Skip to main content
Log in

Resonances in a Chaotic Attractor Crisis of the Lorenz Flow

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

Local bifurcations of stationary points and limit cycles have successfully been characterized in terms of the critical exponents of these solutions. Lyapunov exponents and their associated covariant Lyapunov vectors have been proposed as tools for supporting the understanding of critical transitions in chaotic dynamical systems. However, it is in general not clear how the statistical properties of dynamical systems change across a boundary crisis during which a chaotic attractor collides with a saddle. This behavior is investigated here for a boundary crisis in the Lorenz flow, for which neither the Lyapunov exponents nor the covariant Lyapunov vectors provide a criterion for the crisis. Instead, the convergence of the time evolution of probability densities to the invariant measure, governed by the semigroup of transfer operators, is expected to slow down at the approach of the crisis. Such convergence is described by the eigenvalues of the generator of this semigroup, which can be divided into two families, referred to as the stable and unstable Ruelle–Pollicott resonances, respectively. The former describes the convergence of densities to the attractor (or escape from a repeller) and is estimated from many short time series sampling the state space. The latter is responsible for the decay of correlations, or mixing, and can be estimated from a long times series, invoking ergodicity. It is found numerically for the Lorenz flow that the stable resonances do approach the imaginary axis during the crisis, as is indicative of the loss of global stability of the attractor. On the other hand, the unstable resonances, and a fortiori the decay of correlations, do not flag the proximity of the crisis, thus questioning the usual design of early warning indicators of boundary crises of chaotic attractors and the applicability of response theory close to such crises.

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
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. A saddle is defined here as a possibly chaotic invariant set, which is attracting neither in forward nor in backward time (see [14] for a more rigorous definition).

  2. This attractor must be only partially hyperbolic, since it contains the singularity at the origin, which prevents the stable and unstable manifolds to be continuous.

  3. Following the algorithm described in [67], also yielding the CLVs.

  4. The flow will be assumed to be nonsingular for the Lebesgue measure, so that it maps sets of null measure into sets of null measure.

  5. For a probability density f, i.e. \(f \ge 0\) and \(\int _X f(x) \eta (dx)= 1\), (3.3) expresses the fact that the probability to find a member x sampled from an initial ensemble f in a set A after some time t is none-other than the probability of this member to be initially in the preimage of this set by the flow.

  6. As explained in “Spectral Theory of Ergodicity and Mixing” section of Appendix A, resonances to the left of the imaginary axis are associated to eigenfunctions on spaces of distributions rather than on spaces of integrable functions.

  7. In the uniform discretization of the ball \(B_o\) in spherical coordinates, in Sect. 3.4, different grid boxes will have different Lebesgue volumes. The component \(f_i\) of an observable f being the integral of f over the box \(B_i\), this component will tend to be larger for boxes of larger volume. This does not affect the eigenvalues, but clearly affects the eigenvectors, where large values of the components might be associated to large boxes, ceteris paribus.

  8. The dependence of the \(\lambda _k\)’s on \(\tau \) thus constitute an important test for the quality of the approximations. This is discussed in Appendix B.

  9. Note that the secondary eigenvectors being orthogonal to the leading adjoint eigenvectors, which is constant (see “Multiple Attractors” section of Appendix A), they integrate to zero with respect to the Lebesgue measure.

  10. The sample correlation function is estimated from a long time series initialized in the basin of attraction of \(\varLambda \) as a discrete approximation of the time average (A.6) (see e.g. [94]).

  11. In particular, the ergodicity of the invariant measure implies that each set of positive measure is visited infinitely often by orbits starting from \(\mu \)-almost every point [100, Lemma 6.15].

  12. Contraction by semigroups can be interpreted in physical systems as determining entropy production. Recognizing this has been essential to understand how reversible microscopic evolutions can lead to irreversible macroscopic properties of systems out of thermal equilibrium (see the pioneering work [102,103,104] for reviews).

  13. There is no general result regarding the convergence with the grid resolution of eigenvalues of transition matrices to the Ruelle–Pollicott resonances. This is, however, the case for uniformly hyperbolic systems [33, 34] for which Ulam’s method may converge (see [107], for the case of hyperbolic maps).

