• 検索結果がありません。

Special Issue on Space Dynamics

N/A
N/A
Protected

Academic year: 2022

シェア "Special Issue on Space Dynamics"

Copied!
28
0
0

読み込み中.... (全文を見る)

全文

(1)

CONJECTURE

TADEUSZ DOBROWOLSKI Received 18 April 2002

The Schauder conjecture that every compact convex subset of a metric linear space has the fixed-point property was recently established by Cauty (2001). This paper elaborates on Cauty’s proof in order to make it more detailed and there- fore more accessible. Such a detailed analysis allows us to show that the convex compacta in metric linear spaces possess the simplicial approximation property introduced by Kalton, Peck, and Roberts. The latter demonstrates that the orig- inal Schauder approach to solve the conjecture is in some sense “correctable.”

1. Introduction

Throughout most of this paper,X will be a compact convex subset of a sepa- rable metric linear space (E, · ) which is not necessarily locally convex. We can always assume (and we will) that · is anF-norm onE; hence, we have x+yx+y,txx for allx, yE and1t1; in general, · is not homogeneous. Generalizing the classical Brouwer theorem, in 1930, Schauder [28] claimed the proof of the fixed-point property ofX; unfortunately, his argument contained a gap. As an effect, the question of whether every com- pact convex subset of a metric linear space has the fixed-point property was put, in August 1935, [25, Problem 54]; it became known as Schauder’s conjec- ture. Since then, many partial results were obtained but the general case went unsettled. Almost all those partial results were based on the so-called finite- dimensional approximation property that some convex compacta possess. The property requires that, idX, the identity map onX, is, in a uniform way, arbitrar- ily closely approximated by mapsψ:XXsuch that dim(ψ(X))<. Here, the dimension “dim” can be understood in any reasonable (i.e., “linear” or purely topological) way; for further information, see [7]. In the case thatEis locally convex, such an approximation can be obtained via partitions of unity. Such an

Copyright©2003 Hindawi Publishing Corporation Abstract and Applied Analysis 2003:7 (2003) 407–433

2000 Mathematics Subject Classification: 54H25, 47H10, 55M20, 46A16, 46A55, 46T20, 52A07 URL:http://dx.doi.org/10.1155/S1085337503211015

(2)

approach predominates in the textbook proofs of the Schauder-Tychonofftheo- rem which states that every compact convex subset of a locally convex topolog- ical vector space has the fixed-point property. This most transparent fact in the area was published by Tychonoff[29] in 1935.

Note that, as a consequence of some advanced results of infinite-dimensional topology, a convex compactum X having the finite-dimensional property not only has the fixed-point property, but actually is homeomorphic either to [0,1]n, nN, or to [0,1], the Hilbert cube, see [7, 12]. As later becomes clear, it is remarkable that Schauder in his failed attempt did not intend to verify the finite-dimensional property, but a weaker property which was later formalized by Kalton et al. in [21]. This “local” property, which is called the simplicial approximation property in [21], yields the fixed-point property, but itself is rather far from the finite-dimensional property, which in turn could be viewed as a “global” property. In [21], the simplicial approximation property was used to verify the fixed-point property of certain convex compacta without extreme points. Those compacta were earlier discovered by Roberts.

Finally, in 1999, Cauty [5] provided a proof of the Schauder conjecture. His proof did not rely on any of the above properties, but was based on the exis- tence of a certain resolution mapϕ:ZX from a certain countable-dimen- sional compactumZontoX. Actually, Cauty has proved a more general result stating that, for an arbitrary convex subsetCof a topological vector space, every map f :CCsuch that f(C) is contained in a compact subset ofC(i.e., ev- eryrelatively compactmap f :CC) has the fixed-point property. We mention that a similar result for the locally convex case was obtained by Mazur [26] and Hukuhara [19] in 1938 and 1950, respectively.

This paper elaborates on Cauty’s proof of the metric case only; that is, on the following result.

