Xis isomorphic to a Hilbert space.

2. Restrictive metric regularity. Recall that metric regularity is a concept defined for mappings between metric spaces; see [11] and the references therein. Given two metric spaces(E1, d1)and(E2, d2)and a mappingf:E1→E2, we say thatfismetrically regulararound ¯xif there are neighborhoodsUof ¯xandVoff (x)¯ and a constantµ >0 such that

dist

x;f−1(y)

:= inf

v∈f−1(y)

d1(x, v)≤µd2

f (x), y

∀x∈U , y∈V . (2.1) When both spacesEiin (2.1) are Banach (withX:=E1,Y :=E2, and the same notation

·for the norms onXandY), (2.1) obviously reduces to (1.1). If in this case,f:X→Y is strictly differentiable at ¯x, then the surjectivity of∇f (x)¯ :X→Y isnecessary and sufficient for the metric regularity of f around ¯x. What could be said about metric regularity of f when∇f (x)¯ is not surjective? We suggest to consider the following property concerning metric regularity of therestrictivemappingf:X→f (X).

Definition2.1. Letf:X→Ybe a mapping between Banach spaces.fis said to have the RMR property around ¯x, orf is RMR around this point, if the restrictive mapping f:X→f (X)betweenXand the metric spacef (X)⊂Y, whose metric is induced by the norm onY, is metrically regular around ¯xin the sense of (2.1).

One can easily see, by the classical open mapping theorem, that forlinearmappings f, the RMR property always holds if the subspacef (X)is closed in Y. However, the situation is much more complicated fornonlinear mappings when the RMR property may be violated even in the simplest cases as, for example, forf (x)=x2around ¯x= 0R.

(4)

Note that, although metric regularity is a local property, the imagef (X)is a nonlocal object in the sense that some pointsxsituated far from ¯xmay contribute to the image offaroundf (x). Indeed, consider the mapping¯ f:R2R2defined by

f (x):=



x1−1,0

ifx1>1, x1, x21

otherwise, (2.2)

for all (x1, x2)∈R2. Thenf is not RMR around(0,0), while the localized mapping f:BR2→f (BR2)satisfies the metric regularity property (2.1) around this point. Thus it might be more appropriate to consider the metric regularity off:U→f (U )for some neighborhoodU of ¯x. However, the latter property is obviously equivalent to RMR of the modified mappingf:X→Y defined by

f (x) :=



f (x) ifx∈U ,

y otherwise, (2.3)

wherey is any fixed point of Y different from f (x). This allows us to confine our¯ consideration to the RMR property introduced above.

In the remaining part of this section, we establish effectivenecessary conditions, sufficientconditions, andcharacterizationsfor the RMR property of mappings between general Banach spaces. We start with an important necessary condition, which is widely used in what follows.

Theorem2.2. LetXandY be Banach spaces, and letf:X→Y be RMR aroundx¯ and Fréchet differentiable at this point. Then∇f (x)(X)¯ is closed inY.

Proof. Chooseδ >0 such that for anyx∈x¯B, there isx∈f1(f (x))satisfying x−x¯ ≤µf (x)−f (x0)for some constantµ >0. Lety0clA(X)forA:= ∇f (x).¯ Then there areyk→y0withyk∈A(X)and such thatyk+1−yk<1/2kfor allk∈ N:= {1,2, . . .}. To proceed, we build a sequence{xk} ⊂Xwith the following properties:

xk+1−xk<

2k, yk−A

xk< 1

2k, k∈N. (2.4)

Definexkiteratively. First letx1be any point withA(x1)=y1. Then havingx1, . . . , xk

that satisfy (2.4), definexk+1as follows. Fixu∈−1(yk+1)−xkand choose a smallt >0 such thattu< δand

f (x¯+th)−f (¯x)

t −A(h)

< 1

2k+2 wheneverh∈max

u,

2k B. (2.5) This gives, in particular, that

f (¯x+tu)−f (x)¯

t −A(u)

< 1

2k+2, (2.6)

(5)

which implies the estimates

f (x¯+tu)−f (x)¯ < tA(u)+ 1 2k+2

=tyk+1−A

xk+ 1 2k+2

≤tyk+1−yk+yk−A

xk+ 1 2k+2

< t 1

2k+ 1 2k+ 1

2k+2

<3t 2k.