References

  1. Guckenheimer, J.M., Holmes, P.: Nonlinear Oscillations, Dynamical Systems, and Bifurcation of Vector Fields. Springer, New York (1983)

    Book  MATH  Google Scholar 

  2. Ruelle, D.: Elements of Differentiable Dynamics and Bifurcation Theory. Academic Press, San Diego (1989). https://doi.org/10.1016/B978-0-12-601710-6.50001-8

    MATH  Google Scholar 

  3. Kuznetsov, Y.A.: Elements of Applied Bifurcation Theory, 2nd edn. Springer, New York (1998)

    MATH  Google Scholar 

  4. Kloeden, P.E., Rasmussen, M.: Nonautonomous Dynamical Systems. American Mathematical Society, Providence (2011)

    Book  MATH  Google Scholar 

  5. Arnold, L.: Random Dynamical Systems. Springer, Berlin (2003)

    Google Scholar 

  6. Held, H., Kleinen, T.: Detection of climate system bifurcations by degenerate fingerprinting. Geophys. Res. Lett. 31(December), 1 (2004). https://doi.org/10.1029/2004GL020972

    Google Scholar 

  7. Kleinen, T., Held, H., Petschel-Held, G.: The potential role of spectral properties in detecting thresholds in the Earth system: application to the thermohaline circulation. Ocean Dyn. 53, 53 (2003). https://doi.org/10.1007/s10236-002-0023-6

    Article  ADS  Google Scholar 

  8. van Nes, E.H., Scheffer, M.: Slow recovery from perturbations as a generic indicator of a nearby catastrophic shift. Am. Nat. 169(6), 738 (2007)

    Article  Google Scholar 

  9. Scheffer, M., Bascompte, J., Brock, W.A., Brovkin, V., Carpenter, S.R., Dakos, V., Held, H., van Nes, E.H., Rietkerk, M., Sugihara, G.: Early-warning signals for critical transitions. Nature 461(7260), 53 (2009). https://doi.org/10.1038/nature08227

    Article  ADS  Google Scholar 

  10. Lenton, T.M.: Early warning of climate tipping points. Nat. Clim. Change 1(4), 201 (2011). https://doi.org/10.1038/nclimate1143

    Article  ADS  Google Scholar 

  11. Eckmann, J.P., Ruelle, D.: Ergodic theory of chaos and strange attractors. Rev. Mod. Phys. 57(July), 617 (1985). https://doi.org/10.1103/RevModPhys.57.617

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Young, L.S.: What are SRB measures, and which dynamical systems have them? J. Stat. Phys. 108(5), 733 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  13. Grebogi, C., Ott, E., Yorke, J.A.: Crises, sudden changes in chaotic attractors, and transient chaos. Physica D Nonlinear Phenom. 7(1–3), 181 (1983). https://doi.org/10.1016/0167-2789(83)90126-4

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Ashwin, P., Buescu, J., Stewart, I.: From attractor to chaotic saddle: a tale of transverse instability. Nonlinearity 9, 703 (1996). https://doi.org/10.1088/0951-7715/9/3/006

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Skufca, J.D., Yorke, J.A., Eckhardt, B.: Edge of chaos in a parallel shear flow. Phys. Rev. Lett. 96(17), 5 (2006). https://doi.org/10.1103/PhysRevLett.96.174101

    Article  Google Scholar 

  16. Schneider, T.M., Eckhardt, B., Yorke, J.A.: Turbulence transition and the edge of chaos in pipe flow. Phys. Rev. Lett. 99(3), 1 (2007). https://doi.org/10.1103/PhysRevLett.99.034502

    Article  Google Scholar 

  17. Eckhardt, B.: Turbulence transition in pipe flow: some open questions. Nonlinearity 21, T1 (2008). https://doi.org/10.1088/0951-7715/21/1/T01

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Bódai, T., Lucarini, V., Lunkeit, F., Boschi, R.: Global instability in the Ghil–Sellers model. Clim. Dyn. 44, 3361 (2015)

    Article  Google Scholar 

  19. Lucarini, V., Bodai, T.: Edge states in the climate system: exploring global instabilities and critical transitions. Nonlinearity 30, R32 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Oseledets, V.I.: A multiplicative ergodic theorem. Characteisitic Lyapunov, exponents of dynamical systems. Tr. Mosk. Mat. Obs. 19, 179 (1968)

    Google Scholar 

  21. La Salle, J.P.: The Stability of Dynamical Systems. Society for Industrial and Applied Mathematics, Philadelphia (1976)

    Book  Google Scholar 

  22. Faranda, D., Lucarini, V., Manneville, P., Wouters, J.: On using extreme values to detect global stability thresholds in multi-stable systems: the case of transitional plane Couette flow. Chaos Solitons Fractals 64(1), 26 (2014). https://doi.org/10.1016/j.chaos.2014.01.008