Theorem1.1 (Cauty [5]). LetXbe a convex compactum of a metric linear space (E, · )and let f :XXbe a map. There existsxXwith f(x)=x.

Revisiting Cauty’s proof, we were able to isolate its two basic ingredients.

The first ingredient is of purely topological (or, better to say, metric) nature;

it deals with the construction of the resolutionϕ. The second ingredient has a linear (affine) flavor and, employing the resolution mapϕ, it establishes a certain approximation property. We hope that our analysis makes Cauty’s proof more accessible. Surprisingly, such an analysis enables us to verify the simplicial ap- proximation property ofX (cf. [8]). This shows that the original approach of Schauder is “correctable.” As it stands, the proof of the fixed-point property is very much complex. Now, knowing that the simplicial approximation property holds for everyX, it is reasonable to ask for a simple way of verifying it.

According to a result of [10], every noncompact convex subsetCof a met- ric linear space contains a topological copy of [0,) (this observation, for the normed case, was made long before by Klee [22]). As easily observed, it follows

(3)

that such aC fails the fixed-point property. Combining this with Cauty’s re- sult, we infer that for convex subsets of metric linear spaces, only compacta have the fixed-point property. On the other hand, there are examples of noncompact convex subsets of locally convex topological vector spaces that do have the fixed- point property, see [10, Example 4.1].

Cauty has extended his technique to show the fixed-point property for com- pacta that are uniformly contractible. Recall that a spaceXis locally uniformly contractible if there exists a continuous mapµ(x, y;t) (which is referred to as an equiconnecting map) defined for all (x, y) in a neighborhoodUof the diagonal inX×XandtI=[0,1], such thatµ(x, y; 0)=x,µ(x, y; 1)=y, andµ(x, x;t)= xfor all (x, y;t)U×I. IfUcan be taken as the wholeX×X, thenXis called uniformly contractible. All convex sets are uniformly contractible. Ifr:CX is a retraction of a convex setContoX, then Xis uniformly contractible be- causeµ(x, y;t)=r((1t)x+t y) is as required. If a topological groupGis con- tractible, that is, there exists a homotopyH:G×IGsuch thatH(g,0)=gand H(g,1)=e(eis the unit element ofG) for all (g, t)G×I, thenGis uniformly contractible viaµ(g, h;t)=(H(e, t))1·H(g·h1, t)·h, see [7]. It follows that every retract of such aGis uniformly contractible. On the other hand, a metriz- able uniformly contractible compactum is a retract of a contractible metrizable topological group [3]. Generalizing the classical Lefschetz-Hopf fixed-point the- orem, Cauty [4] has shown that every self-map f of a locally uniformly con- tractible compactum X has a fixed-point if the Lefschetz number Λ(f)=0.

In particular, all uniformly contractible compacta have the fixed-point prop- erty. Necessary modifications for obtaining the proof of this fact are presented at the end ofSection 2(see also the end ofSection 3). Observe that every uni- formly contractible compactum is a contractible and locally contractible space.

The question of whether a contractible and locally contractible metrizable com- pactum has the fixed-point property remains unanswered.

For the locally convex case, the fixed-point property ofXyields, for an upper semicontinuous (USC) convex-valued mapF:XX, the fixed-point property ofF, that is, there exists xX such thatxF(x). (Here, by a convex-valued map F, we mean a multivalued map such thatF(x) is a convex compactum;

such a map is USC if {xX|F(x)U}is open for every open setUX.) Again, known techniques require partitions of unity, a tool that does not work for the nonlocally convex case. Our approach enabled us to show that, for ev- ery dense convex subsetCof the convex compactumX, every map f :CX admits approximate fixed-points; that is, there exists a sequence (xn)Csuch that limf(xn)xn =0, seeCorollary 2.6. In [9], we used this fact, together with yet another approximation result that was based on a certain technique de- veloped by Cellina-Lasota [6], to obtain the fixed-point property for every USC convex-valued mapFof a compact convex set of an arbitrary topological vector space. The details of such a generalization of Cauty’s result go beyond the scope of this paper and will not be included here. For the record, we mention that

(4)

the convex-valued case has been established, among others, by Kakutani [20], Bohnenblust and Karlin [2], Fan [15], and Glicksberg [16].

