Skip to main content
Log in

Topological entropy of switched linear systems: general matrices and matrices with commutation relations

  • Original Article
  • Published:
Mathematics of Control, Signals, and Systems Aims and scope Submit manuscript

Abstract

This paper studies a notion of topological entropy for switched systems, formulated in terms of the minimal number of trajectories needed to approximate all trajectories with a finite precision. For general switched linear systems, we prove that the topological entropy is independent of the set of initial states. We construct an upper bound for the topological entropy in terms of an average of the measures of system matrices of individual modes, weighted by their corresponding active times, and a lower bound in terms of an active-time-weighted average of their traces. For switched linear systems with scalar-valued state and those with pairwise commuting matrices, we establish formulae for the topological entropy in terms of active-time-weighted averages of the eigenvalues of system matrices of individual modes. For the more general case with simultaneously triangularizable matrices, we construct upper bounds for the topological entropy that only depend on the eigenvalues, their order in a simultaneous triangularization, and the active times. In each case above, we also establish upper bounds that are more conservative but require less information on the system matrices or on the switching, with their relations illustrated by numerical examples. Stability conditions inspired by the upper bounds for the topological entropy are presented as well.

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.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. In information theory, entropy notions are often formulated using binary logarithms due to their connection with binary signals. In this paper, we use natural logarithms to avoid generating additional multiplicative constants \( \ln 2 \) when computing the topological entropy.

  2. In all examples, we denote by \( t_1< t_2 < \cdots \) the sequence of switches and let \( t_0 := 0 \), with \( \sigma = 1 \) on \( [t_{2k}, t_{2k+1}) \) and \( \sigma = 2 \) on \( [t_{2k+1}, t_{2k+2}) \).

  3. We can apply [11, Th. 27, p. 34] to the switched linear system (14) as the switching signal \( \sigma \) is piecewise constant.

  4. Following (23), the upper bound (24) is tighter than the one in the previous result [43, eq. (19)].

  5. In particular, for each \( p \in \mathcal P\), the diagonal entries \( a_p^i \) are also the eigenvalues of the original system matrix \( A_p \) [17, Th. 2.4.8.1, p. 117].

  6. For example, when the switching signal \( \sigma \) is periodic; see [36, Sec. 3.2.1] for more conditions.

  7. A sufficient condition for simultaneous triangularizability is that the matrices \( A_p \) are pairwise commuting (see, e.g., [17, Th. 2.3.3, p. 103]). More sufficient conditions can be found in [25]. A necessary and sufficient condition is that the Lie algebra \( \{A_p: p \in \mathcal P\}_{LA} \) is solvable (see, e.g., [18, pp. 10, 16]). More necessary and sufficient conditions can be found in [13, 31].

References

  1. Adler RL, Konheim AG, McAndrew MH (1965) Topological entropy. Trans Am Math Soc 114(2):309–319

    Article  MathSciNet  Google Scholar 

  2. Agrachev AA, Liberzon D (2001) Lie-algebraic stability criteria for switched systems. SIAM J Control Optim 40(1):253–269

    Article  MathSciNet  Google Scholar 

  3. Agrachev AA, Baryshnikov Y, Liberzon D (2012) On robust Lie-algebraic stability conditions for switched linear systems. Syst Control Lett 61(2):347–353

    Article  MathSciNet  Google Scholar 

  4. Bowen R (1971a) Entropy for group endomorphisms and homogeneous spaces. Trans Am Math Soc 153:401–414

    Article  MathSciNet  Google Scholar 

  5. Bowen R (1971b) Periodic points and measures for axiom A diffeomorphisms. Trans Am Math Soc 154:377–397

    MathSciNet  MATH  Google Scholar 

  6. Chicone C (2006) Ordinary differential equations with applications, vol 34, 2nd edn. Springer, Berlin

    MATH  Google Scholar 

  7. Colonius F (2012) Minimal bit rates and entropy for exponential stabilization. SIAM J Control Optim 50(5):2988–3010

    Article  MathSciNet  Google Scholar 

  8. Colonius F, Kawan C (2009) Invariance entropy for control systems. SIAM J Control Optim 48(3):1701–1721

    Article  MathSciNet  Google Scholar 

  9. Colonius F, Kawan C, Nair GN (2013) A note on topological feedback entropy and invariance entropy. Syst Control Lett 62(5):377–381

    Article  MathSciNet  Google Scholar 

  10. Cover TM, Thomas JA (2005) Elements of information theory, 2nd edn. Wiley, Hoboken

    Book  Google Scholar 

  11. Desoer CA, Vidyasagar M (2009) Feedback systems: input–output properties. SIAM, Philadelphia

    Book  Google Scholar 

  12. Dinaburg EI (1970) The relation between topological entropy and metric entropy. Dokl Akad Nauk SSSR 190(1):19–22 in Russian

    MathSciNet  MATH  Google Scholar 

  13. Drazin MP, Dungey JW, Gruenberg KW (1951) Some theorems on commutative matrices. J Lond Math Soc 26(3):221–228

    Article  MathSciNet  Google Scholar 

  14. Gurvits L (1995) Stability of discrete linear inclusion. Linear Algebra Appl 231:47–85

    Article  MathSciNet  Google Scholar 

  15. Haimovich H, Braslavsky JH (2013) Sufficient conditions for generic feedback stabilizability of switching systems via Lie-algebraic solvability. IEEE Trans Autom Control 58(3):814–820

    Article  MathSciNet  Google Scholar 

  16. Hespanha JP, Ortega A, Vasudevan L (2002) Towards the control of linear systems with minimum bit-rate. In: 15th international symposium on mathematical theory of networks and systems, pp 1–15

  17. Horn RA, Johnson CR (2013) Matrix analysis, 2nd edn. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  18. Humphreys JE (1972) Introduction to lie algebras and representation theory. Springer, Berlin

    Book  Google Scholar 

  19. Katok A, Hasselblatt B (1995) Introduction to the modern theory of dynamical systems. Cambridge University Press, Cambridge

