Skip to content
BY 4.0 license Open Access Published by De Gruyter April 16, 2019

The E-Cohomological Conley Index, Cup-Lengths and the Arnold Conjecture on T2n

  • Maciej Starostka EMAIL logo and Nils Waterstraat

Abstract

We show that the E-cohomological Conley index, that was introduced by the first author recently, has a natural module structure. This yields a new cup-length and a lower bound for the number of critical points of functionals on Hilbert spaces. When applied to the setting of the Arnold conjecture, this paves the way to a short proof on tori, where it was first shown by C. Conley and E. Zehnder in 1983.

MSC 2010: 37J10; 53D40; 58E05

1 Introduction

Motivated by questions of celestial mechanics from the beginning of the 20th century, Arnold conjectured in the sixties that every Hamiltonian diffeomorphism on a compact symplectic manifold (M,ω) has at least as many fixed points as a function on M has critical points. Let us recall that a diffeomorphism ψ:MM is called Hamiltonian if there exists a smooth map H:×M, H(t+1,x)=H(t,x), such that ψ=η1, where the family {ηt}t satisfies

(1.1) { d d t η t = X H ( η t ) , η 0 = id ,

and XH stands for the time-dependent vector field given by

d H ( ) = ω ( X H , ) .

Consequently, p is a fixed point of ψ if and only if it is the initial condition of a 1-periodic solution of (1.1), and so Arnold’s famous conjecture can be reformulated dynamically as follows.

Arnold Conjecture.

The Hamiltonian system

(1.2) x ˙ ( t ) = X H ( x ( t ) )

has at least as many 1-periodic orbits as a function on M has critical points.

The aim of this paper is to point out a new approach to the Arnold conjecture which proves it on tori, where it was first shown by C. Conley and A. Zehnder in [3]. Let us point out that several approaches to the Arnold Conjecture have appeared since then. We refer to [10, pp. 215–216], but want to mention in particular that Chaperon proved it in [2] for tori by a concise geometric argument (cf. also [9]). Our proof is short as well, and it will be future work to investigate if our methods also apply to cases where the conjecture is still open. For example, the Arnold conjecture has not been proved for T2n×Pm where a similar analytical setting can be introduced (see [6]). To the best of our knowledge, the previous methods only work to some extent in this case (see, however, [12] for partial results), and therefore it is worthwhile to develop new approaches.

However, let us point out that, apart from these important applications, our methods are of independent interest and can be outlined as follows. In [15] the first author introduced the E-cohomological Conley index for isolated invariant sets of flows in Hilbert spaces. Roughly speaking, it is a generalization of the classical Conley index for flows on locally compact spaces by using E-cohomology, which is a generalized cohomology theory for subsets of Hilbert spaces that was constructed by Abbondandolo in [1] (cf. also [7]). The first aim of this paper is to introduce a module structure for the E-cohomological Conley index, which allows us to define a relative cup-length for triples of closed and bounded subsets of Hilbert spaces. Secondly, we consider this numerical invariant for isolating neighborhoods of 𝒮-flows in Hilbert spaces (cf. [8, 16]), and show that it is a lower bound for the number of critical points of gradient flows as in classical Ljusternik–Schnirelman theory. Here we substantially use properties of the E-cohomological Conley index that were recently obtained by the first author in a joint work with Izydorek, Rot, Styborski and Vandervorst in [11]. Finally, we apply our Ljusternik–Schnirelman-type theorem to the functionals in the setting of the Arnold conjecture on T2n. This yields an estimate from below for the number of contractible 1-periodic solutions of (1.2), and the obtained bound is indeed the one that Arnold conjectured.

This paper is organized with the intention of guiding the reader through our proof of the Arnold conjecture in as straightforward a manner as possible. Therefore, in the second section, we only introduce the material that is necessary to understand the basics of our approach and postpone more technical proofs to Section 4. Our discussion of the Arnold conjecture can be found in between, in the third section.

2 The E-Cohomological Conley Index and Cup-Lengths

2.1 Module Structure for E-Cohomology

We begin this section by recalling E-cohomology from [1], where we slightly modify the definition as in [11]. Let E be a separable real Hilbert space and E+, E- closed subspaces such that E=E+E-. In what follows we denote by H Alexander–Spanier cohomology with compact supports, for which we refer to [14] and the nice survey in [1, Section 1]. Moreover, we let 𝒱 be the set of all finite-dimensional subspaces of E-, which is partially ordered by inclusion and directed.