This introduction alludes to the author’s observations that were made while revisiting Cauty’s paper [5]. It contains only a handful of historical remarks on the fixed-point property of convex compacta that were needed to put those ob- servations in some historical perspective. The author did not intend to make an expository paper on the subject with such overwhelming literature.

2. The simplicial approximation property

Following [21], we say that a convex setCEhas thesimplicial approximation propertyif for everyε >0, there exists a finite-dimensional compact convex set CεCsuch that ifSis any finite-dimensional simplex inC(i.e.,Sis a convex hull of finitely many vectors inC), then there exists a continuous mapγ:SCε

withγ(x)x< ε,xS. According to [21, Remark (2), page 217] (see also [11, Lemma 5.1]), if a convex compactumXhas the simplicial approximation property, then it has the fixed-point property. We have the following obvious generalization of that fact.

Lemma 2.1. Assume that a convex setCE has the simplicial approximation property. Then every map f :CC¯ has an approximate fixed-point, that is,lim f(xn)xn =0for some sequence(xn)C.

Proof. According to the definition above, for eachnN, pick a convex com- pactumCnCsuch that for a simplexS, we have a sequenceγn:SCnwith limγnidS =0. Approximate f |Cnby a mapfn:CnSn, whereSnis a sim- plex inC, see [7]. By Brouwer fixed-point theorem, there existsxnCnsuch that γn(fn(xn))=xn. It is easy to see that limf(xn)xn =0.

Say that a convex setCEhas the0-simplicial approximation propertyif for everyε >0, there exists a countable-dimensional convex setCεCsuch that if Sis any0-simplex inC(i.e.,Sis a convex hull of countable many vectors of C), then there exists a relatively compact mapγ:SCε withγ(x)x< ε, xS. Actually, this property is equivalent to the one that requiresCεto be an 0-simplex. To see this, use the separability ofEto find an0-simplexCεCε

that is dense inCε. Now, letγ be a relatively compact map ofSintoCε. Using the fact thatCε is countable dimensional, find a map ofCε intoCε that is as close to the identity as we wish, see [7]. Composing this map withγ, we obtain a relatively compact map ofSintoCε.

Lemma2.2. If a convex setCEhas the0-simplicial approximation property, then it has the simplicial approximation property.

Proof. Fixε >0. Use the above modification of the0-simplex approximation property to find, forε/3, an0-simplexCεCas required in the definition. Let

(5)

γ0:CεCεbe a relatively compact map withγ0(x)x< ε/3,xCε. There exists a mapg :γ0(Cε)Cε whose range is contained in a finite-dimensional compact convex setKεand that satisfiesg(x)x< ε/3 for allx, see [7, Lemma 2]. Now, letSbe a finite-dimensional simplex inC. Assuming thatCεis dense (here we employ the separability ofE), we find a mapγ:SCε withγ(x) x< ε/3,xS. Then, the compositiongγ0γ:SKεis as required.

Remark 2.3. Replacing the convexity assumption onCεby the assumption that CεANR, we obtain a weaker version of the0-simplicial approximation prop- erty. Such a weaker property also yields the approximate fixed-point property ofC. To see this, we employ the ANR property ofCε (use a factorization of an ANR through a locally finite-dimensional, locally compact, separable space) to approximate a mapCεCby one whose range isS, an0-simplex. Next, using the contractibility ofS, we extend this map toc(Cε), the metric cone overCε. Composing such a map with a relatively compactγ:SCec(Cε), which is provided by the definition (of such a weaker version of the0-simplicial approx- imation property), we obtain a relatively compact self-map ofc(Cε). An applica- tion of the so-called generalized Schauder theorem (stating that every relatively compact self-map of an AR space has the fixed-point property [13, page 94]) yields a fixed-pointxeCe, which is a counterpart ofxnthat was obtained in the proof ofLemma 2.1.

