• 検索結果がありません。

A DEGREE THEORY FOR COMPACT PERTURBATIONS OF PROPER C1 FREDHOLM MAPPINGS OF INDEX

N/A
N/A
Protected

Academic year: 2022

シェア "A DEGREE THEORY FOR COMPACT PERTURBATIONS OF PROPER C1 FREDHOLM MAPPINGS OF INDEX"

Copied!
25
0
0

読み込み中.... (全文を見る)

全文

(1)

PROPER C

1

FREDHOLM MAPPINGS OF INDEX 0

PATRICK J. RABIER AND MARY F. SALTER Received 14 June 2004

We construct a degree for mappings of the formF+K between Banach spaces, where FisC1Fredholm of index 0 andK is compact. This degree generalizes both the Leray- Schauder degree whenF=Iand the degree forC1Fredholm mappings of index 0 when K=0. To exemplify the use of this degree, we prove the “invariance-of-domain” prop- erty whenF+K is one-to-one and a generalization of Rabinowitz’s global bifurcation theorem for equationsF(λ,x) +K(λ,x)=0.

1. Introduction

We generalize the Leray-Schauder degree to mappingsF+Kbetween real Banach spaces X andY, whereF isC1 Fredholm of index 0 andK is compact. Recall that the Leray- Schauder theory addresses the case whenX=Y andF(x)=x. Throughout this paper, a (nonlinear) compact operator is a continuous operator mapping bounded subsets to relatively compact ones.

Under rather restrictive additional assumptions, it is sometimes possible to reduce a problem involving an operatorF+K as above to one having the desired structure in the Leray-Schauder theory. For instance, ifFis a homeomorphism ofXontoYandF1 maps bounded subsets to bounded subsets, the equationF(x) +K(x)=zbecomesy+K F1(y)=zandKF1is compact. More generally, this reduction is possible if the above assumptions are made aboutF+N, whereNC1(X,Y) is compact (just writeF+K= (F+N) + (KN)). However, the existence of such an operatorNis not guaranteed when Fis nonlinear.

The possible nonexistence of an equivalent Leray-Schauder formulation is already an issue whenK=0 and the question is to define a degree forC1 Fredholm mappingsF of index 0. The first investigations by Caccioppoli [5,6] in this direction, resulting in a mod 2 degree, go back to 1936, only two years after the work of Leray and Schauder. In the C2case, the mod 2 degree was subsequently rediscovered and more rigorously justified by Smale [26].

The first instance of aZ-valued degree forC2 Fredholm mappings of index 0 can be found in [8,9] by Elworthy and Tromba. Although its definition via Fredholm structures

Copyright©2005 Hindawi Publishing Corporation Abstract and Applied Analysis 2005:7 (2005) 707–731 DOI:10.1155/AAA.2005.707

(2)

made this degree of limited practical value, this work revealed that an integer-valued degree theory for Fredholm mappings of index 0cannotcomply with the homotopy in- variance property and, more specifically, that the degreemust be allowed to change sign under some Fredholm homotopies. An attempt to combine the Elworthy-Tromba ap- proach with the ideas of Caccioppoli to handle theC1 case can be found in the work of Borisovich et al. [4], where compact perturbations are also discussed. A theory for the nonnegative index case, in which the degree is defined as a framed cobordism, was further developed in [28] by Zvyagin and [29] by Zvyagin and Ratiner.

The construction of a “user-friendly” degree was described by Fitzpatrick et al. [11], based on the concept of parity of a path of linear Fredholm operators of index 0 (see Section 2), which also makes it possible to assess whether a sign change occurs during homotopy. The technical difficulties to extend this construction to theC1case were re- solved in [17,18] by Pejsachowicz and Rabier. An approach technically simpler in places and in a setting where homotopy invariance holds was proposed by Benevieri and Furi [1,2]. Several applications to existence or bifurcation questions in ODEs or PDEs on unbounded domains (which, typically, give rise to operators beyond the Leray-Schauder theory) have been described in [14,15,20,21,25], among others. Further comments on the development of the degree theory for Fredholm mappings and additional references, notably to the work of the Russian school, can be found in [12].

The starting point of this paper is the degree described in [18] forC1Fredholm map- pings of index 0. A concise but complete review of its definition and properties is given in Section 2. A special feature is that the numerical value of the degree depends upon the choice of abase-pointof the mapping Fof interest, which is simply a point pX such thatDF(p) is invertible (in these introductory comments, there is no need to con- sider the case when no base-point exists). Then, ifΩXis an open subset such thatFis proper onΩ, the base-point degreedp(F,Ω,y)Zis defined for ally /F(∂Ω). Different choices of the base-point lead to the same degree up to sign and the possible sign change is characterized as a parity.

If nowF above is replaced byF+K withKC0(X,Y) compact andΩis bounded, the general idea is to replaceKby a suitableC1approximationKε, so thatF+Kε isC1 Fredholm of index 0 and proper onΩ. As a result, a base-point degree is available for F+Kε which, in principle, can be used to define the desired degree for F+K. How- ever, the justification of this natural approach is faced with various nontrivial difficul- ties.