    Book  Google Scholar 

  20. Kawan C, Latushkin Y (2016) Some results on the entropy of non-autonomous dynamical systems. Dyn Syst 31(3):251–279

    Article  MathSciNet  Google Scholar 

  21. Kolmogorov AN (1958) A new metric invariant of transitive dynamical systems and automorphisms of Lebesgue spaces. Dokl Akad Nauk SSSR 119(5):861–864 in Russian

    MathSciNet  MATH  Google Scholar 

  22. Kolyada S, Snoha L (1996) Topological entropy of nonautonomous dynamical systems. Random Comput Dyn 4(2–3):205–233

    MathSciNet  MATH  Google Scholar 

  23. Kutepov SA (1982) Algebraic criterion of absolute stability. Kibernetika i Vychislitel’naia Tekhnika 54:52–55 in Russian

    MathSciNet  MATH  Google Scholar 

  24. Kutepov SA (1984) Absolute stability of bilinear systems with a compact Levi factor. Kibernetika i Vychislitel’naia Tekhnika 62(118):28–33 in Russian

    MathSciNet  MATH  Google Scholar 

  25. Laffey TJ (1978) Simultaneous triangularization of matrices-low rank cases and the nonderogatory case. Linear Multilinear Algebra 6(4):269–305

    Article  MathSciNet  Google Scholar 

  26. Liberzon D (2003) Switching in systems and control. Birkhäuser, Basel

    Book  Google Scholar 

  27. Liberzon D (2014) Finite data-rate feedback stabilization of switched and hybrid linear systems. Automatica 50(2):409–420

    Article  MathSciNet  Google Scholar 

  28. Liberzon D, Mitra S (2018) Entropy and minimal bit rates for state estimation and model detection. IEEE Trans Autom Control 63(10):3330–3344

    Article  MathSciNet  Google Scholar 

  29. Liberzon D, Hespanha JP, Morse AS (1999) Stability of switched systems: a Lie-algebraic condition. Syst Control Lett 37(3):117–122

    Article  MathSciNet  Google Scholar 

  30. Matveev AS, Pogromsky AY (2016) Observation of nonlinear systems via finite capacity channels: constructive data rate limits. Automatica 70:217–229

    Article  MathSciNet  Google Scholar 

  31. McCoy NH (1936) On the characteristic roots of matric polynomials. Bull Am Math Soc 42(8):592–601

    Article  MathSciNet  Google Scholar 

  32. Nair GN, Evans RJ (2003) Exponential stabilisability of finite-dimensional linear systems with limited data rates. Automatica 39(4):585–593

    Article  MathSciNet  Google Scholar 

  33. Nair GN, Evans RJ, Mareels IMY, Moran W (2004) Topological feedback entropy and nonlinear stabilization. IEEE Trans Autom Control 49(9):1585–1597

    Article  MathSciNet  Google Scholar 

  34. Ornstein DS (1974) Ergodic theory, randomness, and dynamical systems. Yale University Press, London

    MATH  Google Scholar 

  35. Savkin AV (2006) Analysis and synthesis of networked control systems: topological entropy, observability, robustness and optimal control. Automatica 42(1):51–62

    Article  MathSciNet  Google Scholar 

  36. Schmidt AJ (2016) Topological entropy bounds for switched linear systems with Lie structure. M.S. thesis, University of Illinois at Urbana-Champaign

  37. Shannon CE (1948) A mathematical theory of communication. Bell Syst Tech J 27(3):379–423

    Article  MathSciNet  Google Scholar 

  38. Shorten R, Wirth FR, Mason O, Wulff K, King C (2007) Stability criteria for switched and hybrid systems. SIAM Rev 49(4):545–592

    Article  MathSciNet  Google Scholar 

  39. Sibai H, Mitra S (2017) Optimal data rate for state estimation of switched nonlinear systems. In: 20th ACM international conference on hybrid systems: computation and control, pp 71–80

  40. Tatikonda S, Mitter S (2004) Control under communication constraints. IEEE Trans Autom Control 49(7):1056–1068

    Article  MathSciNet  Google Scholar 

  41. Yang G, Hespanha JP (2018) On topological entropy of switched linear systems with pairwise commuting matrices. In: 56th annual Allerton conference on communication, control, and computing, pp 429–436

  42. Yang G, Liberzon D (2018) Feedback stabilization of a switched linear system with an unknown disturbance under data-rate constraints. IEEE Trans Autom Control 63(7):2107–2122

    Article  Google Scholar 

  43. Yang G, Schmidt AJ, Liberzon D (2018) On topological entropy of switched linear systems with diagonal, triangular, and general matrices. In: 57th IEEE conference on decision and control, pp 5682–5687

  44. Yang G, Hespanha JP, Liberzon D (2019) On topological entropy and stability of switched linear systems. In: 22nd ACM international conference on hybrid systems: computation and control, pp 119–127

Download references

Acknowledgements

The work of G. Yang and J. P. Hespanha was supported by the Office of Naval Research under the MURI Grant N00014-16-1-2710, and by the National Science Foundation under the Grants CNS-1329650 and EPCN-1608880. The work of D. Liberzon was supported by the National Science Foundation under the Grant CMMI-1662708, and by the Air Force Office of Scientific Research under the Grant FA9550-17-1-0236. The authors thank Raphaël M. Jungers for his comments on a preliminary version of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Guosong Yang.

Additional information

Publisher's Note

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

Appendices

A Proof of Lemma 1

As a direct consequence of the definition (12), the maximal weighted average \( \bar{a} \) satisfies

$$\begin{aligned} \bar{a}(T) \ge \max \Bigg \{ \dfrac{1}{T} \sum _{p \in \mathcal P} a_p \tau _p(T),\, 0 \Bigg \} = \max \Bigg \{ \sum _{p \in \mathcal P} a_p \rho _p(T),\, 0 \Bigg \} \qquad \forall T > 0, \end{aligned}$$

and thus \( \limsup _{T \rightarrow \infty }\, \bar{a}(T) \ge \max \{\hat{a},\, 0\} \).