    Article  ADS  MATH  Google Scholar 

  23. Faranda, D., Dubrulle, B., Pons, F.M.E.: Statistical early-warning indicators based on autoregressive moving-average models. J. Phys. A Math. Theor. 47(25), 252001 (2014). https://doi.org/10.1088/1751-8113/47/25/252001

    Article  ADS  MATH  Google Scholar 

  24. Lasota, A., Mackey, M.C.: Chaos, Fractals and Noise. Springer, Berlin (1994)

    Book  MATH  Google Scholar 

  25. Halmos, P.R.: Lectures on Ergodic Theory. Chelsea Publishing Company, New York (1956)

    MATH  Google Scholar 

  26. Arnold, V.I., Avez, A.: Ergodic Problems of Classical Mechanics (Advanced Book Classics). W. A. Benjamin, New York (1968). https://doi.org/10.1002/zamm.19700500721

    Google Scholar 

  27. Baladi, V.: Spectrum and statistical properties. In: Casacuberta, C., Miró-Roig, R.M., Verdera, J., Xambó-Descamps, S. (eds.) European Congress of Mathematics, pp. 203–223. Basel, Birkhäuser (2001). https://doi.org/10.1007/978-3-0348-8268-2_11

    Chapter  Google Scholar 

  28. Young, L.S.: Understanding chaotic dynamical systems. Commun. Pure Appl. Math. 66(9), 1439 (2013). https://doi.org/10.1002/cpa.21468

    Article  MathSciNet  MATH  Google Scholar 

  29. Pollicott, M.: On the rate of mixing of Axiom A flows. Invent. Math. 81(3), 413 (1985). https://doi.org/10.1007/BF01388579

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Ruelle, D.: Resonances of chaotic dynamical systems. Phys. Rev. Lett. 56(5), 405 (1986)

    Article  ADS  MathSciNet  Google Scholar 

  31. Liverani, C.: Decay of correlations. Ann. Math. 142, 239 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  32. Blank, M., Keller, G., Liverani, C.: Ruelle–Perron–Frobenius spectrum for Anosov maps. Nonlinearity 15(6), 58 (2001). https://doi.org/10.1088/0951-7715/15/6/309

    MathSciNet  MATH  Google Scholar 

  33. Gouëzel, S., Liverani, C.: Banach spaces adapted to Anosov systems. Ergod. Theory. Dyn. Syst. 26(01), 26 (2006). https://doi.org/10.1017/S0143385705000374

    MathSciNet  MATH  Google Scholar 

  34. Butterley, O., Liverani, C.: Smooth Anosov flows: correlation spectra and stability. J. Mod. Dyn. 1(2), 301 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  35. Fauré, F., Tsujii, M.: In: Dal’Bo, F., Peigné, M., Sambusetti, A. (eds.) Analytic and Probabilistic Approaches to Dynamics in Negative Curvature, chap. 2. Springer, Cham, pp. 65–138 (2014)

  36. Baladi, V.: The quest for the ultimate anisotropic Banach space. J. Stat. Phys. 166(3), 525 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Collet, P., Eckmann, J.P.: Liapunov multipliers and decay of correlations in dynamical systems. J. Stat. Phys. 115(April), 217 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Alves, F., Luzzatto, S., Pinheiro, V.: Lyapunov exponents and rates of mixing for one-dimensional maps. Ergod. Theory. Dyn. Syst. 24(3), 637 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  39. Pires, C.J.A., Saa, A., Venegeroles, R.: Lyapunov statistics and mixing rates for intermittent systems. Phys. Rev. E 84, 066210 (2011). https://doi.org/10.1103/PhysRevE.84.066210

    Article  ADS  Google Scholar 

  40. Slipantschuk, J., Bandtlow, O.F., Just, W.: On the relation between Lyapunov exponents and exponential decay of correlations. J. Phys. A 46, 1 (2013). https://doi.org/10.1088/1751-8113/46/7/075101

    Article  MathSciNet  MATH  Google Scholar 

  41. Vaidya, U., Mehta, P.G., Trans, I.E.E.E.: Lyapunov measure for almost everywhere stability. Autom. Contrib. 53(1), 307 (2008). https://doi.org/10.1109/TAC.2007.914955