(2.7)

Using now the RMR property offaround ¯x, selectx∈Xwithf (x)=f (x¯+tu)and x−x¯<3µt

2k . (2.8)

Puttingv:=(x−x)/t¯ andxk+1:=xk+v, we observe thatxk,xk+1 satisfy the first inequality in (2.4). It remains to show that

yk+1−A

xk+1< 1

2k+1. (2.9)

To furnish this, note that

f (x¯+tv)−f (x)¯

t −A(v)

< 1

2k+2 (2.10)

by (2.5). Thus one has the estimate

f (¯x+tu)−f (x)¯

t −A(v)

< 1

2k+2 (2.11)

due to ¯x+tv=xandf (x)=f (x¯+tu). Combining the latter inequality with that in (2.6), we arrive at

A(u)−A(v)< 1

2k+1. (2.12)

Then the required estimate (2.9) follows from the observation thatA(u)−A(v)=yk+1 A(xk+1), which justifies (2.4).

It is clear from (2.4) that{xk}is a Cauchy sequence inX, and hence it converges to some pointx∈X. On the other hand, we have from (2.4) that limA(xk)=limyk=y0. ThusA(x) =y0, that is,y0∈A(X). By the choice ofy0, we finally conclude thatA(X) is closed inY, which completes the proof of the theorem.

Remark 2.3. It is easy to observe from the proof of Theorem 2.2 that the same conclusion holds true if the RMR property off around ¯xis relaxed as follows:there are a neighborhoodV off (x)¯ and a constantµ >0such that, given anyy∈V∩f (X), there existsx∈f−1(y)satisfying x−x¯< µy−f (x)¯ . Also we do not need to require the completeness of the normed spaceY. Indeed, a slight modification of the proof allows us to show thatA(X)is complete in this case, which is all we need.

(6)

To continue, for the given mappingf:X→Y, we define theLipschitzian modulusof f in the ball ¯x+δBXby

f(x;¯δ):= sup

x1,x2x+δB¯ X x1=x2

f x1

−f x2

x1−x2 (2.13)

and observe thatf is Lipschitz continuous around ¯x if and only if f(¯x;δ) <∞for someδ >0. Given another mappingg:X→Y, we denote for simplicityf ,g(¯x;δ):= f−g(¯x;δ). In this notation, a mappingf:X→Y isstrictly differentiableat ¯x if there is a bounded linear operator∇f (¯x):X→Y with f ,∇f (¯x)(¯x;δ)→0 asδ↓0. For con- venience, we denoterf(¯x;δ):=f ,f (¯x)(x;¯δ)and call the functionrf(¯x;·):(0,∞)→ (0,∞]therate of strict differentiabilityoff at ¯x.

Theorem2.4. LetXandY be Banach spaces, and letA:X→Y be a bounded linear operator such that the spaceA(X)is closed and complemented in Y. Then there exist positive constantsγandµwith the following properties.

Givenf:X→Y withf ,A(x;¯δ) < γ for some δ >0, there are neighborhoods U of

¯

x andV off (x)¯ such that for anyx∈U andy∈V, there is xy ∈X satisfying the estimates

y−f

xy≤µdist y−f

xy

;A(X)

, x−xy≤µf (x)−y. (2.14) To be precise,xycan be chosen so that the first estimate in (2.14) is replaced with

π y−f

xy

=0 (2.15)

for any given projectionπfromY toA(X).

Proof. SinceA(X)is closed and complemented in Y, there is a closed subspace Y1⊂Y such thatY =Y1⊕A(X). Picking anyy∈Y, we uniquely represent it asy= y1+y2with somey1∈Y1,y2∈A(X)and define projectionsπi:Y→Y byπi(y)=yi, i=1,2. It is well known that the norm

y1:1(y)+π2(y) (2.16) is equivalent to the original one. This gives us a constantµ1>0 withy1≤µ1yfor ally∈Y. Thus

π1(y)=π1(y−y)˜ ≤π1(y−y)˜ 2(y−y)˜

= y−y˜1≤µ1y−y˜ ∀y∈Y ,y˜∈A(X). (2.17) Applying the classical open mapping theorem to the operatorA:X→A(X), we find µ2>0 such that for anyy∈A(X), there isx∈A1(y)withx ≤µ2y. Now denote

γ:= 1 2µ1µ2+2

, µ:=max

µ1,2µ1µ2

(2.18) and show that these are the constants we are looking for.