It remains to prove that the reverse inequality holds as well. The definition (11) of the asymptotic weighted average \( \hat{a} \) implies that for each \( \delta > 0 \), there is a large enough \( T_\delta ' \ge 0 \) such that \( \sum _{p \in \mathcal P} a_p \rho _p(t) < \hat{a} + \delta \) for all \( t > T_\delta ' \). For a \( T > T_\delta ' \), let

$$\begin{aligned} t^*(T) := \mathrm {arg\,max}_{t \in [0, T]} \sum _{p \in \mathcal P} a_p \tau _p(t). \end{aligned}$$

Then, \( \sum _{p \in \mathcal P} a_p \tau _p(t^*(T)) \ge 0 \). If \( t^*(T) \in (T_\delta ', T] \), then

$$\begin{aligned} \begin{aligned} \bar{a}(T)&= \dfrac{1}{T} \sum _{p \in \mathcal P} a_p \tau _p(t^*(T)) \le \dfrac{1}{t^*(T)} \sum _{p \in \mathcal P} a_p \tau _p(t^*(T)) \\&= \sum _{p \in \mathcal P} a_p \rho _p(t^*(T)) < \hat{a} + \delta . \end{aligned} \end{aligned}$$

Otherwise \( t^*(T) \in [0, T_\delta '] \), and thus

$$\begin{aligned} \bar{a}(T) = \dfrac{1}{T} \sum _{p \in \mathcal P} a_p \tau _p(t^*(T)) \le \dfrac{a_m t^*(T)}{T} \le \dfrac{\max \{a_m, 0\}\, T_\delta '}{T} \end{aligned}$$

with the constant \( a_m := \max _{p \in \mathcal P} a_p \). Combining the two cases above yields

$$\begin{aligned} \bar{a}(T) \le \max \{\hat{a} + \delta ,\, \max \{a_m, 0\}\, T_\delta '/T\} \qquad \forall T > T_\delta '. \end{aligned}$$

Hence

$$\begin{aligned} \bar{a}(T) \le \max \{\hat{a},\, 0\} + \delta \qquad \forall T > T_\delta := \max \{T_\delta ',\, \max \{a_m, 0\}\, T_\delta '/\delta \}. \end{aligned}$$

As \( \delta > 0 \) is arbitrary, we have \( \limsup _{T \rightarrow \infty }\, \bar{a}(T) \le \max \{\hat{a},\, 0\} \).

B Proof of Lemma 2

  1. 1.

    As \( R(\hat{x}) \subset B_{A_\sigma }(\hat{x}, \varepsilon , T) \) for all \( \hat{x} \in G(\theta ) \), we have

    $$\begin{aligned} K = \bigcup _{\hat{x} \in G(\theta )} R(\hat{x}) \subset \bigcup _{\hat{x} \in G(\theta )} B_{A_\sigma }(\hat{x}, \varepsilon , T). \end{aligned}$$

    Then, (4) implies that the grid \( G(\theta ) \) is \( (T, \varepsilon ) \)-spanning, and thus

    $$\begin{aligned} \log S(A_\sigma , \varepsilon , T, K) \le \log |G(\theta )| = \sum _{i=1}^{n} \log (2 \lfloor 1/\theta _i \rfloor + 1) \le \sum _{i=1}^{n} \log (2/\theta _i + 1). \end{aligned}$$

    Consequently, the definition of entropy (5) implies

    $$\begin{aligned} h(A_\sigma )\le & {} \lim _{\varepsilon \searrow 0} \limsup _{T \rightarrow \infty } \sum _{i=1}^{n} \dfrac{\log (2/\theta _i + 1)}{T} \\= & {} \lim _{\varepsilon \searrow 0} \limsup _{T \rightarrow \infty } \sum _{i=1}^{n} \dfrac{\log (1/\theta _i)}{T}+ \lim _{\varepsilon \searrow 0} \limsup _{T \rightarrow \infty } \sum _{i=1}^{n} \dfrac{\log (2 + \theta _i)}{T}, \end{aligned}$$

    where the last term equals 0 if (17) holds.

  2. 2.

    For all distinct points \( \hat{x}, \hat{x}' \in G(\theta ) \), as \( \hat{x}' \notin R(\hat{x}) \) and \( B_{A_\sigma }(\hat{x}, \varepsilon , T) \subset R(\hat{x}) \), we have \( \hat{x}' \notin B_{A_\sigma }(\hat{x}, \varepsilon , T) \). Then, (6) implies that the grid \( G(\theta ) \) is \( (T, \varepsilon ) \)-separated, and thus

    $$\begin{aligned}&\log N(A_\sigma , \varepsilon , T, K) \ge \log |G(\theta )| \\&= \sum _{i=1}^{n} \log (2 \lfloor 1/\theta _i \rfloor + 1) > \sum _{i=1}^{n} \log (\max \{2/\theta _i - 1,\, 1\}). \end{aligned}$$

    Consequently, the definition of entropy (5) implies

    $$\begin{aligned} \begin{aligned} h(A_\sigma )&\ge \lim _{\varepsilon \searrow 0} \limsup _{T \rightarrow \infty } \sum _{i=1}^{n} \dfrac{\log (\max \{2/\theta _i - 1,\, 1\})}{T} \\&= \lim _{\varepsilon \searrow 0} \limsup _{T \rightarrow \infty } \sum _{i=1}^{n} \dfrac{\log (1/\theta _i)}{T} + \lim _{\varepsilon \searrow 0} \limsup _{T \rightarrow \infty } \sum _{i=1}^{n} \dfrac{\log (\max \{2 - \theta _i,\, \theta _i\})}{T}, \end{aligned} \end{aligned}$$

    where the last term equals 0 if (17) holds.

C Proof of Corollary 2

First, the definition (11) of \( \hat{a} \) and the subadditivity of limit suprema imply

$$\begin{aligned} \hat{a} \le \sum _{p \in \mathcal P} \limsup _{t \rightarrow \infty }\, a_p \rho _p(t) \le \sum _{p \in \mathcal P} \max \{a_p,\, 0\} \limsup _{t \rightarrow \infty } \rho _p(t) = \sum _{p \in \mathcal P} h(a_p) \hat{\rho }_p; \end{aligned}$$

which, combined with (28), implies the upper bound (29). For the case where the limits \( \lim _{t \rightarrow \infty } \rho _p(t) \) exist and \( a_p \ge 0 \) for all \( p \in \mathcal P\), the inequalities in the derivation above becomes equalities due to the additivity of limits and \( \max \{a_p,\, 0\} = a_p \). Second, the definition (11) of \( \hat{a} \) implies

$$\begin{aligned} \hat{a} \le \limsup _{t \rightarrow \infty } \bigg ( \max _{p \in \mathcal P}\, a_p \bigg ) \sum _{p \in \mathcal P} \rho _p(t) = \max _{p \in \mathcal P}\, a_p \le \max _{p \in \mathcal P}\, h(a_p), \end{aligned}$$

which, combined with (28), implies the upper bound (30).

