Skip to main content
Log in

Variable Time Step Dynamics with Choice

  • Published:
Journal of Dynamics and Differential Equations Aims and scope Submit manuscript

Abstract

We are interested in systems that can and possibly do switch between different regimes; the durations of staying in each regime (dwell times) may vary. We work with the deterministic picture and study global attractors both in the state space and, one of the novelties, in the hyperspace. We give an example of a simple ODE system where the existence or non-existence of attractors depends on the intervals from which the dwell times are chosen. Another novelty is the description of restricted iterated function systems and their attractors. An unexpected corollary is that for the restricted to a subshift hyperbolic IFS, its attractor is the image under Hutchinson’s map \(p\) of a different, in general, subshift (we call it dual to the original one).

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

References

  1. Akhmerov, R.R., Kamenskiĭ, M.I., Potapov, A.S., Rodkina, A.E., Sadovskiĭ, B.N.: Measures of Noncompactness and Condensing Operators. Operator Theory: Advances and Applications, vol. 55. Birkhäuser Verlag, Basel (1992)

    MATH  Google Scholar 

  2. Andres, J., Fišer, J., Gabor, G., Leśniak, K.: Multivalued fractals. Chaos Solitons Fractals 24(3), 665–700 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  3. Bandt, C.: Fractals in dynamical systems. In: Proceedings of the Conference on Ergodic Theory and Related Topics, II. Teubner-Texte Math., vol. 94 (1986, Georgenthal, pp. 12–23). Teubner, Leipzig (1987)

  4. Bandt, C.: Self-similar sets. I. Topological Markov chains and mixed self-similar sets. Math. Nachr. 142, 107–123 (1989)

    Article  MATH  MathSciNet  Google Scholar 

  5. Bandt, C.: Self-similar sets III. Constructions with sofic systems. Monatsh. Math. 108(2–3), 89–102 (1989)

    Article  MATH  MathSciNet  Google Scholar 

  6. Barnsley, M.F.: Fractals Everywhere, 2nd edn. Academic Press Professional, Boston, MA (1993)

    MATH  Google Scholar 

  7. Barnsley, M.F.: Superfractals. Cambridge University Press, Cambridge (2006)

    Book  MATH  Google Scholar 

  8. Belbruno, E.: Fly Me to the Moon. An Insider’s Guide to the New Science of Space Travel. With a Foreword by Neil deGrasse Tyson. Princeton University Press, Princeton, NJ (2007)

    Google Scholar 

  9. Branicky, M.: Stability of switched and hybrid systems. In: Proceedings of the 33rd Conference on Design and Control, Lake Buena Vista, FL, pp. 3498–3503 (1994)

  10. Branicky, M.: Studies in hybrid systems: modeling, analysis, and control. Sc.D. thesis, MIT (1995)

  11. Branicky, M.: Introduction to hybrid systems. In: Hristu-Varsakelis, D., Levine, W.S. (eds.) Handbook of Networked and Embedded Control Systems, pp. 91–116. Birkhauser, Boston (2005)

    Chapter  Google Scholar 

  12. Cheban, D.N.: Compact global attractors of control systems. J. Dyn. Control Syst. 16(1), 23–44 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  13. Cheban, D.N.: Global Attractors of Set-Valued Dynamical and Control Systems. Nova Science Publishers Inc, New York (2010)

    Google Scholar 

  14. Hutchinson, J.E.: Fractals and self-similarity. Indiana Univ. Math. J. 30(5), 713–747 (1981)

    Article  MATH  MathSciNet  Google Scholar 

  15. Kapitanski, L., Kostin, I.N.: Attractors of nonlinear evolution equations and their approximations. Leningrad Math. J. 2(1), 97–117 (1991). In Russian, Algebra i Analiz 2(1), 114–140 (1990)

  16. Kapitanski, L., Živanović, S.: Dynamics with choice. Nonlinearity 22, 163–186 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  17. Kapitanski, L., Živanović, S.: Dynamics with a range of choice. Reliab. Comput. 15(4), 290–299 (2010)

    Google Scholar 

  18. Kapitanski, L., Živanović Gonzalez, S.: Attractors in hyperspace. Topol. Methods Nonlinear Anal. (2013, to appear)

  19. Kapitanski, L., Živanović Gonzalez, S.: Continuous limit in dynamics with choice. (submitted)

  20. Kitchens, B.P.: Symbolic dynamics. One-sided, two-sided and countable state Markov shifts, vol. Universitext. Springer-Verlag, Berlin (1998)