(7)

It is clear that everyfsatisfying the assumptions of the theorem is Lipschitz contin- uous around ¯x. Define

V:=f (x)+¯ δ1µ2

B, U:=f1(V )∩ x¯

2

B

(2.19)

and fixx∈U,y∈V. Starting withx0:=x, we find x1∈Xsuch that A

x1−x0

2

˜ y−f

x0 , x1−x0≤µ2π2

y˜−f

x0≤µ1µ2y˜−f (x) ≤δ

4, (2.20)

which impliesx1∈x¯B. Next we iteratively define a sequence{xk} ⊂Xas follows.

Givenxk,xk+1, choosexk+2satisfying A

xk+2−xk+1

=A

xk+1−xk

−π2 f

xk+1

−f xk xk+2−xk+1≤µ2A ,

xk+1−xk

−π2

f xk+1

−f

xk. (2.21) We proceed to show thatxk∈x¯Bby induction. Assumex0, . . . , xk+1∈x¯Band get

xk+2−xk+1≤µ2A

xk+1−xk

−π2

f xk+1

−f xk

2 f

xk+1

−f xk

−A

xk+1−xk 1

f xk+1

−f xk

≤µ2f xk+1

−f xk

−A

xk+1−xk1

f xk+1

−f xk

≤µ2

f ,A(¯x;δ)xk+1−xk1f xk+1

−f xk

−A

xk+1−xk

≤µ2

11

f ,A(¯x;δ)xk+1−xk<1

2xk+1−xk,

(2.22) which clearly implies that

xk+2−xk+12−k−1x1−x0, (2.23) xk+2−x0

k+1 i=0

xi+1−xi

k+1 i=0

2ix1−x0<2x1−x0≤δ

2. (2.24) Hencexk+2∈x¯B, and (2.23) holds for allk≥0 by induction. The latter implies that for anym∈N, one has

xk+m−xk

m1 i=0

xk+i+1−xk+i

m1 i=0

2kix1−x0 →0 ask→ ∞. (2.25) Therefore,{xk}is a Cauchy sequence that converges to some point ˜x∈X. By (2.24), we observe that

x˜−x ≤2x1−x01µ2y˜−f (x) ≤µy˜−f (x) , (2.26)

(8)

which ensures the second estimate in (2.14). It remains to show that ˜y−f (x)˜ ∈Y1

which gives (2.15) and hence the first estimate in (2.14) by (2.17). It follows from (2.20) and (2.21) that

A

xk+2−xk+1

= −π2 f

xk+1

−y˜

∀k∈N. (2.27)

Passing there to the limit ask→ ∞, we getπ2(f (x)˜ −y)˜ =0 and complete the proof.

The proof ofTheorem 2.4is based on a modified iteration procedure that goes back to the original proofs of Lyusternik and Graves. The above result, in contrast to other generalizations of the Lyusternik-Graves theorem, does not require the surjectivity as- sumption on underlying linear operatorA, while it requires thatA(X)be complemented inY. The latter condition is essential for property (2.15). We strongly believe that it can be relaxed to establish the metric regularity estimates (2.14).

WhenAissurjective,Theorem 2.4implies the following well-known result that was mainly obtained by Ioffe and Tihomirov [12] with a different proof; see also [6] and the references therein.

Corollary2.5. LetA:X→Y be a surjective bounded linear operator between Ba- nach spaces. Then there isγ >0such that every mappingf:X→Y withf ,A(¯x;δ) < γ for someδ >0 is metrically regular aroundx¯ and the constantµ >0in the metric regularity property can be chosen independent off.

WhenAisinjectiveinTheorem 2.4, we have the next corollary ensuring theuniform RMR property off around ¯x.

Corollary2.6. LetA:X→Y be an injective bounded linear operator between Ba- nach spaces. Assume thatA(X)is closed and complemented inY. Then there isγ >0 such that, for every mappingf :X →Y satisfying f ,A(¯x;δ) < γ with some number δ >0, the localized mappingf:[x¯+δB]→f (x¯+δB)is metrically regular aroundx¯ with a constantµ >0independent off.