D Proof of Lemma 6

First, we establish the upper bound in (36). For each \( p \in \mathcal P\), as \( N_p \) is nilpotent, there is a positive integer \( k_p \) such that \( N_p^{k_p} = 0 \). Let \( k_s := \sum _{p \in \mathcal P} k_p \), which is finite as the index set \( \mathcal P\) is finite. Define the weighted average matrix over [0, t] by

$$\begin{aligned} N(t) := \sum _{p \in \mathcal P} N_p \rho _p(t) \in \mathbb C^{n \times n}. \end{aligned}$$

For all \( t \ge 0 \), as \( \{N_p: p \in \mathcal P\} \) is a commuting family, we have \( (N(t))^{k_s} = 0 \). Also, \( \Vert N(t)\Vert \le \eta _M := \max _{p \in \mathcal P} \Vert N_p\Vert \). Hence for all \( v \in \mathbb C^n \), we have

$$\begin{aligned} \big \Vert e^{N(t)\, t} v \big \Vert= & {} \Bigg \Vert \left( \sum _{k=0}^{k_s-1} \dfrac{(N(t))^k t^k}{k!} \right) v \Bigg \Vert \le \left( \sum _{k=0}^{k_s-1} \dfrac{\eta _M^k t^k}{k!} \right) \Vert v\Vert \\\le & {} c_\delta \left( \sum _{k=0}^{k_s-1} \dfrac{\delta ^k t^k}{k!} \right) \Vert v\Vert \le c_\delta e^{\delta t} \Vert v\Vert \qquad \forall t \ge 0 \end{aligned}$$

with \( c_\delta := \max \{(\eta _M/\delta )^{k_s - 1},\, 1\} > 0 \).

Second, we establish the lower bound in (36). As \( \Vert {-N(t)}\Vert = \Vert N(t)\Vert \le \eta _M \) for all \( t \ge 0 \), the proof above also implies that for all \( v \in \mathbb C^n \), we have

$$\begin{aligned} \Vert v\Vert = \big \Vert e^{-N(t)\, t} e^{N(t)\, t} v \big \Vert \le c_\delta e^{\delta t} \big \Vert e^{N(t)\, t} v \big \Vert , \end{aligned}$$

that is, \( \Vert e^{N(t)\, t} v\Vert \ge c_\delta ^{-1} e^{-\delta t} \Vert v\Vert \) for all \( t \ge 0 \).

E Computation of \( h(D_{\sigma _2}) \) using (34) in Example 3

Recall from footnote 2 that \( \sigma = 1 \) on \( [t_{2k}, t_{2k+1}) \) and \( \sigma = 2 \) on \( [t_{2k+1}, t_{2k+2}) \), where \( t_0 = 0 \), \( t_1 = 1 \), and \( t_k = 9^{k-1} + 9^{k-2} \) for all \( k \ge 2 \). Hence

(68)

and thus

$$\begin{aligned} \begin{aligned} a_1^1 \tau _1(t) + a_2^1 \tau _2(t) = 3 \tau _2(t) - \tau _1(t)&= {\left\{ \begin{array}{ll} 3.6 t_{2k} - t, &{}t \in [t_{2k}, t_{2k+1}), \\ 3 t - 3.6 t_{2k+1}, &{}t \in [t_{2k+1}, t_{2k+2}) \end{array}\right. } \\ a_1^2 \tau _1(t) + a_2^2 \tau _2(t) = 2 \tau _1(t) - \tau _2(t)&= {\left\{ \begin{array}{ll} 2 t - 2.7 t_{2k}, &{} t \in [t_{2k}, t_{2k+1}), \\ 2.7 t_{2k+1} - t, &{}t \in [t_{2k+1}, t_{2k+2}). \end{array}\right. } \end{aligned} \end{aligned}$$

Then, \( \bar{a}_1 \) and \( \bar{a}_2 \) in (34) satisfy

$$\begin{aligned} \begin{aligned} \bar{a}_1(T)&= \dfrac{1}{T} \max _{t \in [0, T]} a_1^1 \tau _1(t) + a_2^1 \tau _2(t) \\&= {\left\{ \begin{array}{ll} 2.6 t_{2k}/T, &{}T \in [t_{2k}, t_{2k+1} + 8 t_{2k}/3), \\ 3 - 3.6 t_{2k+1}/T, &{}T \in [t_{2k+1} + 8 t_{2k}/3, t_{2k+2}) \end{array}\right. } \\ \bar{a}_2(T)&= \dfrac{1}{T} \max _{t \in [0, T]} a_1^2 \tau _1(t) + a_2^2 \tau _2(t)\\&= {\left\{ \begin{array}{ll} 1.7 t_{2k+1}/T, &{}T \in [t_{2k+1}, t_{2k+2} + 4 t_{2k+1}), \\ 2 - 2.7 t_{2k+2}/T, &{}T \in [t_{2k+2} + 4 t_{2k+1}, t_{2k+3}). \end{array}\right. } \end{aligned} \end{aligned}$$

Hence

$$\begin{aligned} \bar{a}_1(T) + \bar{a}_2(T) = {\left\{ \begin{array}{ll} 17.9 t_{2k}/T, &{}T \in [t_{2k+1}, t_{2k+1} + 8 t_{2k}/3), \\ 3 - 1.9 t_{2k+1}/T, &{}T \in [t_{2k+1} + 8 t_{2k}/3, t_{2k+2}), \\ 25.1 t_{2k+1}/T, &{}T \in [t_{2k+2}, t_{2k+2} + 4 t_{2k+1}), \\ 2 - 0.1 t_{2k+2}/T, &{}T \in [t_{2k+2} + 4 t_{2k+1}, t_{2k+3}). \end{array}\right. } \end{aligned}$$