If U,V,W𝒱 are such that W=VU and dim(U)=1, then we can decompose W into two subspaces by setting

W + = { w W : w , u 0 } ,
W - = { w W : w , u 0 } ,

where u0 is a fixed element in U. Note that the choice of u corresponds to an orientation of the one-dimensional space U, and changing this orientation swaps W+ and W-.

Figure 1 
            Decomposition of X by W.
Figure 1

Decomposition of X by W.

We set for a closed and bounded subset X of E,

X W = X ( E + × W ) , X W + = X ( E + × W + ) , X W - = X ( E + × W - )

and note that XW=XW+XW- as well as XV:=X(E+×V)=XW+XW-. (See Figure 1.) If now AX is closed, then we obtain a relative Meyer–Vietoris sequence

H k ( X W + , A W + ) H k ( X W - , A W - ) H k ( X V , A V ) Δ V , W k H k + 1 ( X W , A W )
H k + 1 ( X W + , A W + ) H k + 1 ( X W - , A W - ) .

In the more general case that W=VU and dim(U)=n2, we decompose U into n one-dimensional subspaces U=U1Un and set Wi=VU1Ui for 1in as well as W0=V. Then the previous construction yields n Mayer–Vietoris homomorphisms

Δ W i - 1 , W i k + i - 1 : H k + i - 1 ( X W i - 1 , A W i - 1 ) H k + i ( X W i , A W i )

and their composition is a homomorphism Hk(XV,AV)Hk+n(XW,AW). Hence we have constructed for any q and V,W𝒱, VW, a homomorphism

Δ V , W q ( X ) : H q + dim ( V ) ( X V , A V ) H q + dim ( W ) ( X W , A W ) .

As noted in [1, Proposition 2.2], these maps do not depend on the choice of the one-dimensional subspaces Ui and their orientations. In summary, {Hq+dim(V)(XV,AV),ΔVWq(X,A)} is a direct system of abelian groups over the directed set 𝒱.

Definition 1.

Let AX be closed and bounded subsets of E. The E-cohomology group of index q of (X,A) is the direct limit

H E q ( X , A ) = lim V 𝒱 { H q + dim ( V ) ( X V , A V ) , Δ V , W q ( X , A ) } ,

and we set as usual HEq(X):=HEq(X,).

The inclusions ιV,W:XVXW for V,W𝒱 yield an inverse system {Hp(XV),ιV,W} over 𝒱. We define for p the group H0p(X) as the inverse limit

H 0 p ( X ) := lim V 𝒱 { H p ( X V ) , ι V , W } .

In what follows, we denote elements of H0p(X) by [αV]0 if αVHp(XV), and correspondingly elements of HEq(X,A) by [αV]E if αVHq+dim(V)(XV,AV).

Let us point out that H0(X) is a ring if we define the product of [αV]0H0p(X) and [βV]0H0q(X) by

[ α V ] 0 [ β V ] 0 = [ α V β V ] 0 H 0 p + q ( X ) .

It is readily seen from the naturality of the cup product that this is a sensible definition.

Proposition 2.

The group HE(X,A) is a right module over H0(X), where the module multiplication is induced by the cup product.

Proof.

We define for [αV]0H0r(X) and [βV]EHEq(X,A),

[ β V ] E [ α V ] 0 := [ β V α V ] E H E q + r ( X , A ) .

This product is well defined, as if βW=ΔV,WqβV and αV=ιV,WαW, then

Δ V , W q + r ( β V α V ) = Δ V , W q + r ( β V ι V , W α W ) = ( Δ V , W q β V ) α W = β W α W ,

where we have used that the coboundary operators of the Mayer–Vietoris sequence commute with products in multiplicative cohomology theories (cf. [4, Proposition 17.2.1]). ∎

Let now ΩE be closed and bounded and such that XΩ. The inclusions jV:XVΩV induce homomorphisms jV:Hp(ΩV)Hp(XV) for V𝒱, and it is readily seen that they actually yield a ring homomorphism

j : H 0 ( Ω ) H 0 ( X ) .