The first problem is of course the veryexistenceof such “suitable” approximationsKε

ofKof classC1, which turns out to be a delicate technical issue. When the dual spaceXis separable and henceXis separable and has an equivalent norm of classC1away from the origin (Restrepo [23]), the problem can be greatly simplified thanks to an approximation theorem of Bonic and Frampton [3]. Details can be found in the dissertation [24]. Our strategy when noC1 norm is available consists in first considering the case whenK is finite dimensional. TheC1approximation remains delicate and cannot be established on all ofΩ, but only on well-chosen compact subsets.

The approximation ofKbyC1mappingsKεis not the only difficulty, even whenK is finite dimensional: if the base-point degree forF+Kε is to be used to define the degree

(3)

forF+K, it must naturally be shown that it is independent of the approximationKε. But dp(F+Kε,Ω,y) makes sense only ifpis a base-point ofF+Kε, andpneed not remain a base-point afterKεis changed. This makes it impossible to compare the degreesdp(F+ Kε,Ω,y) for all the possible approximationsKε. A proof of the independence ofdp(F+ Kε,Ω,y) withpdepending uponKεis even more hopeless since, even whenKε is fixed, different choices ofpmay yield degrees of opposite signs.

The way to overcome this ambiguity is to choosepabove to be a base-point ofFin the first place and to confine attention to approximationsKε “based atp”, that is, satisfying DKε(p)=0. Thus,pis a base-point ofF+Kεfor all the approximationsKεof interest and it now makes sense to ask whetherdp(F+Kε,Ω,y) is independent of the approximation Kεbased atp. This happens to be true, along with the existence of such approximations.

This yields a reasonable definition for a degreedp(F+K,Ω,y) depending upon the base-pointpofF(not ofF+K, since this makes no sense ifKis not differentiable). The next obvious question is whether this degree is independent of the decompositionF+K:

ifT=F+K=G+LwithF andG Fredholm of index 0 andK andLcompact, and if pis a base-point ofF andofG, does the above construction provide the same value for dp(T,Ω,y) whenT=F+K orT=G+L? (This is not an issue with the Leray-Schauder degree sinceF=G=IandK=Lin that theory.) The answer isnegative, but this does not induce any new complication, as we now explain.

First, this negative answer shows that the degree forF+Kconstructed above depends not only upon the base-point pofFbut also upon the decomposition, that is, uponF.

As a result,dp(F+K,Ω,y) is not a correct notation for the degree anddF,p(F+K,Ω,y) should be used instead. At a first sight, this opens up the possibility that a plethora of inte- gers might represent the degree of the same mapping, depending upon which decompo- sition is chosen. But the reality is much simpler: two decompositionsT=F+K=G+L and two base-points p and q of F andG, respectively, can only lead to the same de- gree (dF,p(T,Ω,y)=dG,q(T,Ω,y)) or to two degrees of opposite signs (dF,p(T,Ω,y)=

dG,q(T,Ω,y)). Furthermore, the possible change of sign can once again be character- ized as a parity.

The degreedF,p(F+K,Ω,y) satisfies all the properties expected of a topological degree, except (just as when K=0) that it may change its sign under homotopies. However, special homotopies, notably those affecting onlyK, leave the degree invariant. WhenF is alocal diffeomorphism,dF,p(F+K,Ω,y) is actually independent ofpXand can then be denoted bydF(F+K,Ω,y). In particular, whenX=Y andF=I, we obtain a degree dI(I+K,Ω,y) for the compact perturbations of the identity. Not surprisingly, this degree coincides with the Leray-Schauder degree, which thus appears as a genuine special case ofdF,p(F+K,Ω,y).

To demonstrate the fact that, in its general form, the degree of this paper is as versatile as the Leray-Schauder degree, we use it inSection 8to prove the “invariance-of-domain”

property and inSection 9 to generalize the well-known Rabinowitz global bifurcation theorem to problems of the formF(λ,x) +K(λ,x)=0.

Throughout the paper,ᏸ(X,Y) is the space of bounded linear operators fromXtoY, GL(X,Y) the subset of linear isomorphisms,Φn(X,Y) the subset of Fredholm operators of indexnZ, and᏷(X,Y) the subspace of compact operators. FormNandnZ, we

(4)

letΦnCm(X,Y) denote the set ofCm Fredholm mappings of indexnfromX toY. As is customary, whenX=Y, we useᏸ(X),GL(X),Φn(X),᏷(X), andΦnCm(X), respectively.

2. Background

We briefly review the definition and properties of the base-point degree for properC1 Fredholm mappings of index 0. Details can be found in [18] and the references therein.