    Article  MathSciNet  MATH  Google Scholar 

  42. Mauroy, A., Mezić, I., Trans, I.E.E.E.: Global stability analysis using the eigenfunctions of the Koopman operator. Autom. Contrib. 61(11), 3356 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  43. Tantet, A., Lucarini, V., Lunkeit, F., Dijkstra, H.A.: Crisis of the chaotic attractor of a climate model: a transfer operator approach (2017). arXiv:1507.02228v2

  44. Ruelle, D.: A review of linear response theory for general differentiable dynamical systems. Nonlinearity 22, 855 (2009). https://doi.org/10.1088/0951-7715/22/4/009

    Article  ADS  MathSciNet  MATH  Google Scholar 

  45. Cessac, B., Sepulchre, J.A.: Linear response, susceptibility and resonances in chaotic toy models. Physica D Nonlinear Phenom. 225, 13 (2007). https://doi.org/10.1016/j.physd.2006.09.034

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. Gritsun, A.S., Lucarini, V.: Fluctuations, response, and resonances in a simple atmospheric model. Physica D Nonlinear Phenom. 349, 62–76 (2017). https://doi.org/10.1016/j.physd.2017.02.015

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. Hairer, M., Majda, A.J.: A simple framework to justify linear response theory. Nonlinearity 23(4), 909 (2010). https://doi.org/10.1088/0951-7715/23/4/008

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. Lucarini, V.: Stochastic perturbations to dynamical systems: a response theory approach. J. Stat. Phys. 146(4), 774 (2012). https://doi.org/10.1007/s10955-012-0422-0

    Article  MathSciNet  MATH  Google Scholar 

  49. Gaspard, P.: Trace formula for noisy flows. J. Stat. Phys. 106(1–2), 57 (2002). https://doi.org/10.1023/A:1013167928166

    Article  MathSciNet  MATH  Google Scholar 

  50. Tantet, A., Chekroun, M.D., Neelin, J.D., Dijkstra, H.A.: Mixing spectrum in reduced space of differential equations. Part II: Stochastic Hopf bifurcation (2017). arXiv:1705.07573

  51. Lorenz, E.N.: Deterministic nonperiodic flow. J. Atmos. Sci. 20, 130 (1963)

    Article  ADS  Google Scholar 

  52. Sparrow, C.: The Lorenz Equations: Bifurcations, Chaos and Strange Attractors. Springer, New York (1982)

    Book  MATH  Google Scholar 

  53. Guckenheimer, J.M., Williams, R.F.: Structural stability of Lorenz attractors. Publ. Math. l’I.H.É.S 50, 59 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  54. Araújo, V., Pacifico, M.J.: Three-Dimensional Flows. Springer, Heidelberg (2010). https://doi.org/10.1007/978-3-642-11414-4

    Book  MATH  Google Scholar 

  55. Tucker, W., Acad, C.R.: The Lorenz attractor exists. Sci. Paris 328, 1197 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  56. Reick, C.H.: Linear response of the Lorenz system. Phys. Rev. E Stat. Nonlinear Soft Matter Phys 66, 1 (2002). https://doi.org/10.1103/PhysRevE.66.036103

    Article  MathSciNet  Google Scholar 

  57. Lucarini, V.: Evidence of dispersion relations for the nonlinear response of the Lorenz 63 system. J. Stat. Phys. 134, 381 (2009). https://doi.org/10.1007/s10955-008-9675-z

    Article  ADS  MathSciNet  MATH  Google Scholar 

  58. Rollins, R.W., Hunt, E.R.: Intermittent transient chaos at interior crises in the diode resonator. Phys. Rev. A 29(6), 3327 (1984)

    Article  ADS  Google Scholar 

  59. Pompe, B., Leven, R.W.: Behavior of Lyapunov exponents near crisis points in the dissipative standard map. Phys. Scr. 38(5), 651 (1988). https://doi.org/10.1088/0031-8949/38/5/003

    Article  ADS  MATH  Google Scholar 

  60. Mehra, V., Ramaswamy, R.: Maximal Lyapunov exponent at crises. Phys. Rev. E 53(4), 3420 (1996)

    Article  ADS  Google Scholar 

  61. Beims, M.W., Gallas, J.A.C.: Alignment of Lyapunov vectors: a quantitative criterion to predict catastrophes? Sci. Rep. 6(November), 37102 (2016). https://doi.org/10.1038/srep37102

    Article  ADS  Google Scholar 

  62. Lorenz, E.N.: On the prevalence of aperiodicity in simple systems. In: Grmela, M., Marsden, J.E. (eds.) Global Analysis, pp. 53–75. Springer, Berlin (1979)