Proof. Consider the mapping ˜f := (π◦f ):X →A(X), where π is a projection fromY toA(X). UsingCorollary 2.5, we conclude that ˜f is open around ¯x. Since the linear operator A :X →A(X) is surjective and injective, its inverse operator A−1 : A(X)→X is single-valued and bounded. We show that ˜f is injective on ¯x+δBwhen γ <π1A11. Indeed, if ˜f (x1)=f (x˜ 2)forx1, x2∈x¯Bwithx1=x2, then

f x1

−f x2

−A

x1−x2 x1−x2

≥π1f˜ x1

−f˜ x2

−A

x1−x2 x1−x2

−1A

x1−x2

x1−x2 ≥ π1A1−1,

(2.28)

which contradicts the choice ofγ. Then we applyTheorem 2.4to get the conclusion.

(9)

For further results, we need to introduce another modulus involving the mapping f:X→Y and a linear operatorA:X→Y between Banach spaces:

ϑf ,A(¯x;δ):= sup

y1,y2f (X)(f (¯x)+δB) y1=y2

dist

y2−y1;A(X)

y2−y1 . (2.29)

The following useful relationship can be easily derived from the definitions.

Proposition 2.7. Taking a mapping f :X→Y and a linear operator A:X→Y between Banach spaces, assume thatfis RMR aroundx. Then there are positive numbers¯ µ >0andδ0such that

ϑf ,A(x;¯δ)≤µf ,A(¯x;µδ) ∀δ < δ0. (2.30)

The next result shows that, roughly speaking,f:X→f (X)⊂Y is RMR around ¯xif and only if f (X)islocally homeomorphic toA(X)for some linear bounded operator A:X→Y close tof, that is, the modulusf ,Ais sufficiently small.

Proposition2.8. LetA:X→Ybe a linear bounded operator between Banach spaces such thatA(X)is closed and complemented inY, and letπ be a projection fromY to A(X). Then the following hold.

(a)There isγ >0such that, for every mappingf:X→Y satisfying the RMR property aroundx¯and the estimatef ,A(x;¯δ) < γwith someδ >0, the projectionπ is a local homeomorphism betweenf (X)andA(X)aroundf (x).¯

(b)There isγ >0such that every mappingf:X→Y satisfyingf ,A(¯x;δ) < γ with someδ >0is RMR aroundx¯provided that the projectionπ is a local homeomorphism betweenf (X)andA(X)aroundf (¯x).

Proof. It follows fromTheorem 2.4andProposition 2.7.

Corollary2.9. Letf:X→Y be a mapping between Banach spaces that is strictly differentiable atx. Assume that¯ ∇f (x)(X)¯ is closed and complemented inY. Thenf is RMR aroundx¯if and only if each projection fromY to∇f (¯x)(X)is a local homeomor- phism betweenf (X)and∇f (x)(X).¯

The next result givesnecessary and sufficient conditions for RMR of strictly differ- entiable mappings. In its formulation, we use, beside the modulusϑfrom (2.29), the following construction of thetangent coneto a setΩ⊂Xat ¯x∈Ω:

T (¯x;):=

v∈X| ∃vkv, tk0, xk →x¯withxk+tkvk

, (2.31)

wherexk

→x¯means thatxk→x¯withxkΩask→ ∞. Note that this tangent cone is an enlargement of the well-known (Bouligand) contingent cone corresponding to (2.29) withxk=x¯for allk∈N.

(10)

Theorem2.10. Letf:X→Y be a mapping between Banach spaces that is strictly differentiable atx. Consider the following conditions:¯

(a) fis RMR aroundx;¯

(b) ϑf ,f (¯x)(x;¯δ)→0asδ↓0and∇f (x)(X)¯ is closed;

(c) T (f (x);¯ f (X))= ∇f (x)(X).¯

Then (a)⇒(b)⇒(c). Moreover, (b)⇒(a) if∇f (¯x)(X)is closed and complemented inY. Also (c)⇒(a) whencodim∇f (x)(X) <¯ ∞.