Consequently, we obtain the following corollary from Proposition 2.

Corollary 3.

For every XΩE, HE(X,A) is a right H0(Ω)-module.

Henceforth we denote the module product of αH0r(Ω) and βHEp(X,A) by

β α H E p + r ( X , A ) .

We conclude this section with the following crucial definition of a relative cup-length.

Definition 4.

Let AXΩ be closed and bounded subsets of E.

  1. If HE(X,A)=0, we set

    CL ( Ω ; X , A ) = 0 .

  2. If HE(X,A)0 but βα=0 for every βHE(X,A) and αH0>0(Ω), then we set

    CL ( Ω ; X , A ) = 1 .

  3. If there are k2, β0HE(X,A) and α1,α2,,αk-1H0>0(Ω) such that

    β 0 α 1 α k - 1 0 ,

    then

    CL ( Ω ; X , A ) k .

In order to keep the definition short we have not defined when actually CL(Ω;X,A)=k for k2 as the stated estimate in the final part of Definition 4 is good enough for our purposes (see Theorem 10).

2.2 The E-Cohomological Conley Index and Critical Points

The first aim of this subsection is to introduce the E-cohomological Conley index and to define a module structure for it. Let E be a real separable Hilbert space and L:EE an invertible selfadjoint operator for which there exists a sequence {En}n of finite-dimensional subspaces of E such that L(En)=En, EnEn+1 and nEn¯=E. Let UE be open. Following [8], we call a vector field F:UEE, F(u)=Lu+K(u) an 𝒮-vector field if K:UEE is a locally Lipschitz compact operator. Note that every 𝒮-vector field generates a local flow ηt satisfying

(2.1) d d t η t = - F η t , η 0 = id ,

which we call an 𝒮-flow.

Let us now assume that η is a global 𝒮-flow on U, and let us denote by

Inv ( Ω , η ) = { x Ω : η t ( x ) Ω , t }

the maximal η-invariant subset of ΩU.

Definition 5.

A closed and bounded set ΩU is called an isolating neighborhood of η if Inv(Ω,η)int(Ω), where int(Ω) denotes the interior of Ω.

Let now Ω be an isolating neighborhood of η and S:=Inv(Ω,η).

Definition 6.

We call a closed and bounded pair (X,A) of subsets of Ω an index pair for S if

  1. A is positively invariant with respect to X, i.e. given xA and t>0 with η[0,t](x)X, then η[0,t](x)A,

  2. S = Inv ( X A ¯ , η ) int ( X A ¯ ) ,

  3. if yX, t>0 and ηt(y)X, then there exists t<t such that η[0,t](y)X and η(t,y)A.

It was shown in [11, Lemma 2.7] that every isolated invariant set S as above has an index pair.

Note that the space E splits as E=E+E-, where E± are the spectral subspaces with respect to the positive and negative part of the spectrum of L. Henceforth, we denote by HE the E-cohomology with respect to this splitting. The following crucial result was proved in [11, Proposition 2.8].

Proposition 7.

If (X,A) and (X,A) are index pairs for S, the groups HE(X,A) and HE(X,A) are isomorphic.

Hence the next definition is sensible (cf. [11, Definition 2.9]).

Definition 8.

The E-cohomological Conley index of S is defined by

ch E ( S ) = H E ( X , A ) ,

where (X,A) is an index pair for S.

If we want to emphasize the isolating neighborhood Ω instead of the isolated invariant set S, we will also write chE(Ω) to denote the E-cohomological Conley index.

When taking the module structure from Section 2 into account, it is readily seen by arguing as in [11, Proposition 2.8] that HE(X,A) and HE(X,A) are actually isomorphic as H0(Ω)-modules. Hence we obtain as a consequence of Proposition 7 the following important result.

Corollary 9.

The cup-length CL(Ω;X,A) does not depend on the choice of the index pair (X,A) such that XΩ.

Consequently, we can define

CL ( Ω , S ) := CL ( Ω ; X , A ) ,

where (X,A) is any index pair for S such that XΩ. As S is uniquely determined by Ω and the flow η, we will sometimes denote this cup-length by CL(Ω,η) if we want to emphasize η.

