Skip to main content
Log in

Conley index theory without index pairs. I: The point-set level theory

  • Published:
Journal of Fixed Point Theory and Applications Aims and scope Submit manuscript

A Correction to this article was published on 28 December 2023

This article has been updated

Abstract

We propose a new framework for Conley index theory. The main feature of our approach is that we do not use the notion of index pairs. We introduce, instead, the notions of compactifiable subsets and index neighbourhoods and formulate and prove basic results in Conley index theory using these notions. We treat both the discrete time case and the continuous time case.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Data availability

Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

Change history

References

  1. Bourbaki, N.: Éléments de Mathématique. Topologie Générale, Ch. 1–4. Hermann, Paris (1971)

  2. Conley, C.: Isolated Invariant Sets and the Morse Index. CBMS Regional Conference Series in Mathematics, vol. 38. American Mathematical Society, Providence (1978)

    Book  Google Scholar 

  3. Franks, J., Richeson, D.: Shift equivalence and the Conley index. Trans. Am. Math. Soc. 352, 3305–3322 (2000)

    Article  MathSciNet  Google Scholar 

  4. Lewis, L.G., Jr.: Open maps, colimits, and a convenient category of fibre spaces. Topol. Appl. 19, 75–89 (1985)

    Article  MathSciNet  Google Scholar 

  5. Manolescu, C.: Seiberg–Witten–Floer stable homotopy type of three-manifolds with \(b_1=0\). Geom. Topol. 7, 889–932 (2003). Errata: arXiv:math/0104024v5

    Article  MathSciNet  Google Scholar 

  6. Morita, Y.: Conley index theory without index pairs. II. In preparation

  7. Mrozek, M.: Leray functor and cohomological Conley index for discrete dynamical systems. Trans. Am. Math. Soc. 318, 149–178 (1990)

    Article  MathSciNet  Google Scholar 

  8. Mrozek, M.: Shape index and other indices of Conley type for local maps on locally compact Hausdorff spaces. Fund. Math. 145, 15–37 (1994)

    MathSciNet  Google Scholar 

  9. Robbin, J.W., Salamon, D.: Dynamical systems, shape theory and the Conley index. Ergodic Theory Dyn. Syst. 8*(Charles Conley Memorial Issue), 375–393 (1988)

    MathSciNet  Google Scholar 

  10. Salamon, D.: Connected simple systems and the Conley index of isolated invariant sets. Trans. Am. Math. Soc. 291, 1–41 (1985)

    Article  MathSciNet  Google Scholar 

  11. Sánchez-Gabites, J.J.: An approach to the shape Conley index without index pairs. Rev. Mat. Complut. 24, 95–114 (2011)

    Article  MathSciNet  Google Scholar 

  12. The Stacks project authors, The Stacks project.https://stacks.math.columbia.edu (2021)

  13. Szymczak, A.: The Conley index for discrete semidynamical systems. Topol. Appl. 66, 215–240 (1995)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

I would like to thank Hokuto Konno for his helpful comments on Floer theory. This work was supported by JSPS KAKENHI Grant Number 19K14529.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yosuke Morita.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Part 3. Appendix

Appendix A. Proper maps of compactly generated weak Hausdorff spaces

In this appendix, we prove some results on proper maps that are needed in this paper. As in the other parts of the paper, we assume every topological space to be compactly generated weak Hausdorff.

We define proper maps as follows:

Definition A.1

Let \(f :X \rightarrow Y\) be a continuous map of topological spaces. We say that f is proper if, for any compact Hausdorff subset L of Y, the inverse image \(f^{-1}(L)\) is compact Hausdorff.

Lemma A.2

Let \(f :X \rightarrow Y\) and \(g :Y \rightarrow Z\) be two continuous maps.

  1. (1)

    If f and g are proper, gf is also proper.

  2. (2)

    If gf is proper, f is also proper.

Proof

  1. (1)

    Obvious.

  2. (2)

    Let L be a compact Hausdorff subset of Y. We have

    $$\begin{aligned} f^{-1}(L) \subset (gf)^{-1}(g(L)). \end{aligned}$$

    Since Z is weak Hausdorff, g(L) is a compact Hausdorff subset of Z. Since gf is proper, the inverse image \((gf)^{-1}(g(L))\) is a compact Hausdorff subset of X. On the other hand, since Y is weak Hausdorff, L is closed in Y. Hence, \(f^{-1}(L)\) is closed in X (and therefore, in \((gf)^{-1}(g(L))\)). Thus, \(f^{-1}(L)\) is compact Hausdorff. \(\square \)

Lemma A.3

Proper maps are stable under base change.