A basic concept is that ofparityof a continuous path of linear Fredholm operators of index 0. LetXandY be real Banach spaces and, given a compact interval [a,b]R, letAC0([a,b],Φ0(X,Y)) be a path of linear Fredholm operators of index 0 such that A(a),A(b)GL(X,Y). As shown in [10, Theorem 1.3.6] or [27, Theorem 2.10], there is PC0([a,b],GL(Y,X)) (parametrix) such thatPA=ICwhereCC0([a,b],᏷(X)).

Then,IC(a)GL(X) andIC(b)GL(X) have well-defined indices (=(1)where is the number of negative eigenvalues, counted with multiplicity)i(IC(a))∈ {−1, 1} andi(IC(b))∈ {−1, 1}and the parityσ(A) ofA(orσ(A, [a,b]) if displaying the inter- val is important) is defined by

σ(A) :=iIC(a)iIC(b)∈ {−1, 1}. (2.1) Of course, it can be shown that this formula is independent of the parametrixP.

To understand the meaning ofσ(A), it is helpful to consider the caseX=Y=RN when, as is easily seen,σ(A)=sgn detA(a) sgn detA(b). In the infinite-dimensional set- ting,σ(A) may be thought of as a generalization of this formula in the absence of any determinant function. However, in contrast to the finite-dimensional case,σ(A) neednot depend only upon the endpointsaandb(see [10]). Alternatively,σ(A) may be viewed (generically) as the mod 2 count of the number of timesA(t) crosses the subset of nonin- vertible linear Fredholm operators of index 0 astruns over [a,b].

The parity has a number of interesting properties, including homotopy invariance (provided that the endpoints remain invertible during the homotopy), multiplicativity with respect to consecutive intervals (i.e., σ(A, [a,c])=σ(A, [a,b])σ(A, [b,c]) if A(a), A(b), andA(c) are invertible), and multiplicativity with respect to composition. Also important in practice, the parity is unchanged by reparametrizations and the parity of a path of linear isomorphisms is always 1.

Let nowFΦ0C1(X,Y) be given and letΩX be an open subset. Assume thatF is proper onΩand let y /F(∂Ω) be a regular value ofF|, that is,DF(x)GL(X,Y) wheneverxΩandF(x)=y. Then, by properness, the setF1(y)=F1(y)Ωis finite, say

F1(y)=

x1,. . .,xk

, (2.2)

for some integerk0.

GivenpXsuch thatDF(p)GL(X,Y) (abase-pointofF), the degreedp(F,Ω,y) is defined by the sum of parities:

dp(F,Ω,y)= k i=1

σDFγi

, (2.3)

(5)

whereγiis any continuous curve inXjoiningptoxi, 1ik. Ifk=0 (so thatF1(y)=

), we setdp(F,Ω,y)=0. The homotopy invariance of the parity ensures that the above definition of dp(F,Ω,y) is independent of γi. It does, however, depend upon p, but passing from a base-point to another can only leave the degree unchanged or change it into its negative (seeCorollary 2.4). Thus, the “absolute” degree|d|defined by|d|(F,Ω, y)= |dp(F,Ω,y)|is independent ofp. This makes it possible to define|d|even when no base point exists, by setting|d|(F,Ω,y)=0 in this case.

Remark 2.1. It follows from the above definition that whenF is a linear isomorphism, thendp(F,Ω,y)=1 regardless ofpXwheneveryF(Ω). This is often useful in prac- tical calculations, together with homotopy arguments (see below). Thatdp(F,Ω,y) can never be 1 whenF is linear points to differences—not contradictions or incompati- bilities—between the base-point degree and the Leray-Schauder degree.

The most technical step consists in definingdp(F,Ω,y) wheny /F(∂Ω) is not nec- essarily a regular value ofF|. WhenFisC2, this is done by approximatingyby regular values (see [11]). WhenFis onlyC1, this approach fails and must be replaced by approx- imatingFrather thany(see [17,18]).

The properties of the base-point degree, listed below, are almost the expected ones, the notable exception being that it is only invariant up to sign under homotopies. However, the sign change (or lack thereof) can be fully monitored by the parity, as indicated in Theorem 2.2.

Theorem2.2. LethΦ1C1([0, 1]×X,Y)be proper on[0, 1]×and lety /h([0, 1]×

∂Ω)be given. The following properties hold.

(i)Ifp0Xandp1Xare base-points ofh(0,·)andh(1,·), respectively, then

dp0

h(0,·),Ω,y=νdp1

h(1,·),Ω,y, (2.4)

whereν:=σ(DxhΓ)∈ {−1, 1}andΓis any continuous curve in[0, 1]×Xjoining (0,p0)to(1,p1).

(ii)Ifp0Xis a base-point ofh(0,·)and ifh(1,·)1(y)= ∅, then

dp0

h(0,·),Ω,y=0. (2.5)

(iii)Ifp0Xis a base-point ofh(0,·)and ifh(1,·)has no base-point, then

dp0

h(0,·),Ω,y=0. (2.6)

The following corollary gives a simple but useful case when homotopy invariance holds.