    MATH  Google Scholar 

  21. Kloeden, P.E.: Nonautonomous attractors of switching systems. Dyn. Syst. 21(2), 209–230 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  22. Kuratowski, K.: Topology, vol. I. Polish Scientific Publishers, New York (1966)

    Google Scholar 

  23. Ladyzhenskaya, O.A.: The Mathematical Theory of Viscous Incompressible Flow, 2 English edn. Gordon and Beach, New York (1969)

    MATH  Google Scholar 

  24. Ladyzhenskaya, O.A.: On a dynamical system generated by Navier–Stokes equations, Boundary-value problems of mathematical physics and related problems of function theory. J. Soviet Math. 3(4), 458–479 (1975). (Part 6, Zap. Nauchn. Sem. LOMI, 27, “Nauka”, Leningrad. Otdel., Leningrad, pp. 91–115 (1972))

  25. Ladyzhenskaya, O.A.: Attractors for Semigroups and Evolution Equations, Lezioni Lincee. [Lincei Lectures] Cambridge University Press, Cambridge (1991)

  26. Leśniak, K.: Stability and invariance of multivalued iterated function systems. Math. Slovaca 53(4), 393–405 (2003)

    MATH  MathSciNet  Google Scholar 

  27. Liberzon, D.: Switching in Systems and Control. Systems & Control: Foundations & Applications. Birkhäuser Boston Inc., Boston, MA (2003)

  28. Lind, D., Marcus, B.: An Introduction to Symbolic Dynamics and Coding. Cambridge University Press, Cambridge (1995)

    Book  MATH  Google Scholar 

  29. Lions, J.-L.: Quelques Méthodes de Résolution des Problèmes aux Limites Non linéaires. Dunod, Paris (1969)

    MATH  Google Scholar 

  30. Liu, X., Stechlinski, P.: Pulse and constant control schemes for epidemic models with seasonality. Nonlinear Anal. 12, 931–946 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  31. Mallet-Paret, John, Nussbaum, Roger D.: Inequivalent measures of noncompactness. Ann. Math. Pura Appl. (4) 190(3), 453–488 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  32. Margaliot, M.: Stability analysis of switched systems using variational principles: an introduction. Autom. J. IFAC 42(12), 2059–2077 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  33. Martinón, A.: A system of axioms for measures of noncompactness. Extr. Math. 4(1), 42–44 (1989)

    Google Scholar 

  34. Massatt, P.: Some properties of condensing maps. Ann. Math. Pura Appl. 125(4), 101–115 (1980)

    Article  MATH  MathSciNet  Google Scholar 

  35. Melnik, V.S., Valero, J.: On attractors of multi-valued semi-flows and differential inclusions. Set-Valued Anal. 6, 83–111; (2008). (Addendum to: On attractors of multivalued semiflows and differential inclusions [Set-Valued Anal. 6(1), 1998, pp. 83–111]. Set-Valued Anal. 16(4), 507–509 (1998))

  36. Sell, G.R.: Nonautonomous differential equations and topological dynamics, I, II. Trans. Am. Math. Soc. 127(241–262), 263–283 (1967)

    Article  MathSciNet  Google Scholar 

  37. Sell, G.R., You, Y.: Dynamics of Evolutionary Equations. Applied Mathematical Sciences, vol. 143. Springer-Verlag, New York (2002)

    Book  Google Scholar 

  38. Živanović, S.: Attractors in dynamics with choice. Ph.D. Thesis, University of Miami, Coral Gables, FL (2009)

Download references

Acknowledgments

This work has been completed during the fall 2012 program “Hamiltonians in Magnetic Field” at Institut Mittag-Leffler. LK thanks the organizers of the program and the leadership and staff of Institut Mittag-Leffler for their kind support and hospitality.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lev Kapitanski.

Appendix

Appendix

In this section we collect a few useful facts about the space \(X^\sharp \) and convergence of sets in \(X\) and in \(X^\sharp \). Also, we present some results on the global attractors of systems arising from iterations of a single multi-valued map.

The notation convention is the same as in Sect. 3. As in the main part of the paper, the space \((X, d)\) is complete. Then \((X^\sharp , d^\sharp )\) is also complete, [22, §33, IV].

Lemma 26

  1. (1)

    \(d^\sharp (A, B) = d^\sharp (A, \overline{B})\); if \(d^\sharp (A, B) = 0\), then \(\overline{A} = \overline{B}\).

  2. (2)

    If \(B\) is a nonempty, closed, bounded set in \(X\), then \(B^\sharp \), the family of all nonempty, closed subsets of \(B\), is a bounded set in \(X^\sharp \) (it need not be closed).