Proof

Let \(f :X \rightarrow Y\) be a proper map and \(g :Y' \rightarrow Y\) be a continuous map. Consider the following pullback diagram:

Take any compact Hausdorff subset \(L'\) of \(Y'\). Since Y is weak Hausdorff, \(g(L')\) is a compact Hausdorff subset of Y. Since f is proper, the inverse image \(f^{-1}(g(L')) \ (= g'(f'^{-1}(L')))\) is a compact Hausdorff subset of X. The following is a pullback diagram:

We see that \(f'^{-1}(L)\) is compact Hausdorff because it is a fibre product of two compact Hausdorff spaces. Thus, \(f'\) is proper. \(\square \)

Lemma A.4

Let \(f :X \rightarrow Y\) be a continuous map of topological spaces. Then, the following two conditions are equivalent:

  1. (i)

    f is proper.

  2. (ii)

    For any continuous map \(g :L \rightarrow Y\) with L compact Hausdorff, the fibre product \(L \times _Y X\) is compact Hausdorff.

Proof

(ii) \(\Rightarrow \) (i): Consider the case when g is an inclusion.

(i) \(\Rightarrow \) (ii): By Lemma A.3, the base change \(f' :L \times _Y X \rightarrow L\) is proper. Since L is compact Hausdorff, it follows that \(L \times _Y X \ (= f'^{-1}(L))\) is also compact Hausdorff. \(\square \)

Lemma A.5

Proper maps are universally closed.

Proof

By Lemma A.3, it suffices to verify that proper maps are closed. Let \(f :X \rightarrow Y\) be a proper map. Let A be a closed subset of X. Since Y is compactly generated, it suffices to verify that, for any continuous map \(g :L \rightarrow Y\) with L compact Hausdorff, \(g^{-1}(f(A))\) is closed in L. Consider the following pullback diagram:

By Lemma A.4, \(L \times _Y X\) is compact Hausdorff. We see that \(f' :L \times _Y X \rightarrow L\) is a continuous map between compact Hausdorff spaces, hence closed. Thus, \(g^{-1}(f(A)) \ (= f'(g'^{-1}(A)))\) is closed in L. \(\square \)

Lemma A.6

An inclusion is proper if and only if it is closed.

Proof

The ‘if’ part holds because closed subsets of compact Hausdorff spaces are compact Hausdorff. The ‘only if’ part follows from Lemma A.5. \(\square \)

Lemma A.7

Let \(f :K \rightarrow X\) be a continuous map with K compact Hausdorff. Then, f is proper.

Proof

Take any compact Hausdorff subset L of X. Since X is weak Hausdorff, L is closed in X, and the inverse image \(f^{-1}(L)\) is closed in K. Since K is compact Hausdorff, \(f^{-1}(L)\) is also compact Hausdorff. \(\square \)

Lemma A.8

Let \(f :X \rightarrow Y\) be a continuous map of topological spaces. Suppose that there exists a finite closed cover \((A_i)_{i \in I}\) of X such that \(f|_{A_i} :A_i \rightarrow Y\) is proper for each \(i \in I\). Then, f is proper.

Proof

Take any compact Hausdorff subset L of Y. For each \(i \in I\), we see that \(f^{-1}(L) \cap A_i\) is compact Hausdorff. We have a continuous map

$$\begin{aligned} g :\coprod _{i \in I} (f^{-1}(L) \cap A_i) \rightarrow X \end{aligned}$$

induced from the inclusions. Since X is weak Hausdorff, \(f^{-1}(L)\), which is the image of g, is compact Hausdorff. \(\square \)

Appendix B. Shift equivalences and the Szymczak category

In this appendix, we introduce a class of morphisms, called shift equivalences, and explain its relation to the Szymczak category, following the idea of Franks–Richeson [3]. The results in this appendix are not used in this paper. We include them, however, because they clarify a categorical meaning of the Szymczak category.

We use the language of localization of categories. For basics on this subject, see, e.g. [12, §04VB].

Definition B.1

Let \(\textsf{C}\) be a category.

  1. (1)

    Let \(f :X \rightarrow X\) be an endomorphism in \(\textsf{C}\). Then, we write \(\widehat{f}\) for f seen as a morphism from f to itself in \(\textsf{End}(\textsf{C})\).

  2. (2)

    Let \(f :X \rightarrow X\) and \(g :Y \rightarrow Y\) be two endomorphisms in \(\textsf{C}\). We say that a morphism \(\varphi :f \rightarrow g\) in \(\textsf{End}(\textsf{C})\) is a shift equivalence if there exist a morphism \(\psi :g \rightarrow f\) in \(\textsf{End}(\textsf{C})\) and \(n \in \mathbb {N}\) such that \(\psi \varphi = \widehat{f}^n\) and \(\varphi \psi = \widehat{g}^n\).