Proof. Implication (a)⇒(b) is straightforward, while (b)⇒(a) follows fromProposi- tion 2.8under the assumptions made. To prove (b)(c), we first observe that∇f (x)(X)¯

⊂T (f (x);¯ f (X))by the strict differentiability off at ¯x. Now picku∈T (f (x);¯ f (X)) and find by (2.31) sequencesuk→u,yk→f (x), and¯ tk0 such thatyk+tkuk∈f (X) for allk∈N. Property (b) ensures the existence of ˜yk∈ ∇f (x)(X)¯ withuk−y˜k/tk0 ask→ ∞. This gives ˜yk/tk→uask→ ∞and henceu∈ ∇f (x)(X), which yields (c).¯

Finally, we prove (c)⇒(a) assuming that the space∇f (x)(X)¯ is finite codimensional.

Thus there is a finite-dimensional spaceY1⊂Y withY1⊕ ∇f (¯x)(X)=Y. ByTheorem 2.4, we find neighborhoodsUof ¯xandV off (¯x)as well as a numberµ >0 such that, for anyx∈Uandy∈V, there existsxy∈Xsatisfying

y−f xy

∈Y1, x−xy≤µf (x)−y. (2.32)

To justify (a), one needs to show thaty=f (xy)provided thaty∈f (X)and thatU and V are sufficiently small. Assuming the contrary, we findxk→x¯and yk→f (x)¯ with yk∈f (X)such that 0=yk−f (xk)∈Y1. Denotetk:= yk−f (xk)and uk:= (yk−f (xk))/tk. SinceY1is finite dimensional,ukconverges to some 0=u∈Y1along a subsequence. On the other hand,u∈T (f (x);¯ f (X))= ∇f (x)(X)¯ by (c) and the con- struction ofuk. Thusu=0, which is a contradiction. This completes the proof of (c)(a) and of the whole theorem.

3. First-order calculus. In this section, we give applications of the RMR property to first-order calculus rules for sequential limiting generalized differential constructions in arbitrary Banach spaces. Other applications of metric regularity and related proper- ties to generalized differential calculus can be found in [11,13,15], and the references therein. The results presented below seem to be new even in the case of applications of the classical (not restrictive) metric regularity property for mappings between general Banach spaces.

First we define the generalized differential constructions of our study; see [19,25]

and their bibliographies for the history of these constructions and more discussions.

GivenΩ⊂Xand ¯x∈Ω, we define theset ofε-normalstoΩat ¯xby

Nε(x;¯Ω):=



x∈X|lim sup

xx¯

x, x−x¯ x−x¯ ≤ε



, ε≥0. (3.1)

(11)

Whenε=0, this is a cone called theprenormal coneor theFréchet normal coneand denoted byN( x;¯Ω). Then thebasic/limiting normal conetoΩat ¯xis given by

N(x;¯Ω):=

x∈X| ∃εk0, xk w x, xk →x¯withxk∈Nεk

xk;Ω

, (3.2) where the limits in (3.2) aresequential. When X is Asplund andΩ is closed around

¯

x, one can equivalently putεk=0 in (3.2) and the subsequent limiting constructions;

see [25]. However,εkcannot be removed from the definitions without loss of crucial properties in general Banach space settings, as one can see in the arguments and results below.

A set Ω⊂X is callednormally regular at ¯x∈Ωif N( x;¯Ω)=N(x;¯Ω). This class includes, in particular, all convex sets, sets with smooth boundaries, and so forth. New calculus results for normal regularity are obtained in this section being incorporated in calculus rules for normal cones.

Given a set-valued mapping F : XY between Banach spaces, we define its ε- coderivative, normal coderivative, andmixed coderivativeat(¯x,y)¯ gphF by, respec- tively,

DεF (¯x,y)¯ y

:=

x∈X|

x,−y

∈Nε

(¯x,y); gph¯ F

, (3.3)

DNF (¯x,y)¯ y

:=

x∈X|

x,−y

∈N

(¯x,y); gphF¯

, (3.4)