  3. (3)

    If \({\fancyscript{A}}\) is a nonempty, bounded subset of \(X^\sharp \), then \({\fancyscript{A}}^\flat = \bigcup _{C\in {\fancyscript{A}}} C\) is a bounded subset of \(X\).

  4. (4)

    If \(A\subset X\) is compact, then \(A^\sharp \) is compact in \(X^\sharp \).

The proof is an easy exercise. \(\square \)

  • We say that a set \(A\) attracts the sets \(A_n\) if \(\lim _n e(A_n, A) = 0\).

The following observation proves to be quite useful.

Lemma 27

Let \(A\subset X\) be a compact set. Let \(B_n\) be a sequence of sets attracted by \(A\). Then the sequence \(\overline{B_n}\) has a subsequence converging in the Hausdorff metric. Its limit is a compact subset of \(A\).

Proof

Choose a sequence \(n_k\nearrow +\infty \) such that \(\overline{B_{n_k}}\subset \mathcal{O}_{2^{-k}}(A)\). Because \(A\) is compact, we can define the sets

$$\begin{aligned} A_k = \{y\in A\,:\, \exists x\in B_{n_k} \;\hbox {such that}\; d(x, y) = d(x,A)\}. \end{aligned}$$

Obviously, \(d^\sharp (B_{n_k},A_k)\le 2^{-k}\). Since \(A^\sharp \) is compact in \(X^\sharp \), the sequence \(\overline{A_k}\) has a convergent subsequence with a limit in \(A^\sharp \). The corresponding subsequence of \(B_{n_k}\) will have the same limit. \(\square \)

The limit of a Cauchy sequence, \(A_n\), in the hyperspace \(X^\sharp \) can be described with the help of the formula due to Hausdorff, [22, §29, IV(8)]:

$$\begin{aligned} A_n\rightarrow {A} = \bigcap _n \overline{\bigcup _{m\ge n} A_m} = \{x\in X: \,\exists n_k\nearrow \infty \,\exists x_k\in A_{n_k} \;\text {such that}\; x = \lim x_k\}. \end{aligned}$$
(66)

The basic convergence results stem from a generalization of the Cantor nested intervals theorem. The following theorem is a version of the Kuratowski generalization of Cantor’s result, [22, §34]. The first part is in [34, Theorem 3.1] and the second follows from [26, Prop. 3(viii)].

Theorem 28

Assume \(A_1\supset A_2\supset \dots \) is a nested sequence of nonempty, closed, bounded subsets in \(X\). Assume that \(\psi (A_n)\rightarrow 0\) for some mnc \(\psi \). Then,

  1. (1)

    the set \(A = \bigcap _n A_n\) is nonempty and compact;

  2. (2)

    \(A_n\rightarrow A\) in \(X^\sharp \).

Combining this with Lemma 27, we obtain the following corollary.

Corollary 29

Suppose \(A_1\supset A_2\supset \dots \) is a nested sequence of non-empty, closed, and bounded sets, and assume that \(\lim _n\psi (A_n) = 0\). Then any sequence of closed subsets \(B_n\subset A_n\) has a convergent in the Hausdorff metric subsequence (with the limit a subset of \(A\)).

The implications of convergence to a compact set are as follows.

Lemma 30

Assume \(A_n\rightarrow A\) in \(X^\sharp \) and \(A\) is compact. Then

  1. (1)

    any sequence \(C_k\) of nonempty, closed subsets of \(A_{n_k}\), where \(n_k\nearrow \infty \), has a convergent in \(X^\sharp \) subsequence, and its limit is a compact subset of \(A\);

  2. (2)

    \( \psi (A_n) \rightarrow 0\,; \)

  3. (3)

    \( \overline{\bigcup _{m\ge n} A_m} \rightarrow A\,; \)

  4. (4)

    \( \psi (\overline{\bigcup _{m\ge n} A_m})\rightarrow 0. \)

Proof

The first statement follows from Lemma 27. For the second observe that \( \psi (A_n) = \psi (A_n) - \psi (A) \le c_\psi d^\sharp (A_n, A)\rightarrow 0\). The third statement is a consequence of the first and formula (66). Finally, the fourth statement follows from the second. \(\square \)

It is convenient to have a different description of the condition \(\lim _n\psi (A_n) = 0\).

Lemma 31

Let \(A_1\supset A_2\supset \dots \) be a nested sequence of nonempty, closed, and bounded sets. The following conditions are equivalent:

  1. (1)

    \(\lim _n\psi (A_n) = 0\).