Let us now assume that η is the gradient flow with respect to a differentiable functional f:U, i.e. the map F:UEE is of the form F=f. As before, we assume that η is global. Let Ω be an isolating neighborhood of η and S=Inv(Ω,η). We denote by Crit(f,Ω) the set of critical values of f|Ω and can now state the main theorem of this paper.

Theorem 10.

If f has only finitely many critical points in Ω, then the number of critical values of f|Ω is bounded below by the cup-length of Ω with respect to S, i.e.

(2.2) # Crit ( f , Ω ) CL ( Ω , S ) .

Note that by Theorem 10, the right-hand side in (2.2) is obviously also a lower bound for the number of critical points of f in Ω. We will prove Theorem 10 in Section 4.

3 The Arnold Conjecture on the Torus T2n

Let T2n denote the standard Torus of dimension 2n and let ω0 be its standard symplectic structure. Let HC2(S1×T2n,) be a 1-periodic Hamiltonian and XH the induced vector field on T2n given by

d H ( ) = ω ( X H , ) .

We consider the Hamiltonian equation

(3.1) x ˙ ( t ) = X H ( x ( t ) ) ,

and the aim of this section is to prove the following deep theorem.

Theorem 1 (Arnold Conjecture on T2n).

For every C2-Hamiltonian on T2n there exist at least 2n+1 contractible solutions of (3.1).

The above theorem was first proved by Conley and Zehnder in [3] (cf. also [10]). We now recall the analytical setting from their proof in the next section, and then use Theorem 10 for a short proof of Theorem 1.

3.1 The Analytical Setting

Before proving Theorem 1, let us first recall the analytical setting from [10] (see also [13]). In what follows, we let J be the symplectic standard matrix

( 0 I n - I n 0 ) ,

which is related to the symplectic form on T2n by

ω 0 ( x , y ) = x , J y ,

where , denotes the standard Euclidean scalar product.

We now start with the case of 2n and consider the space of smooth loops C(S1,2n) in 2n. If we set ek(t):=etk2πJ, k, then any xC(S1,2n) is represented by its Fourier series

(3.2) x ( t ) = k e k ( t ) x k ,

where xk2n, k. The Sobolev space H12(S1,2n) is the Hilbert space which is obtained as the completion of C(S1,2n) with respect to the scalar product

x , y 1 2 = x 0 , y 0 + 2 π k | k | x k , y k .

There is an orthogonal decomposition

H 1 2 ( S 1 , 2 n ) = Z 0 Z - Z +

into a 2n-dimensional subspace Z0 and closed infinite-dimensional subspaces Z+ and Z- which correspond to k=0, k>0 and k<0 in the Fourier-series expansion (3.2), respectively. In what follows, we denote by P0, P+ and P- the corresponding orthogonal projections.

Now let HC2(S1×2n,) be a Hamiltonian such that |H(x)|C|x|2 at infinity and such that the second spatial derivative H′′ is globally bounded. We define a functional ΦH:C(S1,2n) by the formula

(3.3) Φ H ( x ) = a ( x ) - b ( x ) := 1 2 0 1 - J x ˙ ( t ) , x ( t ) 𝑑 t - 0 1 H ( t , x ( t ) ) 𝑑 t .

The importance of ΦH comes from the fact that the critical points of ΦH are periodic solutions of the Hamilton equation (3.1). It is easy to see that ΦH extends to H12(S1,2n), and

(3.4) Φ H = L + K ,

where L=a=P+-P- is a selfadjoint Fredholm operator and K=-b=-j*H is a compact map because of the compactness of the adjoint j:L2H12 of the inclusion.

On a general manifold, it is a delicate problem to define spaces H12(S1,M) as H12(S1,2n) contains non-continuous functions which consequently have no local meaning. However, for a torus one can overcome this problem by using the universal covering 2nT2n=2n/2n. Then smooth Hamiltonians on T2n are in one-to-one correspondence with 2n-invariant smooth Hamiltonians on 2n, where 2n acts on 2n by translations. By a slight abuse of notation, we will denote by H both the Hamiltonian on the torus and the Hamiltonian lifted to 2n. Note that the lifted Hamiltonian on 2n is 2n-invariant and therefore its second spatial derivative is bounded and it obviously satisfies the growth condition mentioned above. Now the corresponding functional ΦH in (3.3) is 2n-invariant as well, and therefore it descends to a functional on the quotient space

:= Z 0 / 2 n × Z + × Z - = T 2 n × Z + × Z - .