(6)

Corollary2.3. LethΦ1C1([0, 1]×X,Y)be proper on[0, 1]×and lety /h([0, 1]×

Ω)be given. IfpXis a base-point ofh(t,·)for allt[0, 1](i.e.,Dxh(t,p)GL(X,Y) for allt[0, 1]), thendp(h(0,·),Ω,y)=dp(h(1,·),Ω,y).

Corollary2.4. LetFΦ0C1(X,Y)be proper onand letp,qXbe base-points ofF.

Ify /F(∂Ω), thendq(F,Ω,y)=νdp(F,Ω,y)whereν:=σ(DFγ)∈ {−1, 1}andγis any continuous curve inXjoiningptoq.

Corollary2.5 (local constancy). LetFΦ0C1(X,Y)be proper onand letpX be a base-point ofF. The degreedp(F,Ω,y)depends only upon the connected component of Y\F(∂Ω)containingy.

Corollary2.6 (normalization). LetFΦ0C1(X,Y)be proper onand letpX be a base-point ofF. Ify /F(∂Ω)anddp(F,Ω,y) =0, thenF1(y) = ∅.

The following two properties of the degree are also important for calculations.

Theorem2.7 (excision). LetFΦ0C1(X,Y)be proper onand letΣbe a closed subset ofΩ. Then,Fis proper on\Σ. Furthermore, ify /F(ΣΩ), theny /F(∂(Ω\Σ))and ifpXis a base-point ofF,dp(F,Ω\Σ,y)=dp(F,Ω,y).

Theorem2.8 (additivity on domain). Suppose that=12 where1 and2 are disjoint open subsets ofXand letFΦ0C1(X,Y)be proper onΩ. Then,Fis proper onα, α=1, 2. Furthermore, ify /F(∂Ω), theny /F(∂Ωα),α=1, 2, and ifpXis a base-point ofF,dp(F,Ω,y)=dp(F,Ω1,y) +dp(F,Ω2,y).

The absolute degree|d|ishomotopy invariant, which is consistent withTheorem 2.2 when base-points exist, but is true in general. Corollaries2.5and2.6as well asTheorem 2.7also hold for|d|.

All of the above can be repeated ifFis only defined on a connected and simply con- nected open subsetᏻofX withΩᏻ(it even suffices that the first cohomology group H1(ᏻ) vanishes; see [12]) and closures are understood relativeᏻ. It will be obvious from the proofs that the results of this paper can also be extended verbatim to this setting.

3. Degree for finite-dimensional perturbations: definition

In this section,FΦ0C1(X,Y) and the compactfinite dimensionalmappingKC0(X,Y) are given andΩdenotes aboundedopen subset ofX. We define a degree forF+K that generalizes the base-point degree ofSection 2whenK=0. After this degree is constructed and its main properties established, it will be a simple matter to drop the assumption that Kis finite dimensional (Section 7).

We will further assume thatFis proper onΩand that

F(Ω) is bounded, (3.1)

although (3.1) will be dropped inSection 7.

LetY0be a finite-dimensional subspace ofYsuch thatK(X)Y0. SinceF(Ω) is closed (by the properness ofF), it follows from (3.1) thatF(Ω)Y0is a compact subset ofY

(7)

and hence that

SF,Ω,Y0

:=F1F(Ω)Y0

Ω (3.2)

is a compact subset ofX. For example, ifX=Y andF=I, then S(I,Ω,Y0)=Y0. Note that (F+K)1(0)S(F,Ω,Y0) sinceK(X)Y0.

Definition 3.1. Givenε >0, a mappingKεC1(X,Y0) is called a regularε-approximation ofKonS(F,Ω,Y0) (regular finite-dimensionalε-approximation, for short) if

sup

xS(F,Ω,Y0)

Kε(x)K(x)ε. (3.3)

If alsoDKε(p)=0 for somepX,Kεis said to be based atp.

The (delicate) question about the existence of compact and regular approximations based atp is settled byTheorem 3.2below. For clarity, its proof is postponed until the next section.

Theorem3.2. For every finite-dimensional subspaceY0containingK(X), everyε >0, and everypX, there is a compact and regularε-approximation ofKonS(F,Ω,Y0)based atp.

More generally, there is a compact and regularε-approximation ofKonS(F,Ω,Y0)based at all the points of any given finite sequencep1,. . .,prX.

For the time being, it will suffice to define the degree at the value 0/ (F+K)(∂Ω). To justify the definition given in (3.6) below, it must be checked thatdp(F+Kε,Ω, 0) exists and is independent of the choice of the regularε-approximationKofK onS(F,Ω,Y0) based at pandof the subspaceY0. This is done in the next lemma, where we implicitly use the fact that the properness ofFonΩand the compactness ofKimply thatF+K is proper onΩ, so that (F+K)(∂Ω) is closed.

Lemma 3.3. Suppose that0/ (F+K)(∂Ω), that 0< ε <dist(0, (F+K)(∂Ω)), and that pXis a base-point ofF.