Below, we state the main technical ingredient of Cauty’s proof that establishes a property which, in a sense, is equivalent to the0-simplicial approximation property ofX.

Proposition2.4 (cf. [5, Lemma 3]). For the convex compactumX, there exist a countable-dimensional compactumZ and a mapϕ:ZX such that, if Z is embedded onto a linearly independent subset of a metric linear space(F,| · |)so that the affine extensionϕˆ: conv(Z)Xofϕis continuous (the existence of such an embedding is ensured byLemma 2.5below), then, for everyε >0, every separable metrizable spaceYANR, and every mapξ:YX, there exists a map

η:Y−→conv(Z) (2.1)

such that

(i)ϕ(η(y))ˆ ξ(y)< εfor allyY,

(ii)η(Y)conv2(Z)is a compact subset ofconv(Z); hence,ηis relatively com- pact.

Here byconv2(Z), we mean{t1z1+t2z2|z1, z2Z,0t1, t21, t1+t2=1}. Lemma2.5. Given a mapϕ:ZXof a compactumZ, there exists a metric linear space(F,| · |)topologically containingZas a linearly independent subset such that

(6)

ˆ

ϕ: conv(Z)Xdefined by

ˆ ϕ

k

i=1

tizi

= k i=1

tiϕzi, (2.2)

whereki=1ti=1,ti0,kN, is continuous.

Proof ofLemma 2.5. EmbedZonto a linearly independent subset of (2,| · |).

Extendϕto a linear (not necessarily continuous) map ˆϕ:2E. Define anF- norm on2by

|x| =x+ϕ(x)ˆ . (2.3)

LetF=(2,| · |). Then ˆϕ:FEis continuous, and (Z,| · |) yields the original

topology ofZ.

Corollary2.6 (cf. [8, Theorem]). Every convex setCEwith a compact clo- sureXhas the0-simplicial approximation property. Consequently, byLemma 2.1, every map f :CXhas an approximate fixed-point.

Proof. LetZ,ϕ, and ˆϕbe that ofProposition 2.4. Fixε >0. Since the compactum Zis countable dimensional, so is conv(Z). An application of an argument of [7]

shows that there exist an0-simplexXεCand a mapϕε: conv(Z)Xεsuch thatϕ(z)ˆ ϕε(z)< ε/2,zconv(Z). LetSbe0-simplex inC. By a theorem of Haver [17] (see also [7, Note 4]), we haveSAR. ApplyProposition 2.4to Y=Sand toξ, the inclusion ofSintoX. There exists a relatively compactη: Sconv(Z) withϕˆη(x)x< ε/2,xS. Letγ=ϕεη:XXε. We have thatγ(x)xϕεη(x)ϕˆη(x)+ϕˆη(x)xε/2 +ε/2=ε.

Proof ofTheorem 1.1. Applying Corollary 2.6, we conclude that f has an ap- proximate fixed-point. By the compactness ofX,f has a fixed-point.

Reduction Fact 2.7. The assertion ofProposition 2.4holds provided it does hold for the class of spacesYthat are separable, metrizable, locally finite dimensional, and locally compact.

Proof. LetYbe an arbitrarily separable, metrizable ANR, letξ:YXbe a map, and letε >0. By [18, page 138], there exist a countable locally finite simplicial complex|N|(considered in the metric topology) and maps

α:Y−→ |N|, β:|N| −→Y (2.4) such that

ξβα(y)ξ(y)

2, (2.5)

(7)

yY. The space|N|is separable, locally finite dimensional, and locally com- pact. For the mapξβ:|N| →X, we can find a relatively compact map ˜η:|N| → conv(Z) withϕˆη(p)˜ ξβ(p)< ε/2,p∈ |N|. Setη=η˜α:Y conv(Z).

We see thatηis relatively compact andϕˆη(y)ξ(y) = ϕˆη(α(y))˜ ξ(β α(y))+ξ(βα(y))ξ(y)< ε/2 +ε/2=ε.