3.2 Proof of Theorem 1

We suppose as in the previous subsection that HC2(S1×T2n,) is a given Hamiltonian. Let us note at first that F=ΦH in (3.4) is an 𝒮-vector field, even though the operator L is not invertible. Indeed, if we write F=L^+K^:=(L+P0)+(K-P0), where P0 is the orthogonal projection onto the finite-dimensional kernel of L as introduced above, then F is the sum of an invertible selfadjoint operator and a compact map.

As is a Hilbert manifold, we cannot directly apply the E-cohomological Conley index which we only have defined for flows on open subsets of a Hilbert space. However, if we use a tubular neighborhood, the definition can easily be extended to Hilbert manifolds of the type M×E, where M is a closed manifold and E is a Hilbert space. In the case of , the construction is as follows. We embed into E^=4n×Z+×Z- in such a way that every S1 in T2n=S1××S1 is mapped to the unit circle in 2. We consider the open set

U := D 0 2 n × Z + × Z - E ^

of E^, where D0={(x,y)2:0<x2+y2<4} is a punctured disc of radius 2 in 2, and we let π:𝒩 be the standard projection to T2n on D02n and the identity on Z+ and Z-. The map ΦH can be extended to U by

Ψ H ( x ) = Φ H ( π ( x ) ) + i = 1 2 n ( 1 - r i ( x ) ) 2 ,

where ri(x) denotes the polar coordinate in 2 of the projection of xU to the i-th component of (2)2n. Note that the extension is done in such a way that ΨH and ΦH have the same critical points. We denote by K~ the compact operator which is the sum of K^ and (i=12n(1-ri(x))2).

Note that ΨH=L^+K~ is an 𝒮-vector field, and the negative and positive spectral subspaces of the selfadjoint isomorphism L^ are given by

E + = 4 n Z + , E - = Z - .

Now Theorem 1 can be obtained as follows. Since K is bounded, there is R>0 such that R>K(x) for all xU. We set

X = C 2 n × B ( Z + , R ) × B ( Z - , R ) ,

where B(Z±,R) are the closed balls of radius R in Z± and CD0 is a closed annulus containing S1. Note that the boundary X is given by the non-disjoint union

X = ( C 2 n × B ( Z + , R ) × B ( Z - , R ) ) ( C 2 n × B ( Z + , R ) × B ( Z - , R ) ) ( C 2 n × B ( Z + , R ) × B ( Z - , R ) ) .

Let now X1,X2,X3 denote the three parts of X in the above order and let η be the flow induced by -ΨH as in (2.1). Firstly, if xX1 but neither in X2 nor in X3, then there is some t>0 such that η(0,t](x)XX. Secondly, if xX2, then P+x=R and we have

L ^ x + K ~ ( x ) , P + x R = L P + x , P + x R + K ( x ) , P + x R > R - K ( x ) > 0 .

Consequently, the vector field L^x+K~(x) is pointing outwards the sphere B(Z+,R). Hence, if xX1X2 but xX3, then η, which is the flow induced by -ΨH, moves x into the interior of X. Finally, if xX3, we analogously see that

L ^ x + K ~ ( x ) , P - x R < 0

as P-x=R. Hence those x leave X under η. It is now readily seen that

( X , A ) = ( C 2 n × B ( Z + , R ) × B ( Z - , R ) , C 2 n × B ( Z + , R ) × B ( Z - , R ) )

is an index pair for S=Inv(X,η) in the sense of Definition 6, where we have set A:=X3 for simplicity of notation.