Therefore,

$$\begin{aligned} h(D_{\sigma _2}) = \limsup _{T \rightarrow \infty }\, \bar{a}_1(T) + \bar{a}_2(T) = \max \{1.99,\, 2.79\} = 2.79. \end{aligned}$$

F Proof of Lemma 7

We regard (43) as a family of scalar differential equations (recall that here \( \xi _\sigma ^k \) denotes the k-th scalar component of \( \xi _\sigma \)):

$$\begin{aligned} \begin{aligned} \dot{\xi }_\sigma ^1&= a_\sigma ^1 \xi _\sigma ^1 + b_\sigma ^{1, 2} \xi _\sigma ^2 + \cdots + b_\sigma ^{1, n} \xi _\sigma ^n, \\ \dot{\xi }_\sigma ^2&= a_\sigma ^2 \xi _\sigma ^2 + b_\sigma ^{2, 3} \xi _\sigma ^3 + \cdots + b_\sigma ^{2, n} \xi _\sigma ^n, \\&\,\,\, \vdots \\ \dot{\xi }_\sigma ^{n-1}&= a_\sigma ^{n-1} \xi _\sigma ^{n-1} + b_\sigma ^{n-1, n} \xi _\sigma ^n, \\ \dot{\xi }_\sigma ^n&= a_\sigma ^n \xi _\sigma ^n, \end{aligned} \end{aligned}$$

and prove Lemma 7 by mathematical induction. For brevity, let

$$\begin{aligned} \psi _{i,j}(t) := b^{i, j}_{\sigma (t)} e^{\eta _j(t) - \eta _i(t)}, \qquad i, j \in \left\{ 1, \ldots , n\right\} . \end{aligned}$$

Then, \( \varPsi \) in (50) can be written as

$$\begin{aligned} \varPsi \left( t, \mathcal C_{k,l,i}\right) = \sum _{(c_0, \ldots , c_i) \in \mathcal C_{k,l,i}} \int _0^t \int _0^{s_1} \cdots \int _0^{s_{i-1}} \prod _{j=1}^{i} \left( \psi _{c_{j-1}, c_j}(s_j) \hbox {d}s_j \right) . \end{aligned}$$
(69)

1.1 F.1 The basis of induction

For the n-th scalar differential equation \( \dot{\xi }_\sigma ^n = a_\sigma ^n \xi _\sigma ^n \), the state-transition function is defined by

$$\begin{aligned} \phi _n(t, s) := e^{\eta _n(t) - \eta _n(s)}, \qquad t \ge s \ge 0. \end{aligned}$$

Hence the n-th scalar component of \( \xi _\sigma (x, t) \) satisfies \( \xi _\sigma ^n(x, t) = e^{\eta _n(t)} x_n \), that is, (50) holds for \( k = n \).

1.2 F.2 The inductive step

For an arbitrary \( m \in \{1, \ldots , n-1\} \), suppose that \( \xi _\sigma ^k(x, t) \) satisfy (50) for all \( k \in \{m+1, \ldots , n\} \). For the m-th differential equation

$$\begin{aligned} \dot{\xi }_\sigma ^m = a_\sigma ^m \xi _\sigma ^m + \sum _{k=m+1}^{n} b_\sigma ^{m, k} \xi _\sigma ^k, \end{aligned}$$

the state-transition function is defined by

$$\begin{aligned} \phi _m(t, s) := e^{\eta _m(t) - \eta _m(s)}, \qquad t \ge s \ge 0. \end{aligned}$$

By variation of constants, the m-th scalar component of \( \xi _\sigma (x, t) \) satisfies

$$\begin{aligned} \begin{aligned} \xi _\sigma ^m(x, t)&= e^{\eta _m(t)} \left( x_m + \sum _{k=m+1}^{n} \int _0^t e^{-\eta _m(s_1)} b_{\sigma (s_1)}^{m, k} \xi _\sigma ^k\left( x, s_1\right) \hbox {d}s_1 \right) \\&= e^{\eta _m(t)} \left( x_m + \sum _{k=m+1}^{n} \int _0^t \psi _{m,k}(s_1) \left( x_k + \sum _{l=k+1}^{n} \sum _{i=1}^{l-k} x_l \varPsi \left( s_1, \mathcal C_{k,l,i}\right) \right) \hbox {d}s_1 \right) \\&= e^{\eta _m(t)} \left( x_m + \sum _{k=m+1}^{n} x_k \int _0^t \psi _{m,k}(s_1) \hbox {d}s_1\right. \\&\left. \quad + \sum _{k=m+1}^{n} \sum _{l=k+1}^{n} \sum _{i=1}^{l-k} x_l \int _0^t \psi _{m,k}(s_1) \varPsi (s_1, \mathcal C_{k,l,i}) \hbox {d}s_1 \right) . \end{aligned} \end{aligned}$$

Based on the definition (52) of \( \mathcal C_{k,l,i} \) and the formula (69) of \( \varPsi \), we have

$$\begin{aligned} \int _0^t \psi _{m,k}(s_1) \hbox {d}s_1 = \varPsi (t, \mathcal C_{m,k,1}) \end{aligned}$$

and

$$\begin{aligned} \begin{aligned}&\int _0^t \psi _{m,k}(s_1) \varPsi \left( s_1, \mathcal C_{k,l,i}\right) \hbox {d}s_1\\&= \sum _{\left( c_1, \ldots , c_{i+1}\right) \in \mathcal C_{k,l,i}} \int _0^t \int _0^{s_1} \cdots \int _0^{s_i} \psi _{m,k}(s_1) \prod _{j=2}^{i+1} \left( \psi _{c_{j-1},c_j}(s_j) \hbox {d}s_j \right) \hbox {d}s_1 \\&= \sum _{\left( c_0, \ldots , c_{i+1}\right) \in \{m\} \times \mathcal C_{k,l,i}} \int _0^t \int _0^{s_1} \cdots \int _0^{s_i} \prod _{j=1}^{i+1} \left( \psi _{c_{j-1},c_{j}}(s_j) \hbox {d}s_j \right) \\&= \varPsi (t, \{m\} \times \mathcal C_{k,l,i}). \end{aligned} \end{aligned}$$