Since every map of a countable locally finite simplicial metric complex into X can be uniformly, arbitrarily closely approximated by a map whose range is contained in an0-simplexS,Proposition 2.4holds if we verify its assertion for allY =S, whereSis such a simplex. In this way,Proposition 2.4establishes a property that is equivalent to the0-simplicial approximation property. A di- rect argument for the0-simplicial approximation property in caseXenjoys the simplicial approximation property is not known to us. Also the relationship of the finite-dimensional approximation property to both simplicial approxima- tion properties is unclear.

The results ofProposition 2.4,Lemma 2.5, andReduction Fact 2.7can be ex- tended to the uniformly contractible spaces X; consequently, the fixed-point property holds for such X. We say thatψ: conv(Z)X is µ-affine (µ is an equiconnecting mapX) ifµ(ψ(z1), ψ(z2);t)=(1t)z1+tz2for allz1, z2conv(Z) andtI.Proposition 2.4holds for a uniformly contractible compactumX if ϕadmits a continuousµ-affine extension ˆϕ: conv(Z)X, where conv(Z) is a convex subset (of a vector space) with a metric topology that makes the con- vex combination map (z1, z2;t)(1t)z1+tz2continuous. Such an extension can be obtained by inspecting the proof ofLemma 2.5. As in that proof, we em- bed the compactumZonto a linearly independent subset of2. Next, extendϕ to (not necessarily continuous) ˆϕ: conv(Z)Xthat isµ-affine. Finally, letting dfor a metric onX,ρ(z1, z2)= |z1z2|+d( ˆϕ(z1),ϕ(zˆ 2)) defines a required metric on conv(Z), a convex subset of2. Before we show how to obtain ˆϕ, for x1, . . . , xnXand (t1, . . . , tn)sn= {(t1, . . . , tn)|ti0 for alliandni=1ti=1}, we inductively let

µn

x1, . . . , xn;t1, . . . , tn

=µn1

x1, . . . , xn1;t1/1tn

, . . . , tn1/1tn

(2.6)

iftn=1, andµn(x1, . . . , xn;t1, . . . , tn)=xnotherwise; setµ1(x1; 1)=x1. Now, well order the setZby a relation<and, forz=n

i=1tiziwhere (t1, . . . , tn)snand z1, . . . , znZ with z1<···< zn, define ˆϕ(z)=µn(ϕ(z1), . . . , ϕ(zn);t1, . . . , tn). It is easily seen that ˆϕisµ-affine. Using the cross-section method (see [1, page 271]), we can show that conv(Z)=(conv(Z), ρ) is countable dimensional. Since conv(Z) is locally contractible and contractible, by a theorem of Haver [17], conv(Z)AR. To conclude the fixed-point property ofX by such a general- ized version ofProposition 2.4, we find a sequence of relatively compact maps (ηn) ofY=conv(Z) with lim ˆϕηn= fϕ. By a generalized Schauder theoremˆ [13, page 94], there exists ynY such thatη(yn)=yn. We can assume that the

(8)

sequence ( ˆϕ(yn)) converges toxX. Such anx is a fixed-point of f because f(x)=limf( ˆϕ(yn))=lim ˆϕ(ηn(yn))=lim ˆϕ(yn)=x.

3. The proof ofProposition 2.4

This section contains the proof ofProposition 2.4. The proof will rely onLemma 3.1, a purely topological fact, which is an abstraction on Cauty’s construction of the resolutionϕ:ZX (that will be provided inSection 4). However, after completing the proof ofProposition 2.4, we give a rough sketch of the proof of Lemma 3.1for the reader who is not interested in all details.

Lemma3.1. Let(X, d)be a metric space. There exist an inverse sequence(Zn, πnn+1), where eachZn is a (finite-dimensional) compactum, a mapkn:XZn, a finite open cover{Wα|αA(n)}ofZn,nN, and a mapϕ:Z=lim←−(Zn, πnn+1)X that satisfy the following conditions:

(1)Zis a countable-dimensional compactum;

(2)d(ϕ(z),V¯α)2∆nwheneverπn(z)WαandαA(n), whereπn:ZZn is an obvious projection andVα=kn1(Wα),αA(n);

(3)∆n=sup{diamd(Vα)|αA(n)} → ∞asn→ ∞.