To find the E-cohomology of (X,A), let VZ- be of finite dimension. Then

( X V , A V ) = ( C 2 n × B ( Z + , R ) × B ( V , R ) , C 2 n × B ( Z + , R ) × B ( V , R ) ) ,

where B(V,R) denotes the ball of radius R in V. Hence we get for k

H k ( X V , A V ) = H k ( X V / A V ) = H k ( S ( V , R ) T 2 n ) = H k - dim ( V ) ( T 2 n ) ,

where S(V,R) denotes the sphere of radius R in V. Moreover, if WV is another finite-dimensional subspace, then the Mayer–Vietoris homomorphism ΔV,Wk mapping

H k + dim ( V ) ( X V , A V ) = H k + dim ( V ) ( S ( V ) T 2 n )

to

H k + dim ( W ) ( X W , A W ) = H k + dim ( W ) ( S ( W ) T 2 n )

is by definition just the suspension isomorphism. Hence we obtain

H E ( X , A ) = H ( T 2 n ) .

Finally, to find the cup-length, we note at first that for the isolating neighborhood X, and any finite-dimensional subspace VZ-,

H ( X V ) = H ( C 2 n × B ( Z + ; R ) × B ( V ; R ) ) = H ( T 2 n ) .

Hence CL(X,η) is just the ordinary cup-length of the torus T2n, which is 2n+1. By Theorem 10, this is a lower bound for the number of critical points of ΦH in X, and so we have proved the Arnold conjecture on T2n.

4 Proof of Theorem 10

We will need the following two properties of the cup-length CL that we introduced in Definition 4. As the proofs are purely algebraic, we leave it to the reader to check that they follow by obvious modifications from [5, Lemmas 2.2 and 2.3].

Lemma 1.

If BAXY are closed and bounded subsets of E, then

CL ( Y ; X , B ) CL ( Y ; X , A ) + CL ( Y ; A , B ) .

Lemma 2.

If AXY1Y2 are closed and bounded subsets of E, then

CL ( Y 2 ; X , A ) CL ( Y 1 ; X , A ) .

Now let us consider an isolating neighborhood Ω for the flow η generated by the gradient of the function f:U in Theorem 10. As we suppose that there are only finitely many critical points of f in Ω, the set of critical values Crit(f,Ω) is finite as well, say, c1<<ck. Let MiΩ denote the set of stationary points with values ci, and set for 1ijk,

M i j = { x Ω : ω ( x ) α ( x ) M i M i + 1 M j } ,

where α(x) and ω(x) denote as above the α and ω limits of xE under the flow η. Note that M1k consists of all the critical points of f inside Ω and all the orbits connecting them. Consequently,

M 1 k = Inv ( Ω , η ) .

Now let (X,A) be an index pair for M1k.

Lemma 3 (Morse Filtration).

There exist sets

X 0 = A X 1 X k = X

such that (Xj,Xi-1) is an index pair for Mij.

Proof.

We let bi(ci,ci+1), i=1,,k-1, be regular values of f, set bk=, and define X0=A as well as Xi:=Xf-1(-,bi], i=1,,k. Then it is readily seen that (Xj,Xi-1) is an index pair for Mij as Mij consists of all critical points x such that f(x){ci,,cj} and all the orbits connecting them. ∎

If we now apply Lemma 1k times, we get

(4.1) CL ( Ω ; X , A ) i = 1 k CL ( Ω ; X i , X i - 1 ) .

On the other hand, (Xi,Xi-1) is an index pair for Mii, which is a set consisting of a finite number of stationary points. Therefore we can choose an isolating neighborhood Ωi for Mii, where Ωi is a disjoint union of discs. If now (Xi,Xi-1) is an index pair for Mii such that XiΩi, then by Corollary 9 and Lemma 2

CL ( Ω ; X i , X i - 1 ) = CL ( Ω ; X i , X i - 1 ) CL ( Ω i ; X i , X i - 1 ) 1 ,

where the last inequality follows from the fact that the groups H0q>0(Ωi) are trivial. Hence, by (4.1),

CL ( Ω ; X , A ) k

and Theorem 10 is shown, as k is the number of critical values of f in Ω.


Communicated by Paul H. Rabinowitz


Funding statement: Maciej Starostka was supported by the grants Preludium9 of the National Science Centre, no. 2015/17/N/ST1/02527 as well as BEETHOVEN2 of the National Science Centre, Poland, no. 2016/23/G/ST1/ 04081. Nils Waterstraat was supported by the grant BEETHOVEN2 of the National Science Centre, Poland, no. 2016/23/G/ST1/04081.