  2. (2)

    Every sequence \(x_k\), where \(x_k\in A_{n_k}\), \(n_k\nearrow +\infty \), has a convergent subsequence.

Proof

Let the first condition be satisfied. For any \(n\), \(\{x_k\}\subset A_n\) except for at most a finite number of \(x_k\)’s. Hence, \(\psi (\{x_k\})\le \psi (A_n)\), and therefore \(\psi (\{x_k\})=0\), which implies the second condition.

Now suppose the second condition is satisfied. Note that if a set \(C\) has a finite \(\epsilon \)-net \(Q\subset C\), then \(d^\sharp (C, Q)\le \epsilon \), and therefore, by the property mnc(iv),

$$\begin{aligned} \psi (C) \le c_\psi \;\epsilon . \end{aligned}$$

If \(\psi (A_n)\) does not converge to \(0\), then there is subsequence \(n_k\) and an \(\epsilon > 0\) such that

$$\begin{aligned} \psi (A_{n_k}) > c_\psi \,\epsilon . \end{aligned}$$

Hence, the sets \(A_{n_k}\) do not have finite \(\epsilon \)-nets. Take a point \(x_1\in A_{n_1}\). There exists \(x_2\in A_{n_2}\) such that \(d(x_1, x_2) > \epsilon /2\). There exists \(x_3\in A_{n_3}\) such that \(d(x_3, \{x_1, x_2\}) > \epsilon /2\), and so on. The sequence \(\{x_n\}\) is not totally bounded. However, by construction, \(x_k \in A_{n_k}\), and by assumption, the sequence \(\{x_k\}\) must have a convergent subsequence. A contradiction. \(\square \)

Consider now a map \(\Phi \) defined on nonempty bounded subsets of \(X\) with values also bounded subsets of \(X\). The first two assumptions on \(\Phi \) are these:

  • \(\mathbf \Phi 1\): \(\Phi \) maps \(X^\sharp \) into itself.

  • \(\mathbf \Phi 2\): If \(A\subset B\), then \(\Phi (A)\subset \Phi (B)\).

We do not assume that \(\Phi \) is a multi-valued map in the traditional sense. In our considerations the set \(\Phi (A)\) may be larger than the union of the sets \(\Phi (x)\) over all \(x\in A\) (where \(\Phi (x)\) is \(\Phi (\{x\})\), of course).

The iterations \(\Phi ^n\) satisfy the conditions of Definition 17. They generate the dynamics of sets on \(X\) and on \(X^\sharp \). These are the systems \(\left( \{\Phi ^n\}, X\right) \) and \(\left( \{\Phi ^n\}, X^\sharp \right) \) in the notation of Sect. 5. We are interested in the global attractors of these systems. Here is a corollary of Theorems 18, 20, and Lemma 31.

Theorem 32

  1. (1)

    For the system \(\left( \{\Phi ^n\}, X\right) \) to possess a global attractor it is necessary and sufficient that the following two conditions hold.

    1. (a)

      There is a bounded global absorbing set;

    2. (b)

      For some mnc \(\psi \), \(\psi (\Phi ^n(A))\rightarrow 0\) for every bounded set \(A\).

  2. (2)

    If both conditions (a) and (b) are satisfied, the system \(\left( \{\Phi ^n\}, X^\sharp \right) \) possesses a global attractor. If \(K\) is the global attractor of \(\left( \{\Phi ^n\}, X\right) \) and \({\fancyscript{K}}\) is the global attractor of \(\left( \{\Phi ^n\}, X^\sharp \right) \), then \({\fancyscript{K}}^\flat = K\).

  3. (3)

    Assume both conditions (a) and (b) are satisfied. Let \(B\) be a bounded global absorbing set. Then

In the third statement, the formula for \(K\) follows from [6, Corollary 19], and the formula for \({\fancyscript{K}}\) uses a similar argument.

We call the maps \(\Phi \) satisfying the condition (b) asymptotically \(\psi \)-condensing. If \(\Phi \) originates from a map \(\Phi : X\rightarrow 2^X\) so that \(\Phi (A) = \bigcup _{x\in A} \Phi (x)\), then \(\Phi \) is asymptotically \(\psi \)-condensing if it is \(\psi \)-condensing (this requires a proof, see [34]). Recall that \(\Phi \) is \(\psi \)-condensing with respect to some mnc \(\psi \) with the properties mnc(i)–(iv) iff \(\psi (\Phi (B))\le \psi (B)\) for any bounded \(B\), and the inequality is strict if \(\psi (B) > 0\). The condition \(\Phi (A) = \bigcup _{x\in A} \Phi (x)\) can be weakened as follows, see [34, Theorems 4.1 and 4.2]. Consider the following two properties (“compact sets” in \(\mathbf{\Phi 4a}\) are replaced by “finite sets” in \(\mathbf{\Phi 4b}\)).