(i)IfKεC1(X,Y0)is a compact and regular ε-approximation of K onS(F,Ω,Y0) based atp, thenF+KεΦ0C1(X,Y),F+Kεis proper onΩ, 0/ (F+Kε)(∂Ω), and pis a base-point ofF+Kε. In particular,dp(F+Kε,Ω, 0)is defined.

(ii)WithKε as in (i), let Y0 be another finite-dimensional subspace of Y such that K(X)Y0and let KεC1(X,Y0)be a compact and regularε-approximation of KonS(F,Ω,Y0)based atp. Then,dp(F+Kε,Ω, 0)=dp(F+Kε,Ω, 0).

Proof. (i) That F+Kε Φ0C1(X,Y) follows from FΦ0C1(X,Y) and Kε compact (whenceDKε(x)᏷(X,Y) for allxX; see, e.g., [7, page 56]). The properness ofF+Kε

onΩis due to the properness ofF, the compactness ofKε, and the boundedness ofΩ.

Suppose now thatxΩand thatF(x) +Kε(x)=0. Then,F(x)= −Kε(x)Y0 and hencexS(F,Ω,Y0). As a result,(F+K)(x) = (F+K)(x)(F+Kε)(x) = K(x) Kε(x)ε <dist(0, (F+K)(∂Ω)), so that x /Ω. This means that 0/ (F+Kε)(∂Ω).

Thatpis a base-point ofF+Kεis obvious.

(8)

(ii) After replacingY0byY0+Y0andKεby a compact and regular finite-dimensional ε-approximation ofK onS(F,Ω,Y0+Y0) (whose existence follows fromTheorem 3.2), we may assume with no loss of generality thatY0Y0.

(ii-a) Assume first thatY0=Y0. Then,Kεt:=(1t)Kε+tKεis also a compact and regularε-approximation of K onS(F,Ω,Y0) based at pfor every t[0, 1] anddp(F+ Kε,Ω, 0)=dp(F+Kε,Ω, 0) follows readily fromCorollary 2.3.

(ii-b) Consider now the general case whenY0Y0. By (ii-a) andTheorem 3.2, it suf- fices to comparedp(F+Kε,Ω, 0) anddp(F+Kε,Ω, 0) whenKεis also anε-approximation for some 0< εε, sayKε=Kεfor consistency.

LetP0ᏸ(Y) project ontoY0and setε:=(ε/P0)ε. SinceP0K=KandKεis an ε-approximation ofKonS(F,Ω,Y0), we have

sup

xS(F,Ω,Y0)

P0Kε(x)K(x)P0εε. (3.4)

SinceS(F,Ω,Y0)S(F,Ω,Y0),P0Kε is a compact and regualrε-approximation ofK on S(F,Ω,Y0) based atp, so thatdp(F+Kε,Ω, 0)=dp(F+P0Kε,Ω, 0) by (ii-a). It remains to show that

dpF+P0Kε,Ω, 0=dpF+Kε,Ω, 0, (3.5) which follows from Corollary 2.3 after checking that F+ (1t)P0Kε+tKε does not vanish onΩfort[0, 1]. But ifxΩandF(x) + (1t)P0Kε(x) +tKε(x)=0, then xS(F,Ω,Y0) andF(x) +K(x)=(1t)(K(x)P0Kε(x)) +t(K(x)Kε(x)). Thus, by (3.4),F(x) +K(x)(1t)ε+ε <dist(0, (F+K)(∂Ω)), so thatx /Ω.

FromLemma 3.3, if 0/ (F+K)(∂Ω) and ifpXis a base-point ofF, the definition dF,p(F+K,Ω, 0) :=dp

F+Kε,Ω, 0 (3.6)

makes sense whenever 0< ε <dist(0, (F+K)(∂Ω)) andKεis a regular finite dimensional ε-approximation ofKbased atp.

Remark 3.4. In particular, ifK=0, we may chooseKε=0 and (3.6) reads asdF,p(F,Ω, y)=dp(F,Ω,y). Thus, forFΦ0C1(X,Y) proper onΩ, there is no difference between dp(F,Ω,y) defined inSection 2anddF,p(F,Ω,y) defined above.

4. Existence of compact regular approximations

This section is devoted to a generalization ofTheorem 3.2whenF is Fredholm of any index, which will be useful when dealing with homotopies. The finite-dimensional sub- spaceY0ofY and the mappingsFΦnC1(X,Y) for somenZandKC0(X,Y0) as well as the bounded open subsetΩofX are given once and for all. It is also assumed throughout thatF is proper on Ω, thatF(Ω) is bounded, and thatK is compact (i.e., bounded on bounded subsets sinceKis finite dimensional). Givenε >0 and a base-point pXofF, our goal is to findKεC1(X,Y0) such thatDKε(p)=0,Kεis compact, and supxS(F,Ω,Y0)Kε(x)K(x)ε, whereS(F,Ω,Y0) is given by (3.2). The compactness

(9)

ofS(F,Ω,Y0) or the definition of a regular finite-dimensionalε-approximation is not af- fected by the fact that the index ofFis not necessarily 0.