DMF (¯x,y)¯ y

:=

x∈X∃εk0, xk, yk

F

→(¯x,y),¯ xk w x, yk ·→y,withxk∈Dε

kF xk, yk

yk ,

(3.5)

for ally∈Y, where(xk, yk)→F (x,¯y)¯ means that(xk, yk)→(¯x,y)¯ with(xk, yk)∈ gphF. We say thatFisN-regular (resp.,M-regular) at(¯x,y)¯ ifDF (¯x,y)¯ =DNF (x,¯y)¯ (resp.,DF (x,¯y)¯ =DMF (¯x,y)).¯

Given an extended real-valued functionϕ:X→R:=[−∞,∞]finite at ¯x, we define its (first-order)ε-subdifferentialandbasic subdifferentialby, respectively,

εϕ(¯x):=DεEϕ

¯ x, ϕ(¯x)

(1), ∂ϕ(¯x):=DEϕ

¯ x, ϕ(¯x)

(1), (3.6) whereEϕ :X→R is the epigraphical multifunction with gphEϕ =epiϕ and where DEϕ :=DNEϕ=DMEϕ, since there is no difference between the normal and mixed coderivatives for mappings with values in finite-dimensional spaces. A functionϕ is said to belower regular at ¯xif∂ϕ(¯ x)=∂ϕ(¯x).

Note that the above subdifferential constructions admit intrinsic analytic representa- tions not directly involving generalized normals; see [25] and its references. The given definitions allow us to develop ageometric approachto generalized differential calculus based just on calculus rules for generalized normals.

First we obtain two-sideduniform estimatesofε-normals forinverse imagesof sets f1():=

x∈X|f (x)∈

(3.7) under mappingsf:X→Y having the RMR property.

(12)

Theorem3.1. Letf:X→Y be a mapping between Banach spaces that is Lipschitz continuous around some pointx, let¯ A:X→Y be a linear bounded operator, and letbe a subset ofY withy¯:=f (x)¯ . Then there are positive constantsδ¯andµ1such that for everyε≥0,δ∈(0,δ),¯ x∈f−1()∩(x¯B), andy∈Nε(f (x);∩f (X)),

Ay∈Nε1

x; f−1()

ε1:1ε+yf ,A(¯x;δ). (3.8) If, in addition,f is RMR aroundx¯andA(X)is closed inY, then there isµ2>0such that for everyx∈Nε(x;f −1()),x∈x+(ε+µ2(ε+x)f ,A(x;¯δ))B satisfying (A)1(x)= ∅and

y∈Nε2

f (x); Ω∩f (X)

ε2:2ε+µ2x+yf ,A(¯x;δ) (3.9)

whenevery∈(A)1(x).

Proof. Without loss of generality, we assume thatΩ⊂f (X)throughout the proof.

Fixxas in the theorem and puty:=f (x), y∈Nε(y;). Choose ¯δsuch thatf(¯x; ¯δ) <

. Then we have

lim sup

xf−1(Ω)x

Ay, x−x

x−x = lim sup

xf−1(Ω)x

y, A(x−x) x−x

lim sup

xf

−1(Ω)

x

y, f (x)−f (x)

x−x +yf ,A(¯x;δ)

lim sup

yy max

0,

y, y−y f(¯x; ¯δ)1y−y

+yf ,A(¯x;δ)

f(x; ¯¯δ)ε+yf ,A(x;¯δ),

(3.10) which implies (3.8) withµ1:=f(¯x; ¯δ)by definition (3.1).

The proof of (3.9) is more involved. First observe that

f (x+tv)−y≤f ,A(x;¯δ)vt (3.11) for allv∈kerA and smallt >0. By the RMR property of f around ¯x, we findxt f1(y) such thatx+tv−xt ≤µf ,A(¯x;δ)vtfor some constantµ >0 and for each smallt >0. Then for anyγ >0, one has

x, xt−x

≤(ε+γ)xt−x≤(ε+γ)

1+µf ,A(¯x;δ)

vt (3.12)

whentis close to zero. On the other hand, x, xt−x

= x, tv

+

x, xt−x−tv

≥t x, v

−µxf ,A(x;¯δ)vt. (3.13)

(13)

Thus we have the estimate x, v

≤µxf ,A(¯x;δ)v+(ε+γ)

1f ,A(¯x;δ)

v (3.14)

for anyγ >0, which implies that

x, v

≤εv (3.15)

wheneverv∈kerA, with

ε:=ε+µ

ε+xf ,A(¯x;δ). (3.16) The Hahn-Banach theorem allows us to extendx|kerAto some ˜x∈Xwithx˜ ≤ε.