Changing the order of summation, we obtain

$$\begin{aligned} \sum _{k=m+1}^{n} \sum _{l=k+1}^{n} \sum _{i=1}^{l-k} x_l \varPsi (t, \{m\} \times \mathcal C_{k,l,i})&= \sum _{l=m+2}^{n} \sum _{k=m+1}^{l-1} \sum _{i=1}^{l-k} x_l \varPsi \left( t, \{m\} \times \mathcal C_{k,l,i}\right) \\&= \sum _{l=m+2}^{n} \sum _{i'=2}^{l-m} \sum _{k=m+1}^{l-i'+1} x_l \varPsi \left( t, \{m\} \times \mathcal C_{k,l,i'-1}\right) , \end{aligned}$$

where in the last step we also let \( i' = i + 1 \). Next, we prove that the family of sets \( \{\{m\} \times \mathcal C_{k,l,i'-1}: k = m+1, \ldots , l-i'+1\} \) forms a partition of \( \mathcal C_{m,l,i'} \).

  • For all \( (c_1, \ldots , c_{i'}) \in \mathcal C_{k_1,l,i'-1} \) and \( (c'_1, \ldots , c'_{i'}) \in \mathcal C_{k_2,l,i-1} \) with \( k_1 \ne k_2 \), as \( c_1 = k_1 \ne k_2 = c'_1 \), we have \( (c_1, \ldots , c_{i'}) \ne (c'_1, \ldots , c'_{i'}) \). Hence the sets in \( \{\{m\} \times \mathcal C_{k,l,i'-1}: k = m+1, \ldots , l-i'+1\} \) are pairwise disjoint.

  • For all \( (c_1, \ldots , c_{i'}) \in \mathcal C_{k,l,i'-1} \), as \( c_1 = k \ge m+1 \) and \( c_{i'} = l \), we have \( (m, c_1, \ldots , c_{i'}) \in \mathcal C_{m,l,i'} \). Hence \( \bigcup _{k=m+1}^{l-i'-1} \{m\} \times \mathcal C_{k,l,i'-1} \subset \mathcal C_{m,l,i'} \).

  • For all \( (c_0, \ldots , c_i') \in \mathcal C_{m,l,i'} \), as \( c_1 \ge c_0 + 1 = m + 1 \) and \( c_1 \le c_{i'} - (i' - 1) = l - i' + 1 \), we have \( k := c_1 \) satisfies \( m + 1 \le k \le l - i' + 1 \) and \( (c_0, \ldots , c_{i'}) \in \{m\} \times \mathcal C_{k,l,i'-1} \). Hence \( \mathcal C_{m,l,i'} \subset \bigcup _{k=m+1}^{l-i'-1} \{m\} \times \mathcal C_{k,l,i'-1} \).

Therefore,

$$\begin{aligned} \sum _{l=m+2}^{n} \sum _{i'=2}^{l-m} \sum _{k=m+1}^{l-i'+1} x_l \varPsi \left( t, \{m\} \times \mathcal C_{k,l,i'-1}\right) = \sum _{l=m+2}^{n} \sum _{i'=2}^{l-m} x_l \varPsi \left( t, \mathcal C_{m,l,i'}\right) . \end{aligned}$$

Combining the results above, we obtain

$$\begin{aligned} \xi _\sigma ^m(x, t) = e^{\eta _m(t)} \left( x_m + \sum _{l=m+1}^{n} \sum _{i=1}^{l-m} x_l \varPsi \left( t, \mathcal C_{m,l,i}\right) \right) , \end{aligned}$$

that is, (50) holds for \( k = m \). Therefore, mathematical induction implies that (50) holds for all \( k \in \{1, \ldots , n\} \).

G Proof of Lemma 8

For every \( k \in \{1, \ldots , n\} \), following the formula (50) and the triangle inequality, the k-th scalar component of \( \xi _\sigma (x, t) \) satisfies

$$\begin{aligned} |\xi _\sigma ^k(x, t)| \le e^{Re(\eta _k(t))} |x_k| + \sum _{l=k+1}^{n} \left( \sum _{i=1}^{l-k} e^{Re(\eta _k(t))} |\varPsi (t, \mathcal C_{k,l,i})| \right) |x_l|. \end{aligned}$$

First, following the definitions (46) of \( \bar{d}_i \) and (51) of \( \eta _i \), we have

$$\begin{aligned} Re\left( \eta _i(t)\right) \le Re\left( \eta _{i-1}(t)\right) + \bar{d}_i(t)\, t \qquad \forall t \ge 0, \forall i \in \left\{ 2, \ldots , n\right\} . \end{aligned}$$

Hence

$$\begin{aligned} \begin{aligned} |\varPsi (t, \mathcal C_{k,l,i})|&\le \sum _{(c_0, \ldots , c_i) \in \mathcal C_{k,l,i}} b_M^i \int _0^{t} \int _0^{s_1} \cdots \int _0^{s_{i-1}} \prod _{j=1}^{i} \left( e^{Re(\eta _{c_j}(s_j) - \eta _{c_{j-1}}(s_j))} \hbox {d}s_j \right) \\&\le \sum _{\left( c_0, \ldots , c_i\right) \in \mathcal C_{k,l,i}} b_M^i t^i \prod _{j=1}^{i} \left( \max _{s_j \in [0, T]} e^{Re\left( \eta _{c_j}(s_j) - \eta _{c_{j-1}}(s_j)\right) } \right) \\&= \sum _{\left( c_0, \ldots , c_i\right) \in \mathcal C_{k,l,i}} b_M^i t^i e^{\sum _{j=1}^{i} \bar{d}_{c_j}(t)\, t} \\&\le \sum _{\left( c_0, \ldots , c_i\right) \in \mathcal C_{k,l,i}} b_M^i t^i e^{\sum _{j=k+1}^{l} \bar{d}_j(t)\, t}, \end{aligned} \end{aligned}$$