Definition 4.1. The finite-dimensional subspaceY1ofY is said to beS(F,Ω,Y0)-regular ifY0Y1 and there is P1ᏸ(Y) projecting ontoY1 such thatQ1F:XrgeQ1 is a submersion onS(F,Ω,Y0), whereQ1:=IP1.

IfY1 is an S(F,Ω,Y0)-regular subspace ofY, there is an open neighborhood U of S(F,Ω,Y0) inXsuch that

MY1

:=

Q1F1(0)U (4.1)

is a finite-dimensionalC1submanifold ofXcontainingS(F,Ω,Y0). The existence of reg- ular subspaces is settled by the following.

Lemma4.2. There is anS(F,Ω,Y0)-regular subspace ofY.

Proof. GivenxX, it is well known that a finite-dimensional direct complementYx of rge DF(x) remains a complement (though not necessarily direct) ofrge DF(ξ) forξin an open neighborhoodUxofx. SinceS(F,Ω,Y0) is compact, it may be covered by finitely many neighborhoodsUx1,. . .,Uxn. SetY1:=Yx1+···+Yxn+Y0, a finite-dimensional subspace ofYcontainingY0, and letZ1be any closed direct complement ofY1.

Denote byP1andQ1=IP1the projections onto the spacesY1andZ1, respectively.

IfxS(F,Ω,Y0), thenrge DF(x) +Y1=Y, so that, givenz1Z1, there arewX and y1Y1such thatDF(x)w+y1=z1. Thus,Q1DF(x)w=z1, which shows thatQ1F:X Z1is a submersion onS(F,Ω,Y0) and hence thatY1isS(F,Ω,Y0)-regular.

IfY1is anS(F,Ω,Y0)-regular subspace ofY, the functionK|M(Y1 )can be uniformly ap- proximated byC1functions on every compact subset ofM(Y1) and hence onS(F,Ω,Y0).

The tool needed to extendC1functions defined onC1finite-dimensional submanifolds is a variant of the Whitney embedding theorem, proved in [17, Theorem 7.1] and repro- duced inLemma 4.3below, showing that a compact subset of such a submanifold can always be “flattened” by a diffeomorphism of thewholespace.

Lemma4.3. LetX be a real Banach space,MX a finite-dimensionalC1 submanifold, andNM a compact subset. There is a finite-dimensional subspaceX1 ofX and aC1 diffeomorphismΦofXonto itself such thatΦ(N)X1.

Theorem4.4. For everyε >0and every finite sequencep1,. . .,prX, there is a compact and regularε-approximation ofKonS(F,Ω,Y0)with values inY0and based atp1,. . .,pr. Proof. Let Y1 be an S(F,Ω,Y0)-regular subspace of Y (Lemma 4.2) and choose M= M(Y1), N=S(F,Ω,Y0) in Lemma 4.3. The corresponding diffeomorphism Φ maps S(F,Ω,Y0) onto a compact subsetQ:=Φ(S(F,Ω,Y0)) of the finite-dimensional subspace X1 of X. By the Dugundji extension theorem (see, e.g., [7]), (KΦ1)|Q C0(Q,Y0) can be extended to a mappingKC0(X1,Y0) with values in the (compact) convex hull conv(K(S(F,Ω,Y0)))Y0.

(10)

Letπ1ᏸ(X) project ontoX1and let p1i:=π1Φ(pi)X1, 1ir. We claim that, givenε >0, there isKεC1(X1,Y0) such that

sup

x1X1

Kεx1

Kx1ε (4.2)

andDKε(p1i)=0, 1ir. Since dimX1<, this is clear without the requirement that DKε(p1i)=0. But it is also clear that such an approximation exists which is constant in some neighborhood ofp1iand hence satisfiesDKε(p1i)=0 for all indicesi.

Now, defineKε:=Kεπ1C1(X,Y0) andK:=Kπ1. Then, sup

x1Q

Kε(x1)K(x1)

=sup

x1Q

Kε x1

Kx1 sup

x1X1

Kε x1

Kx1ε (4.3) by (4.2) andDKε(Φ(pi))=DKε(p1i)=0, 1ir. Therefore,Kε:=KεΦC1(X,Y0) satisfiesDKε(pi)=0, 1ir, and

sup

xS(F,Ω,Y0)

Kε(x)K(x)

=sup

x1Q

KεΦ1x1

KΦ1x1=sup

x1Q

Kεx1

Kx1ε. (4.4)

To complete the proof, it remains to show thatKεis compact. But this follows at once from the remark that, by (4.2),Kε(X) is a bounded subset ofY0sinceKhas values in the

compact subset conv(K(S(F,Ω,Y0))).

5. Degree for finite-dimensional perturbations: homotopy variance We begin with a convenient definition.

Definition 5.1. The homotopyhC0([0, 1]×X,Y) will be calledΩ-admissible if it can be written in the formh=hΦ+hκwithhΦΦ1C1([0, 1]×X,Y),hΦproper on [0, 1]×Ω, andhκC0([0, 1]×X,Y) compact.