We write \(\textsf{ShiftEq}(\textsf{C})\) for the class of all shift equivalences in the category \(\textsf{End}(\textsf{C})\).

Lemma B.2

Let \(\textsf{C}\) be a category. Then, \(\textsf{ShiftEq}(\textsf{C})\) is a saturated multiplicative system in \(\textsf{End}(\textsf{C})\) (in the sense of [12, Defn. 05Q8]).

Proof

Obviously, \(\textsf{ShiftEq}(\textsf{C})\) is closed under composition and contains the isomorphisms. Suppose that we are given a diagram

in \(\textsf{End}(\textsf{C})\) such that \(\varphi \in \textsf{ShiftEq}(\textsf{C})\). Let us take \(\psi :g \rightarrow f\) and \(n \in \mathbb {N}\) such that \(\psi \varphi = \widehat{f}^n\) and \(\varphi \psi = \widehat{g}^n\). Then, the following diagram is commutative, and the right map \(\widehat{h}^n\) is a morphism in \(\textsf{ShiftEq}(\textsf{C})\):

Suppose that \(\varphi , \psi :f \rightrightarrows g\) are two morphisms in \(\textsf{End}(\textsf{C})\) and \(\chi :h \rightarrow f\) is a morphism in \(\textsf{ShiftEq}(\textsf{C})\) such that \(\varphi \chi = \psi \chi \). Let us take \(\omega :f \rightarrow h\) and \(n \in \mathbb {N}\) such that \(\omega \chi = \widehat{h}^n\) and \(\chi \omega = \widehat{f}^n\). Then, \(\widehat{g}^n \varphi = \widehat{g}^n \psi \), and \(\widehat{g}^n\) is a morphism in \(\textsf{ShiftEq}(\textsf{C})\). Thus, \(\textsf{ShiftEq}(\textsf{C})\) is a left multiplicative system. One can see that \(\textsf{ShiftEq}(\textsf{C})\) is a right multiplicative system in the same way.

Let us prove that \(\textsf{ShiftEq}(\textsf{C})\) is saturated. Suppose that we are given a sequence

$$\begin{aligned} f \xrightarrow {\varphi } g \xrightarrow {\psi } h \xrightarrow {\chi } i \end{aligned}$$

in \(\textsf{End}(\textsf{C})\) such that \(\psi \varphi \) and \(\chi \psi \) are morphisms in \(\textsf{ShiftEq}(\textsf{C})\). We want to prove that \(\psi \) is in \(\textsf{ShiftEq}(\textsf{C})\). Take \(\omega :h \rightarrow f\), \(\tau :i \rightarrow g\), and \(m,n \in \mathbb {N}\) such that

$$\begin{aligned} \omega \psi \varphi = \widehat{f}^m, \qquad \psi \varphi \omega = \widehat{h}^m, \qquad \tau \chi \psi = \widehat{g}^n, \qquad \chi \psi \tau = \widehat{i}^n. \end{aligned}$$

Put \(\sigma = \varphi \omega \widehat{h}^n\). Then, we have \(\sigma \psi = \widehat{g}^{m+n}\) and \(\psi \sigma = \widehat{h}^{m+n}\). \(\square \)

Definition B.3

Let \(\textsf{C}\) be a category. We define a functor \(Q :\textsf{End}(\textsf{C}) \rightarrow \textsf{Sz}(\textsf{C})\) as follows:

  • Q is identity on the objects.

  • \(Q\varphi = \overline{(\varphi , 0)}\) for each morphism \(\varphi \) in \(\textsf{End}(\textsf{C})\).

Proposition B.4

Let \(\textsf{C}\) be a category.

  1. (1)

    \((\textsf{Sz}(\textsf{C}), Q)\) is a localization of the category \(\textsf{End}(\textsf{C})\) at the class of morphisms \(\textsf{ShiftEq}(\textsf{C})\).

  2. (2)

    A morphism \(\varphi \) in \(\textsf{End}(\textsf{C})\) is a shift equivalence if and only if \(Q \varphi \) is an isomorphism.

Proof

(1) Let f and g be two endomorphisms in \(\textsf{C}\). Let \(\varphi :f \rightarrow g\) be a shift equivalence. Take \(\psi :g \rightarrow f\) and \(n \in \mathbb {N}\) such that \(\psi \varphi = \widehat{f}^n\) and \(\varphi \psi = \widehat{g}^n\). We see that \(Q\varphi \ (= \overline{(\varphi , 0)})\) is an isomorphism in \(\textsf{Sz}(\textsf{C})\) with inverse \(\overline{(\psi , n)}\).