Now we letx:=x−x˜and construct a linear functionalyonA(X)by y, y

:= x, x

for anyy∈A(X), x∈A1(y). (3.17) Since kerAker(x), this functional is well defined. The RMR property ofA:X→A(X) (which is automatic due to the closedness ofA(X)) implies thatyis bounded. Using the Hahn-Banach theorem again, we extendyto some functional ˜y∈Y. One clearly hasx=Ay˜, and hence(A)−1(x)= ∅.

Take an arbitrary functionaly∈(A)−1(x). It remains to show thaty∈Nε2(y;

), whereε2is defined in (3.9) with someµ2≥µ. To proceed, we use the assumed RMR property off around ¯x and for anyy Ωclose toy, find xy ∈f−1(y)such thatxy−x ≤µy−ywith some constantµ > 0. Therefore,xy→xwheny→y.

Furthermore,

y−y−A

xy−x≤f ,A(x;¯δ)xy−x (3.18) wheneveryis close toy. Thus we have the estimates

lim sup

yy

y, y−y y−y

lim sup

y y

y, A

xy−x

y−y +f ,A(¯x;δ)xy−x y−y y

lim sup

y y

x−x˜, xy−x

y−y f ,A(x;¯δ)y

lim sup

y y

x, xy−x

y−y +x˜·xy−x y−y

f ,A(¯x;δ)y

lim sup

y y max

0,

x, xy−x

µ−1xy−x

+µε+µ f ,A(¯x;δ)y

≤µε +µε+µ f ,A(¯x;δ)y≤µ2ε+µ2x+yf ,A(x;¯δ)=ε2,

(3.19)

whereµ2:=max{µ,µ+µµ f ,A(x; ¯¯δ),µµ, µ} ≥µ. This completes the proof.

(14)

Remark3.2. (i) Whenf isstrictly differentiableat ¯x, we can takeA:= ∇f (x)¯ with f ,A(¯x;δ)replaced by the rate of strict differentiabilityrf(x;¯δ)inTheorem 3.1. Note that the subspace∇f (x)(X)¯ is closed inY byTheorem 2.2. In this case, necessary and sufficient conditions for the RMR property offaround ¯xare given inTheorem 2.10. If, in addition,∇f (x)¯ issurjective, thenfis metrically regular around ¯xby the Lyusternik- Graves theorem and (3.9) reduces to

x∈ ∇f (x)¯ Nε2

f (x);+

ε+µ2

ε+xrf(x;¯η)

B (3.20)

for anyx∈Nε(x; f1()), whereε2:2ε+µ2xrf(x;¯δ). In particular, whenxis replaced by ¯x, the following holds:

Nε

¯

x;f−1()

⊂ ∇f (x)¯ Nµ2ε

f (x);¯ Ω

B ∀ε≥0. (3.21) (ii) It is important to put Ω∩f (X) in (3.9) but not Ω. To illustrate this, consider f:R2R2andΩR2defined by

f (u, v):=(u,0) ∀(u, v)∈R2, Ω:=

(u, v)∈R2with|u| ≥ |v|

. (3.22) Then (3.9) fails for ¯x=x=0 ifΩ∩f (X)is replaced byΩ.

(iii) Note that the uniform estimates inTheorem 3.1are distinguished from the re- sults of “fuzzy calculus” type available under other assumptions for such constructions in Asplund spaces; see, for example, [11,25] with their references. The main differences are that we getuniform qualitativeestimates ofε-normals forallpoints around the ref- erence ones, while fuzzy calculus results involve only some of them. These advantages ofTheorem 3.1are used in what follows.

Theorem 3.1 directly implies the following two calculus rules ofequality typefor Fréchet normals to inverse images.

Corollary3.3. Letf:X→Y be a mapping between Banach spaces that is strictly differentiable atx¯and such thatf (x)¯ . Assume thatf is RMR aroundx. Then¯

N

¯

x;f1()

= ∇f (x)¯ N

f (x);¯ Ω∩f (X) , ∇f (x)¯ −1N

¯

x;f−1()

=N

f (x);¯ Ω∩f (X)

. (3.23)

Proof. Both equalities follow fromTheorem 3.1, withA= ∇f (x),¯ ε=0,x=x, and¯ δ↓0.