Ifh=hΦ+hκC0([0, 1]×X,Y) is anΩ-admissible homotopy withhκfinite dimen- sionaland if y /h([0, 1]×∂Ω), then the degreedhΦ(t,·),pt(h(t,·),Ω,y) is defined when- everptis a base-point ofhΦ(t,·) (see (3.2)). The next theorem explains how the degrees dhΦ(0,·),p0(h(0,·),Ω,y) anddhΦ(1,·),p1(h(1,·),Ω,y) are related.

Theorem5.2. Leth=hΦ+hκC0([0, 1]×X,Y)be anΩ-admissible homotopy withhκ finite dimensional andhΦ([0, 1]×Ω)bounded and suppose that0/ h([0, 1]×Ω). Ifp0 Xandp1Xare base-points ofhΦ(0,·)andhΦ(1,·), respectively, then

dhΦ(0,·),p0

h(0,·),Ω, 0=νdhΦ(1,·),p1

h(1,·),Ω, 0, (5.1) whereν:=σ(DxhΦΓ)∈ {−1, 1}andΓis any continuous curve in[0, 1]×Xjoining(0,p0) to(1,p1).

(11)

Proof. By the arguments of the proof ofLemma 3.3(i), the set h([0, 1]×Ω) is closed inY, so that dist(0,h([0, 1]×Ω))>0. Let 0< ε <dist(0,h([0, 1]×Ω)) be chosen once and for all. In order to useTheorem 4.4withFreplaced byhΦandKreplaced byhκ(and Ωreplaced by (0, 1)×Ω) so as to obtain regular approximations ofhκ, it is necessary to extend bothhΦ andhκ to the whole spaceR×X. This issue is straightforward forhκ, but it is immediately realized that extendinghΦto a Fredholm mapping is not such an easy matter. However, a notable exception to this statement occurs when DthΦ(0,·)= DthΦ(1,·)=0, for then the extensionhΦ ofhΦdefined byhΦ(t,x) :=hΦ(0,x) fort0 andhΦ(t,x) :=hΦ(1,x) fort1 is inΦ1C1(R×X,Y) and, evidently,hΦ|[0,1]×=hΦ|[0,1]× is proper.

At this stage, the pertinent remark is that neither the assumptions ofTheorem 5.2nor its conclusion is affected by changingh(t,x) intoh(ϕ(t),x) whereϕis anyC1homeomor- phism of [0, 1] onto itself such thatϕ(0)=0 andϕ(1)=1. In particular, this change does not modify the seth([0, 1]×Ω) or the mappingshΦ andhκ whent=0 ort=1. The only slightly less obvious point is thatν:=σ(DxhΦΓ) is unchanged. But this follows from the homotopy invariance of the parity since, as is readily checked,ϕand the iden- tity of [0, 1] are homotopic. Sinceϕcan be chosen so that (dϕ/dt)(0)=(dϕ/dt)(1)=0, it follows that, for the purpose of provingTheorem 5.2, we may and will assume with no loss of generality thatDthΦ(0,·)=DthΦ(1,·)=0.

Let then hΦ be the extension of hΦ introduced above and let hκ extend hκ in the same way, so thathκC0(R×X,Y) is compact and finite dimensional. It follows from Theorem 4.4withXreplaced byR×XandΩreplaced by (0, 1)×Ωthat given any finite- dimensional subspaceY0ofY such thathκ(R×X)Y0, there is a compact and regular finite-dimensionalε-approximationhκ,εC1(R×X,Y0) ofhκonS(hΦ, [0, 1]×Ω,Y0) := hΦ1(hΦ([0, 1]×Ω)Y0)[0, 1]×=hΦ1(hΦ([0, 1]×Ω)Y0)[0, 1]×Ω(thus inde- pendent of the extension) based at both (0,p0) and (1,p1). In particular,

Dxhκ,ε

0,p0

=Dxhκ,ε

1,p1

=0. (5.2)

In particular,hκ,ε(0,·) is a compact and regularε-approximation ofhκ(0,·)=hκ(0,·) on S(hΦ(0,·),Ω,Y0)=hΦ(0,·)1(hΦ({0} ×Ω)Y0)Ωbased atp0andhκ,ε(1,·) is a com- pact and regularε-approximation ofhκ(1,·)=hκ(1,·) onS(hΦ(1,·),Ω,Y0)=hΦ(1,·)1 (hΦ({1} ×Ω)Y0)Ωbased at p1. Also,hΦ({0} ×Ω) andhΦ({1} ×Ω) are bounded (as required by (3.1)) sincehΦ([0, 1]×Ω) is bounded by hypothesis. Therefore, by the definition (3.6),

dhΦ(0,·),p0

h(0,·),Ω, 0=dp0

hΦ(0,·) +hκ,ε(0,·),Ω, 0, (5.3) dhΦ(1,·),p1

h(1,·),Ω,y=dp1

hΦ(1,·) +hκ,ε(1,·),Ω,y. (5.4)