Let \(\Phi :\textsf{End}(\textsf{C}) \rightarrow \textsf{D}\) be a functor that sends shift equivalences to isomorphisms in \(\textsf{D}\). We want to show that there exists a unique functor \(\overline{\Phi } :\textsf{Sz}(\textsf{C}) \rightarrow \textsf{D}\) such that \(\overline{\Phi }Q = \Phi \). Obviously, such \(\overline{\Phi }\) must be equal to \(\Phi \) on the objects. Notice that, for any morphism \(\varphi :f \rightarrow g\) in \(\textsf{End}(\textsf{C})\) and any \(n \in \mathbb {N}\), we have \(\overline{(\varphi , n)} = Q\varphi \circ (Q \widehat{f})^{-n}\). This means that \(\overline{\Phi }\) must satisfy \(\overline{\Phi }\,\overline{(\varphi , n)} = \Phi \varphi \circ (\Phi \widehat{f})^{-n}\). To the contrary, if we define \(\overline{\Phi }\) in this way, it is indeed a well-defined functor satisfying \(\overline{\Phi }Q = \Phi \).

(2) This follows from (1), Lemma B.2, and [12, Lem. 05Q9]. \(\square \)

Next, let us consider the continuous time case.

Definition B.5

Let \(\textsf{C}\) be either \(\textsf{Set}_*\) or \(\textsf{CHaus}_*\).

  1. (1)

    Let F be an object of \(\textsf{SFlow}(\textsf{C})\). Then, for each \(t \in \mathbb {R}_{\geqslant 0}\), we write \(\widehat{f}^t\) for \(f^t\) seen as a morphism from F to itself in \(\textsf{SFlow}(\textsf{C})\).

  2. (2)

    Let F and G be two objects in \(\textsf{SFlow}(\textsf{C})\). We say that a morphism \(\varphi :F \rightarrow G\) in \(\textsf{SFlow}(\textsf{C})\) is a shift equivalence if there exist a morphism \(\psi :G \rightarrow F\) in \(\textsf{SFlow}(\textsf{C})\) and \(s \in \mathbb {R}_{\geqslant 0}\) such that \(\psi \varphi = \widehat{f}^s\) and \(\varphi \psi = \widehat{g}^s\).

We write \(\textsf{ShiftEq}_\textsf{cont}(\textsf{C})\) for the class of all shift equivalences in the category \(\textsf{SFlow}(\textsf{C})\).

Lemma B.6

Let \(\textsf{C}\) be either \(\textsf{Set}_*\) or \(\textsf{CHaus}_*\). Then, \(\textsf{ShiftEq}_\textsf{cont}(\textsf{C})\) is a saturated multiplicative system in \(\textsf{SFlow}(\textsf{C})\).

Proof

The same as the proof of Lemma B.2. \(\square \)

Definition B.7

Let \(\textsf{C}\) be either \(\textsf{Set}_*\) or \(\textsf{CHaus}_*\). We define a functor \(Q :\textsf{SFlow}(\textsf{C}) \rightarrow \textsf{Sz}_\textsf{cont}(\textsf{C})\) as follows:

  • Q is identity on the objects.

  • \(Q\varphi = \overline{(\varphi , 0)}\) for each morphism \(\varphi \) in \(\textsf{SFlow}(\textsf{C})\).

Proposition B.8

Let \(\textsf{C}\) be either \(\textsf{Set}_*\) or \(\textsf{CHaus}_*\).

  1. (1)

    \((\textsf{Sz}_\textsf{cont}(\textsf{C}), Q)\) is a localization of the category \(\textsf{SFlow}(\textsf{C})\) at the class of morphisms \(\textsf{ShiftEq}_\textsf{cont}(\textsf{C})\).

  2. (2)

    A morphism \(\varphi \) in \(\textsf{SFlow}(\textsf{C})\) is a shift equivalence if and only if \(Q \varphi \) is an isomorphism.

Proof

The same as the proof of Proposition B.4. \(\square \)

Remark B.9

The relation between the Szymczak category and shift equivalences are explained in Franks–Richeson [3]. However, in [3], localization is not used; the Conley index is thus defined as a shift equivalence class in the endomorphism category. In other words, it is defined as an isomorphism class of the objects in the Szymczak category. The Conley index defined in Szymczak [13] carries more information: it is a connected simple system in the Szymczak category, i.e. an object of the Szymczak category defined up to unique isomorphism, which retains the information of automorphisms.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Morita, Y. Conley index theory without index pairs. I: The point-set level theory. J. Fixed Point Theory Appl. 25, 15 (2023). https://doi.org/10.1007/s11784-022-01037-5

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11784-022-01037-5

Navigation