Note that the first equality implies the second one inCorollary 3.3when∇f (¯x)is surjective, that is, whenfis metrically regular around ¯x. In general, they are indepen- dent as can be easily illustrated by simple examples.

Next, we intend to derive exact formulas for computing basic normals (3.2) to inverse images by passing to the limit from the estimates of Theorem 3.1. To proceed, we need to introduce first the followingweakextensibility property, which is related but somewhat different from the Banach extensibility property (see, e.g., [5]) and plays an essential role in the subsequent results of this paper.

(15)

Definition3.4. LetLbe a closed linear subspace of a Banach spaceX.Lis w- extensibleinXif every sequence{vk} ⊂L, withvk w 0 ask→ ∞, contains a subse- quence{vk

j}such that eachvk

j can be extended toxj∈Xwithxj w 0 asj→ ∞. The next proposition shows that the w-extensibility property always holds for complemented subspaces of arbitrary Banach spaces and also, unconditionally, in a broad class of Banach spaces including all Asplund spaces, weakly compactly gener- ated spaces, spaces admitting smooth renorms of any kind, and so forth.

Proposition3.5. LetLbe a closed linear subspace of a Banach spaceX. ThenLis w-extensible inXif one of the following conditions holds:

(a) Lis complemented inX;

(b) the closed unit ball ofXis weaksequentially compact.

Proof. LetLbe complemented in X, and letπ:X→L be a projection operator.

Puttingxk:= vk, π (x)onX, we conclude thatxkis an extension ofvkwithxk w 0, that is,Lisw-extensible inXin case (a).

To justify this property in case (b) for everyL⊂X, we take an arbitrary sequencevk fromDefinition 3.4and observe that it is bounded inLdue to the weakconvergence.

By the Hahn-Banach theorem, we extend eachvk to ˜xk ∈Xsuch that the sequence {x˜k}is still bounded inX. SinceBX is assumed to be weaksequentially compact, there arex∈Xand a weakconvergent subsequence ˜xk

j w

→x asj→ ∞. Putting xj:=x˜k

j−x, we complete the proof of the proposition.

We show that the w-extensibility property may not hold even in some classical Banach spaces.

Example3.6. The subspaceL=c0is notw-extensible inX=.

Proof. Recall thatc0 is a Banach space of all real sequences convergent to zero that is endowed with the supremum norm. Letvk:k∈c0, whereξk maps every vector fromc0to itskth component. Assume that there is an increasing sequence of kjNsuch thatvk

j can be extended toxj∈() with xj w 0. Define a closed linear subspace ofby

Z:=

α1, α2, . . .

k=0 ifk

k1, k2, . . .

(3.24) and a linear bounded operatorA:→Zby

A

α1, α2, . . . :=

β1, β2, . . .

α1, α2, . . .

, (3.25)

where one has

βk=



αj ifk=kj, j∈N,

0 otherwise. (3.26)

参照

関連したドキュメント

By using some generalized Riemann integrals instead of ordinary sums and multiplication systems of Banach spaces instead of Banach spaces, we establish some natural generalizations

Interesting results were obtained in Lie group invariance of generalized functions [8, 31, 46, 48], nonlinear hyperbolic equations with generalized function data [7, 39, 40, 42, 45,

proof of uniqueness divides itself into two parts, the first of which is the determination of a limit solution whose integral difference from both given solutions may be estimated

[2] Agmon S., Douglis A., Nirenberg L., Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions, I, Comm..

Arnold This paper deals with recent applications of fractional calculus to dynamical sys- tems in control theory, electrical circuits with fractance, generalized voltage di-

Arnold This paper deals with recent applications of fractional calculus to dynamical sys- tems in control theory, electrical circuits with fractance, generalized voltage di-

[2] , Metric and generalized projection operators in Banach spaces: Properties and applica- tions, Theory and Applications of Nonlinear Operators of Accretive and Monotone Type

[2] , Metric and generalized projection operators in Banach spaces: Properties and applica- tions, Theory and Applications of Nonlinear Operators of Accretive and Monotone Type