Now,hΦ+hκ,εΦ1C1([0, 1]×X,Y) is proper on [0, 1]×Ω(being a finite-dimensional perturbation ofhΦ) and 0/ (hΦ+hκ,ε)([0, 1]×∂Ω). ByTheorem 2.2,

dp1

hΦ(1,·) +hκ,ε(1,·),Ω,y=νdp0

hΦ(0,·) +hκ,ε(0,·),Ω,y, (5.5)

(12)

whereν:=σ(DxhΦΓ+Dxhκ,εΓ)∈ {−1, 1}andΓis any continuous curve in [0, 1]×X joining (0,p0) to (1,p1). We claim that, more simply,

ν=σDxhΦΓ. (5.6)

Indeed, consider the homotopyH(s,t) :=DxhΦ(Γ(t)) +sDxhκ,ε(Γ(t)), so that H(0,t)= DxhΦΓandH(1,t)=DxhΦΓ+Dxhκ,εΓ. Then,H(s, 0)=DxhΦ(0,p0) andH(s, 1)= DxhΦ(1,p1) for all s[0, 1] since Dxhκ,ε(0,p0)=Dxhκ,ε(1,p1)=0 (see (5.2)). This is to say that the endpoints remain invertible during the homotopy, whenceσ(H(0,·))= σ(H(1,·)). This proves (5.6) and thus the theorem by (5.3), (5.4), and (5.5).

Generalizations and corollaries ofTheorem 5.2will be mentioned inSection 7. For the time being, we only clarify theF-dependence of the degreedF,p.

With F and K satisfying the assumptions required in Section 3 to define dF,p(F+ K,Ω, 0) by (3.6), suppose also thatF+K=G+LwithGΦ0C1(X,Y) andLC0(X,Y) compact and finite dimensional. Then,G=F+KLis proper onΩandG(Ω) is bound- ed sinceF(Ω) is bounded by hypothesis. Therefore, withT:=F+K=G+Land assum- ing that 0/ T(∂Ω), we have a degreedF,p(T,Ω, 0) and a degreedG,q(T,Ω, 0) wheneverp andqare base-points ofFandG, respectively. Up to sign, these two degrees coincide, as shown below.

Theorem5.3. Above,

dF,p(T,Ω, 0)=νdG,q(T,Ω, 0), (5.7) whereν∈ {−1, 1}is the parity of any path (although perhaps not apparent, this is a path of Fredholm operators of index0; see the proof){(1t)DF(γ(t)) +tDG(γ(t)) :t[0, 1]} withγC0([0, 1],X)being a curve joiningptoq.

Proof. The formula (5.7) is the special case of Theorem 5.2where Γ(t)=(t,γ(t)) and where p0=p,p1=q,hΦ(t,x)=F(x) +t(KL)(x), andhκ(t,x)=(1t)K(x) +tL(x).

Note thathΦisC1becauseKL=GFisC1(although neitherKnorLneed beC1) and Fredholm of index 0 sinceFis Fredholm of index 0 andKLis compact andC1. Note also thathΦ([0, 1]×Ω) is bounded sinceF(Ω) is bounded and thathΦ+hκ=F+K=T

is independent oft.

Theorem 5.3helps to clarify the question of extending the definition (3.6) in the case that F has no base-point. Indeed, given anyqX, it is always possible to write T= F+K in the form T=G+L in such a way thatq is a base-point of G. For instance, chooseG=F+Aand L=KAwhereAᏸ(X,Y) is a suitable operator with finite rank. Then, dG,q(T,Ω, 0) makes sense (if 0/ T(∂Ω)) and can be used in place of the nonexistingdF,p(T,Ω, 0).

6. Degree for finite-dimensional perturbations: main properties

We now prove that the degree (3.6) possesses the main properties valid in theC1case:

normalization, excision, and additivity on domain. We continue to assume thatΩXis a bounded open subset and thatF(Ω) is bounded.

参照

関連したドキュメント

In this paper we give an improvement of the degree of the homogeneous linear recurrence with integer coefficients that exponential sums of symmetric Boolean functions satisfy..

We prove the existence of weak solutions of higher order degenerated quasihnear elliptic equations The mmn tools are the degree theory for generahzed monotone mappings and

Definition An embeddable tiled surface is a tiled surface which is actually achieved as the graph of singular leaves of some embedded orientable surface with closed braid

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 proof uses a set up of Seiberg Witten theory that replaces generic metrics by the construction of a localised Euler class of an infinite dimensional bundle with a Fredholm

CHANDRA, On the degree of approximation of a class of functions by means of Fourier series, Acta Math. CHANDRA, A note on the degree of approximation of continuous functions,

Since the continuum random tree is a random dendrite, the results of the previous chapter are readily applicable and so we are immediately able to deduce from these heat

Some existence results for this problem are established using the fixed point method and topological degree