where the last inequality follows partially from the definition (52) of the sets \( \mathcal C_{k,l,i} \). As \( b_M^i t^i \) and \( \sum _{j=k+1}^{l} \bar{d}_j(t)\, t \) are independent of the choice of \( (c_0, \ldots , c_i) \in \mathcal C_{k,l,i} \) and the latter is also independent of the choice of \( i \in \{1, \ldots , l-k\} \), and the set \( \mathcal C_{k,l,i} \) can be characterized by the combinations of \( i-1 \) increasing integers from \( k+1 \) to \( l-1 \), we have

$$\begin{aligned} \sum _{i=1}^{l-k} e^{Re(\eta _k(t))} |\varPsi (t, \mathcal C_{k,l,i})|\le & {} \left( \sum _{i=1}^{l-k} |\mathcal C_{k,l,i}| b_M^i t^i \right) e^{Re(\eta _k(t)) + \sum _{j=k+1}^{l} \bar{d}_j(t)\, t} \\= & {} \left( \sum _{i=1}^{l-k} \left( {\begin{array}{c}l-k-1\\ i-1\end{array}}\right) b_M^i t^i \right) e^{Re(\eta _k(t)) + \sum _{j=k+1}^{l} \bar{d}_j(t)\, t}\\\le & {} (b_M t + 1)^{l - k} e^{Re\left( \eta _k(t)\right) + \sum _{j=k+1}^{l} \bar{d}_j(t)\, t}, \end{aligned}$$

where the last inequality follows partially from the binomial formula. Hence

$$\begin{aligned} \begin{aligned} |\xi _\sigma ^k(x, t)|&\le e^{Re(\eta _k(t))} |x_k| + \sum _{l=k+1}^{n} \Big ( (b_M t + 1)^{l - k} e^{Re(\eta _k(t)) + \sum _{j=k+1}^{l} \bar{d}_j(t)\, t} |x_l| \Big ) \\&\le e^{Re(\eta _k(t))} \sum _{l=k}^{n} \Big ( (b_M t + 1)^{l - k} e^{\sum _{j=k+1}^{l} \bar{d}_j(t)\, t} |x_l| \Big ). \end{aligned} \end{aligned}$$

Note that the upper bound for \( |\xi _\sigma ^k(x, t)| \) above is decreasing in k. Indeed, the upper bound for \( |\xi _\sigma ^{k-1}(x, t)| \) satisfies

$$\begin{aligned} \begin{aligned}&e^{Re(\eta _{k-1}(t))} \sum _{l=k-1}^{n} \Big ( (b_M t + 1)^{l-k+1} e^{\sum _{j=k}^{l} \bar{d}_j(t)\, t} |x_l| \Big ) \\&\quad \ge e^{Re(\eta _{k-1}(t)) + \bar{d}_k(t)\, t} \sum _{l=k}^{n} \Big ( (b_M t + 1)^{l-k} e^{\sum _{j=k+1}^{l} \bar{d}_j(t)\, t} |x_l| \Big ) \\&\quad \ge e^{Re(\eta _k(t))} \sum _{l=k}^{n} \Big ( (b_M t + 1)^{l-k} e^{\sum _{j=k+1}^{l} \bar{d}_j(t)\, t} |x_l| \Big ). \end{aligned} \end{aligned}$$

Hence we obtain (53) by taking the upper bound for \( |\xi _\sigma ^1(x, t)| \) (recall that we take \( \Vert \cdot \Vert \) to be the \( \infty \)-norm; see Remark 1).

Second, recall \( c_0 = k \) and \( c_i = l \), and let \( s_0 := t \) and

$$\begin{aligned} a_m^i := \max _{p \in \mathcal P} Re(a_p^i), i \in \{1, \ldots , n\}. \end{aligned}$$

Following the definition (51) of \( \eta _i \), we have

$$\begin{aligned} Re(\eta _i(t) - \eta _i(\tau )) = \sum _{p \in \mathcal P} Re(a_p^i) (\tau _p(t) - \tau _p(\tau )) \le a_m^i (t - \tau ) \qquad \forall t \ge \tau \ge 0, \forall i \in \{1, \ldots , n\}. \end{aligned}$$

Hence

$$\begin{aligned} \begin{aligned} e^{Re(\eta _k(t))} |\varPsi (t, \mathcal C_{k,l,i})|&\le \sum _{(c_0, \ldots , c_i) \in \mathcal C_{k,l,i}} b_M^i \int _0^{s_0} \cdots \int _0^{s_{i-1}} e^{Re(\eta _{c_i}(s_i))} \prod _{j=1}^{i}\\&\quad \left( e^{Re(\eta _{c_{j-1}}(s_{j-1}) - \eta _{c_{j-1}}(s_j))} \hbox {d}s_j \right) \\&\le \sum _{(c_0, \ldots , c_i) \in \mathcal C_{k,l,i}} b_M^i \int _0^{s_0} \cdots \int _0^{s_{i-1}} e^{a_m^{c_i} s_i} \prod _{j=1}^{i} \Big ( e^{a_m^{c_{j-1}}(s_{j-1} - s_j)} \hbox {d}s_j \Big ) \\&\le \sum _{(c_0, \ldots , c_i) \in \mathcal C_{k,l,i}} b_M^i e^{\max _{0 \le j \le i} a_m^{c_j} t} \Bigg ( \int _0^{s_0} \cdots \int _0^{s_{i-1}} \prod _{j=1}^{i} \hbox {d}s_j \Bigg ) \\&\le \sum _{(c_0, \ldots , c_i) \in \mathcal C_{k,l,i}} b_M^i t^i e^{\max _{k \le j \le l} a_m^j t}, \end{aligned} \end{aligned}$$

where the last inequality follows partially from the definition (52) of the sets \( \mathcal C_{k,l,i} \). As \( b_M^i t^i \) and \( \max _{k \le j \le l} a_m^j t \) are independent of the choice of \( (c_0, \ldots , c_i) \in \mathcal C_{k,l,i} \) and the latter is also independent of the choice of \( i \in \{1, \ldots , l-k\} \), and the set \( \mathcal C_{k,l,i} \) can be characterized by the combinations of \( i-1 \) increasing integers from \( k+1 \) to \( l-1 \), we have