    Chapter  Google Scholar 

  63. Hartman, P.: Ordinary Differential Equations, vol. 53. Wiley, New York (1964)

    MATH  Google Scholar 

  64. Kaplan, J.L., Yorke, Ja: Preturbulence: a regime observed in a fluid flow model of Lorenz. Commun. Math. Phys. 67(2), 93 (1979). https://doi.org/10.1007/BF01221359

    Article  ADS  MathSciNet  MATH  Google Scholar 

  65. Katok, A., Hasselblatt, B.: Introduction to the Modern Theory of Dynamical Systems. Cambridge University Press, Cambridge (1996)

    MATH  Google Scholar 

  66. Barreira, L., Pesin, Y. B., Sarig, O.: In: Hasselblatt, B., Katok, A. (eds.) Handbook of Dynamical Systems, vol. 1B. Elsevier, Amsterdam, pp. 57–263 (2006)

  67. Kuptsov, P.V., Parlitz, U.: Theory and computation of covariant Lyapunov vectors. J. Nonlinear Sci. 22(5), 727 (2012). https://doi.org/10.1007/s00332-012-9126-5

    Article  ADS  MathSciNet  MATH  Google Scholar 

  68. Pesin, Y.B.: Lectures on Partial Hyperbolocity and Stable Ergodicity, zurich lec edn. European Mathematical Society, Zurich (2004). https://doi.org/10.4171/003

    Book  MATH  Google Scholar 

  69. Ginelli, F., Poggi, P., Turchi, A., Chaté, H., Livi, R., Politi, A.: Characterizing dynamics with covariant lyapunov vectors. Phys. Rev. Lett. 99(13), 1 (2007). https://doi.org/10.1103/PhysRevLett.99.130601

    Article  Google Scholar 

  70. Yang, H.L., Takeuchi, K.A., Ginelli, F., Chaté, H., Radons, G.: Hyperbolicity and the effective dimension of spatially extended dissipative systems. Phys. Rev. Lett. 102(7), 1 (2009). https://doi.org/10.1103/PhysRevLett.102.074102

    Article  Google Scholar 

  71. Engel, K.J., Nagel, R.: One-Parameter Semigroups for Linear Evolution Equations. Springer, New York (2001)

    MATH  Google Scholar 

  72. Davies, E.B.: Linear Operators and Their Spectra. Cambridge University Press, Cambridge (2007)

    Book  MATH  Google Scholar 

  73. Yosida, K.: Functional Analysis, vol. 123. Springer, Berlin (1980)

    MATH  Google Scholar 

  74. Koopman, B.O.: Hamiltonian systems and transformations in Hilbert space. Proc. Natl Acad. Sci. U.S.A. 17(5), 315 (1931). https://doi.org/10.1073/pnas.17.5.315

    Article  ADS  MATH  Google Scholar 

  75. von Neumann, J.: Proof of the quasi-ergodic hypothesis. Proc. Natl Acad. Sci. U.S.A. 18(2), 70 (1932). https://doi.org/10.1073/pnas.18.1.70

    Article  ADS  MATH  Google Scholar 

  76. Hasegawa, H., Saphir, W.: Unitary and irreversibility in chaotic systems. Phys. Rev. A 46(12), 7401 (1992). https://doi.org/10.1103/PhysRevA.46.7401

    Article  ADS  MathSciNet  Google Scholar 

  77. Gaspard, P., Ramirez, D.A.: Diffusion, effusion, and chaotic scattering: an exactly solvable Liouvillian dynamics. Phys. Rev. A 45(12), 8383 (1992). https://doi.org/10.1103/PhysRevA.45.8383

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. Gaspard, P., Nicolis, G., Provata, A., Tasaki, S.: Spectral signature of the Pitchfork bifurcation: Liouville equation approach. Phys. Rev. E 51(1), 74 (1995)

    Article  ADS  MathSciNet  Google Scholar 

  79. Gaspard, P., Tasaki, S.: Liouvillian dynamics of the Hopf bifurcation. Phys. Rev. E 64(5), 056232 (2001). https://doi.org/10.1103/PhysRevE.64.056232

    Article  ADS  MathSciNet  Google Scholar 

  80. Ulam, S.M.: Problems in Modern Mathematics, science edn. Wiley, New York (1964)

    MATH  Google Scholar 

  81. Dellnitz, M., Junge, O.: On the approximation of complicated dynamical behavior. SIAM J. Numer. Anal. 36(2), 491 (1999). https://doi.org/10.1137/S0036142996313002