The space Z=lim←−(Zn, πnn+1) is equipped with a metric dZ, a restriction of dZ((xn),(yn))=

n=12ndn(xn, yn) defined on n=1Zn, where eachdnis a com- patible, bounded by 1, metric onZn.

Proof ofProposition 2.4. The convex compactumX will be equipped with the metricdthat is induced by anF-norm · .Lemma 3.1provides us with a com- pactumZand a mapϕ:ZXthat satisfy (1), (2), and (3). (Further on, the countable dimensionality ofZwill not be used.) Moreover,Zis assumed to be a linearly independent subset of a metric linear space (F,| · |) such that the affine extension ˆϕ: conv(Z)X is continuous. Letξ:YXbe a map, and letε >0, ε1. ApplyingReduction Fact 2.7, we may assume that Y is a locally finite- dimensional, locally compact, separable, metrizable space. There exists an open cover{Yi}i=1 such that, for everyi,Yi is relatively compact, dim(Yi)i, and YiYj= ∅whenever|ij|>1, see [14, page 291].

By the compactness ofZ, for eachi, there existsδi>0 that satisfies δi ε

2i+ 3, dZ(z, z)< δi=⇒ |zz|< 1

2(2i+ 3)i (3.1) for allz, zZ. ChooseniNsuch thatni1< niand max(2ni,ni)< δi/3. For iand for every 1mni, define Ᏻnmi= {mnikniξ)1(B(z,2ni))|zZm} (here,B(z, r) stands for the opendm-ball inZm that is centered atzwith radius r) and0ni = {(kniξ)1(Wα)=ξ1(Vα)|αA(ni)}. Write Ᏻ0=n0in0i1, and defineᏳi=0ni1

m=1nmi1ni

m=1mni, an open cover ofY. Here, for open coversᏭ1, . . . ,p, we designateᏭ1∩ ··· ∩p= {A1∩ ··· ∩Ap|Aii, i= 1, . . . , p}. WriteᏳ= {GYi|Gifor somei}. Pick an open coverᐁthat is a

(9)

star refinement ofᏳ. Using the fact dimYii, find a partition of unity{λU}U

such that supp(λU)=λ1((0,1])U and i+2k=1supp(λUk)= ∅if each Uk is a subset ofYi. For eachUᐁ, choose the smallestiNwithUYi, and pick zUZsuch thatπni(zU)kniξ(U). Defineη(y)=

UλU(y)zU,yY. FixyYand writeA(y)= {U|ysupp(λU)}. SupposeiNis the first in- dex such thatyYi. It may happen thatyYi+1. Consequently, the cardinality ofA(y) is at most 2i+ 3. It also follows thatU(y)=

{U|UA(y)} ⊂Yi Yi+1. We conclude that for eachl=niorni+1, there existsα(l)A(l) such that U(y)(klξ)1(Wα(l)). On the other hand, for eachUA(y) and for a cer- tain suchl, we haveπl(zU)klξ(U); so, we haveU(klξ)1(Wα)=ξ1(Vα) andπl(zU)klξ(U)WαforαA(l), wherel=niorl=ni+1. It shows that ξ(y)Vαand, by (2) ofLemma 3.1, thatd(ϕ(zU),V¯α)2∆l; hence,ϕ(zU) ξ(y)3∆l<max(δi, δi+1) becauselis eitherniorni+1. We can estimate

ϕˆη(y)ξ(y)=

UA(y)

λU(y)ϕzU

ξ(y)

(2i+ 3)ϕzU

ξ(y)

(2i+ 3) maxδi, δi+1

ε.

(3.2)

The last inequality follows from the first part of (3.1). This shows (i).