$$\begin{aligned} \begin{aligned} \sum _{i=1}^{l-k} e^{Re(\eta _k(t))} |\varPsi (t, \mathcal C_{k,l,i})|&\le \left( \sum _{i=1}^{l-k} |\mathcal C_{k,l,i}| b_M^i t^i \right) e^{\max _{k \le j \le l} a_m^j t} \\&= \left( \sum _{i=1}^{l-k} \left( {\begin{array}{c}l-k-1\\ i-1\end{array}}\right) b_M^i t^i \right) e^{\max _{k \le j \le l} a_m^j t}\\&\le \left( b_M t + 1\right) ^{l - k} e^{\max _{k \le j \le l} a_m^j t}, \end{aligned} \end{aligned}$$

where the last inequality follows partially from the binomial formula. Hence

$$\begin{aligned} |\xi _\sigma ^k(x, t)| \le e^{Re(\eta _k(t))} |x_k| + \sum _{l=k+1}^{n} \Big ( (b_M t + 1)^{l - k} e^{\max _{k \le j \le l} a_m^j t} |x_l| \Big ). \end{aligned}$$

Note that the upper bound for \( |\xi _\sigma ^k(x, t)| \) above is decreasing in k. Indeed, the upper bound for \( |\xi _\sigma ^{k-1}(x, t)| \) satisfies

$$\begin{aligned} \begin{aligned}&e^{Re(\eta _{k-1}(t))} |x_{k-1}| + \sum _{l=k}^{n} \Big ( (b_M t + 1)^{l-k+1} e^{\max _{k-1 \le j \le l} a_m^j t} |x_l| \Big ) \\&\quad \ge e^{a_m^k t} |x_k| + \sum _{l=k+1}^{n} \Big ( (b_M t + 1)^{l-k} e^{\max _{k \le j \le l} a_m^j t} |x_l| \Big ) \\&\quad \ge e^{Re(\eta _k(t))} |x_k| + \sum _{l=k+1}^{n} \Big ( (b_M t + 1)^{l-k} e^{\max _{k \le j \le l} a_m^j t} |x_l| \Big ). \end{aligned} \end{aligned}$$

Hence we obtain (54) by taking the upper bound for \( |\xi _\sigma ^1(x, t)| \) (recall that we take \( \Vert \cdot \Vert \) to be the \( \infty \)-norm; see Remark 1).

H Computation of an upper bound for \( h(U_{\sigma _2}) \) using (44) in Example 4

Following (68) in “Appendix E”, we have

$$\begin{aligned} \begin{aligned} a_1^1 \tau _1(t) + a_2^1 \tau _2(t)&= 3 \tau _2(t) - \tau _1(t) \\&= {\left\{ \begin{array}{ll} 3.6 t_{2k} - t, &{}t \in [t_{2k}, t_{2k+1}), \\ 3 t - 3.6 t_{2k+1}, &{}t \in [t_{2k+1}, t_{2k+2}) \end{array}\right. } \\ \left( a_1^2 - a_1^1\right) \, \tau _1(t) + \left( a_2^2 - a_2^1\right) \, \tau _2(t)&= 3 \tau _1(t) - 4 \tau _2(t) \\&= {\left\{ \begin{array}{ll} 3 t - 6.3 t_{2k}, &{}t \in [t_{2k}, t_{2k+1}), \\ 6.3 t_{2k+1} - 4 t, &{}t \in [t_{2k+1}, t_{2k+2}). \end{array}\right. } \end{aligned} \end{aligned}$$

Then, \( \bar{a}_1 \) and \( \bar{d}_2 \) in (44) satisfy

$$\begin{aligned} \begin{aligned} \bar{a}_1(T)&= \dfrac{1}{T} \max _{t \in [0, T]} a_1^1 \tau _1(t) + a_2^1 \tau _2(t) \\&= {\left\{ \begin{array}{ll} 2.6 t_{2k}/T &{}T \in [t_{2k}, t_{2k+1} + 8 t_{2k}/3); \\ 3 - 3.6 t_{2k+1}/T, &{}T \in [t_{2k+1} + 8 t_{2k}/3, t_{2k+2}) \end{array}\right. } \\ \bar{d}_2(T)&= \dfrac{1}{T} \max _{t \in [0, T]} (a_1^2 - a_1^1)\, \tau _1(t) + \left( a_2^2 - a_2^1\right) \, \tau _2(t)\\&= {\left\{ \begin{array}{ll} 2.3 t_{2k+1}/T, &{}T \in [t_{2k+1}, 2 t_{2k+2} + 5 t_{2k+1}/3); \\ 3 - 6.3 t_{2k+2}/T, &{}T \in [2 t_{2k+2} + 5 t_{2k+1}/3, t_{2k+3}). \end{array}\right. } \end{aligned} \end{aligned}$$

Hence

$$\begin{aligned} 2 \bar{a}_1(T) + \bar{d}_2(T) = {\left\{ \begin{array}{ll} 25.9 t_{2k}/T, &{}T \in [t_{2k+1}, t_{2k+1} + 8 t_{2k}/3); \\ 6 - 4.9 t_{2k+1}/T, &{}T \in [t_{2k+1} + 8 t_{2k}/3, t_{2k+2}); \\ 49.1 t_{2k+1}/T, &{}T \in [t_{2k+2}, 2 t_{2k+2} + 5 t_{2k+1}/3); \\ 3 - 1.1 t_{2k+2}/T, &{}T \in [t_{2k+2}+ 5 t_{2k+1}/3, t_{2k+3}). \end{array}\right. } \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{aligned} h(U_{\sigma _2})&= \limsup _{T \rightarrow \infty }\, 2 \bar{a}_1(T) + \bar{d}_2(T) = \max \{2.88,\, 5.46\} = 5.46. \end{aligned} \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, G., Schmidt, A.J., Liberzon, D. et al. Topological entropy of switched linear systems: general matrices and matrices with commutation relations. Math. Control Signals Syst. 32, 411–453 (2020). https://doi.org/10.1007/s00498-020-00265-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00498-020-00265-9

Keywords

Navigation