\(\mathbf \Phi 4a(b)\)::

For every \(D\in X^\sharp \), there is a sequence of sets \(B_n\subset \Phi ^n(D)\), \(n = 1, 2,\dots \), such that

(1):

\(d^\sharp (B_n, \Phi ^n(D))\rightarrow 0\);

(2):

for \(n = 1, 2,\dots \), for every compact set (\(\mathbf{b:}\) finite set) \(L\in B_n\) there is a compact set (\(\mathbf{b:}\) finite set) \(L^\prime \subset B_{n-1}\) such that

$$\begin{aligned} L\subset \Phi \left( L^\prime \right) . \end{aligned}$$
(67)

[The reason assumption \(\mathbf{\Phi 4}\) precedes \(\mathbf{\Phi 3}\) is that we would like to have the same names for the properties of \(\Phi \) here as in [18].] An example of a map that satisfies \(\mathbf{\Phi 4b}\) is the Hutchinson-Barnsley map (13). The following result is due essentially to Massatt, [34].

Proposition 33

Let \(\Phi \) satisfy the assumptions \(\mathbf{\Phi 1}\), \(\mathbf{\Phi 2}\), and \(\mathbf{\Phi 4a}\) or \(\mathbf{\Phi 4b}\). In addition, let there be a bounded global absorbing set for the iterations of \(\Phi \) in \(X\). Then, if \(\Phi \) is \(\psi \)-condensing, it is asymptotically \(\psi \)-condensing. As a result, both \(\left( \{\Phi ^n\}, X\right) \) and \(\left( \{\Phi ^n\}, X^\sharp \right) \) have global attractors.

It is easier to check that a map is condensing than that it is asymptotically condensing. For example, \(\Phi \) is condensing if it is compact (i.e., it maps bounded sets into relatively compact sets). For another example, assume \(\Phi (A) = \bigcup _{x\in A}S(x)\) and \(S\) is a single-valued map. Then \(\Phi \) is condensing if \(S\) is a contraction, or if \(S\) is compact. If \(X\) is a Banach space and \(\psi \) is the Kuratowski, or the Hausdorff, or any other mnc with the additional property that \(\psi (A + B) \le \psi (A) + \psi (B)\) for any pair of bounded sets, then if \(S\) is a sum of a finite number of \(\psi \)-condensing maps, \(\Phi \) is \(\psi \)-condensing.

The invariance of attractors \(K\) and \({\fancyscript{K}}\) does not come free, some form of continuity of \(\Phi \) is needed. For example, the following assumption will do:

\(\mathbf{\Phi 3a:}\) If \(A_k, A\in X^\sharp \) and \(A\) is compact, then

$$\begin{aligned} A_k\rightarrow A \quad \implies \quad \Phi (A_k)\rightarrow \Phi (A). \end{aligned}$$

Proposition 34

Assume \(\left( \{\Phi ^n\}, X\right) \) has a global attractor. Let \(K\) be this attractor and let \({\fancyscript{K}}\) be the global attractor of \(\left( \{\Phi ^n\}, X^\sharp \right) \). Assume \(\Phi \) has the property \(\mathbf{\Phi 3a}\). Then \(K = \Phi (K)\) and \({\fancyscript{K}} = \Phi ({\fancyscript{K}})\). This, in turn, implies that \(K\) is the union in \(X\) of bounded two-sided set-trajectories of \(\Phi \). [A two-sided trajectory is a sequence of bounded, closed sets \(A_n\), \(n\in \mathbb Z\), such that \(\Phi (A_n) = A_{n+1}\). It is bounded if \(\{A_n\}\) is bounded in \(X^\sharp \).] The attractor \({\fancyscript{K}}\) is the union (in \(X^\sharp \)) of bounded two-sided point-trajectories of \(\Phi \). Also, \(K\) is the maximal compact \(\Phi \)-invariant subset of \(X\) and \({\fancyscript{K}}\) is the maximal \(\Phi \)-invariant compact subset of \(X^\sharp \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kapitanski, L., Živanović Gonzalez, S. Variable Time Step Dynamics with Choice. J Dyn Diff Equat 26, 745–779 (2014). https://doi.org/10.1007/s10884-014-9374-1

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10884-014-9374-1

Keywords

Navigation