To show (ii), we first partition the familyA(y) intoA1(y) andA2(y);U Aj(y) if and only ifi+j1 is the first index so thatUYi+j1, j=1,2. As previously, for eachl=niorni+1andml, there existsz(l,m)Zmsuch that

U(y)

πml klξ1Bz(l,m),2l

=

klξ1πmni1Bz(l,m),2l. (3.3) Hence, ifπl(zU)klξ(U), thenπm(zU)B(z(l,m),2l). It follows that, givenm with 1mni,πm(zU)B(z(ni,m),2ni) for allUA1(y); so, diamdm{πm(zU)| UA1(y)} ≤2ni. Since

diamdZ(S)

m=1

2mdiamdm

πm(S)

ni

m=1

2mdiam(S)

+ 2ni for everySZ,

(3.4)

we conclude that diamdZ{zU|UA1(y)} ≤2ni+ 2niδi. In a similar way, we show that diamdZ{zU|UA2(y)} ≤2ni+1+ 2ni+1δi+1. Hence, by the second part of (3.1),

zU1zU1< 1

2(2i+ 3)i, zU2zU2< 1

2(2i+ 5)(i+ 1) (3.5)

(10)

for allzU1, zU1A1(y) and allzU2, zU2A2(y). FixzUj, whereUjAj(y), and lettj=

{λU(y)|UAj(y)},j=1,2. From (3.5), we obtain η(y)

t1zU1+t2zU2=

UA(y)

λU(y)zUzU1

+

UA2(y)

λU(y)zUzU2

(2i+ 3) 1

2(2i+ 3)i+ 1 2(2i+ 5)(i+ 1)

<1 i.

(3.6) Now, if{η(yk)}is a sequence in conv(Z), then either{yk}has a subsequence contained in someYi(this subsequence, in turn, contains a convergent subse- quence becauseη( ¯Yi) is compact), or else it contains a subsequence{yi(n)}such thatyi(n)Yi(n)for alln. In the latter instance, from the above estimate, we have

|η(yi(n))(t1nz1n+t2nz1n)| ≤1/i(n) for sometn1, tn20 witht1n+tn2=1 and some zn1, zn2Z. By the compactness ofZ, it now easily follows that{η(yi(n))}con- tains a subsequence that converges to conv2(Z). Since conv2(Z) is compact, (ii)

is shown.

In what follows, we present a general overview of the argument that justifies Lemma 3.1. More precisely, we give a sketch of the proof ofLemma 3.1assuming that{Vα}and the inverse sequence (Zn, πnn+1) satisfy conditions (3.7), (3.8), and (3.9). We will not comment on the construction of the sequence (Zn, πnn+1). This will be done inSection 4.

Remark 3.2. The continuity ofϕis easily obtained from condition (2) ofLemma 3.1 as follows. For everyn, {πn1(Wα)|αA(n)} is an open cover of Z. If z, zπn1(Wα), thend(ϕ(z), ϕ(z))d(ϕ(z),V¯α) + diam( ¯Vα) +d(ϕ(z),V¯α) 5∆n0.

Now, we first defineϕin case (3.7) holds.

Remark 3.3. FixzZand letA(z)= {αA(n)|πn(z)Wα}. DefineFn(z)= {V¯α|αA(z)}. Assume

dHFn(z), Fm(z)2∆n (3.7) for 1nm (here,dH stands for the Hausdorffmetric induced by d). Let ϕ(z)=x, where{x} =limdHFn(z). Note that (3.7), together with condition (3) ofLemma 3.1, implies that the sequence{Fn(z)}converges to a singleton in the hyperspace ofX; hence,ϕis well defined. On the other hand,Fn(z)V¯αfor all αA(z). Applying (3.7) again, we easily obtain (2) ofLemma 3.1.

Next, we indicate how the countable dimensionality of Z can be achieved (here, we can compare this argument with that of Zarichny˘ı [30]). We stress that the countable dimensionality ofZ is essential for deducingCorollary 2.6from Proposition 2.4.