Acknowledgements

We would like to thank Kazimierz Gȩba and Marek Izydorek for many inspiring discussions, as well as Thomas Schick for clarifying remarks about our groups H0(X).

References

[1] A. Abbondandolo, A new cohomology for the Morse theory of strongly indefinite functionals on Hilbert spaces, Topol. Methods Nonlinear Anal. 9 (1997), no. 2, 325–382. 10.12775/TMNA.1997.017Search in Google Scholar

[2] M. Chaperon, An elementary proof of the Conley–Zehnder theorem in symplectic geometry, Dynamical Systems and Bifurcations (Groningen 1984), Lecture Notes in Math. 1125, Springer, Berlin (1985), 1–8. 10.1007/BFb0075631Search in Google Scholar

[3] C. C. Conley and E. Zehnder, The Birkhoff–Lewis fixed point theorem and a conjecture of V. I. Arnol’d, Invent. Math. 73 (1983), no. 1, 33–49. 10.1007/BF01393824Search in Google Scholar

[4] T. tom Dieck, Algebraic Topology, EMS Textbk. Math., European Mathematical Society (EMS), Zürich, 2008. 10.4171/048Search in Google Scholar

[5] Z. A. Dzedzej, K. Gȩba and W. Uss, The Conley index, cup-length and bifurcation, J. Fixed Point Theory Appl. 10 (2011), no. 2, 233–252. 10.1007/s11784-011-0065-9Search in Google Scholar

[6] B. Fortune, A symplectic fixed point theorem for 𝐂Pn, Invent. Math. 81 (1985), no. 1, 29–46. 10.1007/BF01388770Search in Google Scholar

[7] K. Gȩba and A. Granas, Infinite-dimensional cohomology theories, J. Math. Pures Appl. (9) 52 (1973), 145–270. Search in Google Scholar

[8] K. Gȩba, M. Izydorek and A. Pruszko, The Conley index in Hilbert spaces and its applications, Studia Math. 134 (1999), no. 3, 217–233. 10.4064/sm-134-3-217-233Search in Google Scholar

[9] H. Hofer, Lagrangian embeddings and critical point theory, Ann. Inst. H. Poincaré Anal. Non Linéaire 2 (1985), no. 6, 407–462. 10.1016/s0294-1449(16)30394-8Search in Google Scholar

[10] H. Hofer and E. Zehnder, Symplectic Invariants and Hamiltonian Dynamics, Birkhäuser Adv. Texts Basler Lehrbücher, Birkhäuser, Basel, 1994. 10.1007/978-3-0348-8540-9Search in Google Scholar

[11] M. Izydorek, T. O. Rot, M. Starostka, M. Styborski and R. C. A. M. Vandervorst, Homotopy invariance of the Conley index and local Morse homology in Hilbert spaces, J. Differential Equations 263 (2017), no. 11, 7162–7186. 10.1016/j.jde.2017.08.007Search in Google Scholar

[12] Y.-G. Oh, A symplectic fixed point theorem on T2n×𝐂Pk, Math. Z. 203 (1990), no. 4, 535–552. 10.1007/BF02570755Search in Google Scholar

[13] P. H. Rabinowitz, Minimax Methods in Critical Point Theory with Applications to Differential Equations, CBMS Reg. Conf. Ser. Math. 65, American Mathematical Society, Providence, 1986. 10.1090/cbms/065Search in Google Scholar

[14] E. H. Spanier, Algebraic Topology, McGraw-Hill, New York, 1966. 10.1007/978-1-4684-9322-1_5Search in Google Scholar

[15] M. Starostka, Morse cohomology in a Hilbert space via the Conley index, J. Fixed Point Theory Appl. 17 (2015), no. 2, 425–438. 10.1007/s11784-015-0216-5Search in Google Scholar

[16] M. Styborski, Conley index in Hilbert spaces and the Leray–Schauder degree, Topol. Methods Nonlinear Anal. 33 (2009), no. 1, 131–148. 10.12775/TMNA.2009.010Search in Google Scholar

Received: 2018-03-12
Revised: 2019-01-18
Accepted: 2019-03-13
Published Online: 2019-04-16
Published in Print: 2019-08-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 28.4.2024 from https://www.degruyter.com/document/doi/10.1515/ans-2019-2044/html
Scroll to top button