(11)

Remark 3.4. Suppose that eachZnis a tower of compactaZnk (inSection 4, de- noted byMnk), 0k < n, such that

dimZnkk, πnn+11ZnkZkn+1 (3.8) for every 0k < n. Furthermore, supposeZn\Znn1=

nfor a certain family ᐁn= {Uβ|βB(n)}that consists of finite, pairwise disjoint open subsets ofZn with

diamdm

πmnUβ

2n+1 (3.9)

for every 0m < nandβB(n). Then, we haveZ=P0

n=1Pn, whereP0=

n=1πn1(n) andPn=πn1(Znn1). Since condition (3.9) yields diamdZn1 (UβB))<2n+2for everyβB(n),nN, it follows that for everyn,P0can be covered by the finite family{πn1(Uβ)|βB(n)}which consists of open pair- wise disjoint sets of diameter<2n+2; hence, dim(P0)0. On the other hand, it can be easily checked thatPn=πn1(Znn1)=lim←−m(Pn), πkm|πk(Pn)). From (3.8), we infer thatπm(Pn)Zmn1(and consequently, dim(πm(Pn))n1 be- cause dim(Zmn1)n1) for allmn. This yields dim(Pn)n1.

In the case thatXis merely a uniformly contractible compactum, the proof ofProposition 2.4requires the following adjustments (that were initiated at the end ofSection 2). Formally, the definition ofη is the same, but to guarantee Proposition 2.4(i) and (ii), we must modify the choice ofδimade in condition (3.1). Assumingδi+1δi, it suffices to haved(µ2i+3(x1, . . . , x2i+3;t1, . . . , t2i+3), x) εford(xi, x)< δiandi=1, . . . ,2i+ 3, andρ(t1z1+···+t2i+5z2i+5+t1z1+···+ t2i+5z2i+5, tz+tz)<1/ifordZ(zi, z)< δianddZ(zi, z)< δi,i=1, . . . ,2i+ 5, where (t1, . . . , t2i+5, t1, . . . , t2i+5 )s4i+10andt=2i+5

i=1 ti,t=2i+5

i=1 ti. 4. The proof ofLemma 3.1

We begin with a statement of the three main points of Cauty’s original construc- tion ofϕ:ZX. Having done this, we show howLemma 3.1can be deduced.

The detailed Cauty’s construction is performed in Sections4.1and4.2.

Step 1. With the metric compactumX=(X, d), we associate an inverse sequence (Kn, qn) of finite simplicial complexesKn, dim(Kn)n, which are nerves of cer- tain open finite coversᐂn= {Vα|αAn}ofX, which is indexed in such a way thatα=αimpliesVα=Vα. WritingKn=nerv(ᐂn), we identify elements ofAn

with the vertices ofKn. Later, the set of the vertices ofKis denoted by Vert(K);

for a vertexαVert(K), st(α, K) stands for the open star ofαin|K|, the body ofK. We further assume that

(K1)ᐂn+1n (i.e., ᐂn+1 is inscribed inᐂn) and the simplicial map qn: Kn+1Knis onto, which is a consequence of (K2);

参照

関連したドキュメント

pole placement, condition number, perturbation theory, Jordan form, explicit formulas, Cauchy matrix, Vandermonde matrix, stabilization, feedback gain, distance to

By an inverse problem we mean the problem of parameter identification, that means we try to determine some of the unknown values of the model parameters according to measurements in

This technique allows us to obtain the space regularity of the unique strict solution for our problem.. Little H¨ older space; sum of linear operators;

We have formulated and discussed our main results for scalar equations where the solutions remain of a single sign. This restriction has enabled us to achieve sharp results on

[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

To derive a weak formulation of (1.1)–(1.8), we first assume that the functions v, p, θ and c are a classical solution of our problem. 33]) and substitute the Neumann boundary

We finally wish to remark that our results can be viewed as a first step towards the regularity theory of obstacle problems with integrands G being not of power growth.. The

This paper presents an investigation into the mechanics of this specific problem and develops an analytical approach that accounts for the effects of geometrical and material data on