    Article  MathSciNet  MATH  Google Scholar 

  82. Klus, S., Koltai, P., Schütte, C.: On the numerical approximation of the Perron-Frobenius and Koopman operator. J. Comput. Dyn. 3(1), 51–79 (2015)

    MathSciNet  MATH  Google Scholar 

  83. Williams, M.O., Kevrekidis, I.G., Rowley, C.W.: A data-driven approximation of the Koopman operator: extending dynamic mode decomposition. J. Nonlinear Sci. 25(6), 1307 (2015). https://doi.org/10.1007/s00332-015-9258-5

    Article  ADS  MathSciNet  MATH  Google Scholar 

  84. Lucarini, V.: Response operators for Markov processes in a finite state space: radius of convergence and link to the response theory for Axiom A systems. J. Stat. Phys. 162(2), 312 (2016). https://doi.org/10.1007/s10955-015-1409-4

    Article  ADS  MathSciNet  MATH  Google Scholar 

  85. Schütte, C.: Conformational dynamics: modelling, theory. algorithm and application to biomolecules. Tech. Rep. July, Konrad-Zuse-Zentrum für Informationstechnik, Berlin (1999)

  86. Chekroun, M.D., Neelin, J.D., Kondrashov, D., McWilliams, J.C., Ghil, M.: Rough parameter dependence in climate models and the role of Ruelle–Pollicott resonances. Proc. Natl Acad. Sci. U.S.A. 111(5), 1684 (2014). https://doi.org/10.1073/pnas.1321816111

    Article  ADS  Google Scholar 

  87. Chekroun, M.D., Tantet, A., Neelin, J.D., Dijkstra, H.A.: Mixing spectrum in reduced space of stochastic differential equations. Part I: Theory (2017, in preparation)

  88. Tantet, A., van der Burgt, F.R., Dijkstra, H.A.: An early warning indicator for atmospheric blocking events using transfer operators. Chaos 25(3), 036406 (2015). https://doi.org/10.1063/1.4908174

    Article  ADS  Google Scholar 

  89. Fishman, S., Rahav, S.: Dynamics of Dissipation, pp. 165–192. Springer, Berlin (2002)

    Book  MATH  Google Scholar 

  90. Billingsley, P.: Statistical Inference for Markov process. University of Chicago Press, Chicago (1961)

    MATH  Google Scholar 

  91. Baker, C.G., Hetmaniuk, U.L., Lehoucq, R.B., Thornquist, H.K., Trans, A.C.M.: Anasazi software for the numerical solution of large-scale eigenvalue problems. Math. Softw. 36(3), 13 (2009). https://doi.org/10.1145/1527286.1527287

    Article  MathSciNet  MATH  Google Scholar 

  92. Heroux, M., Bartlett, R., Howle, V., Hoekstra, R., Hu, J., Kolda, T., Lehoucq, R.B., Long, K., Pawlowski, R., Phipps, E., Salinger, A., Thornquist, H.K., Tuminaro, R., Willenbring, J., Williams, A.: An overview of Trilinos. Tech. Rep. Sandia National Laboratories, Albuquerque (2003)

  93. Froyland, G., Junge, O., Koltai, P.: Estimating long term behavior of flows without trajectory integration: the infinitesimal generator approach. SIAM J. Numer. Anal. 51(1), 223 (2011). https://doi.org/10.1137/110819986

    Article  MathSciNet  MATH  Google Scholar 

  94. von Storch, H., Zwiers, F.: Stastistical Analysis in Climate Research. Cambridge University Press, Cambridge (1999)

    Book  Google Scholar 

  95. Lan, Y., Mezić, I.: Linearization in the large of nonlinear systems and Koopman operator spectrum. Physica D Nonlinear Phenom. 242(1), 42 (2013). https://doi.org/10.1016/j.physd.2012.08.017

    Article  ADS  MathSciNet  MATH  Google Scholar 

  96. Kubo, R.: Statistical-mechanical theory of irreversible processes. I. General theory and simple applications to magnetic and conduction problems. J. Phys. Soc. Jpn. 12(3), 570–586 (1957). https://doi.org/10.1143/JPSJ.12.570

    Article  ADS  MathSciNet  Google Scholar 

  97. Allawala, A., Marston, J.B.: Statistics of the stochastically forced Lorenz attractor by the Fokker–Planck equation and cumulant expansions. Phys. Rev. E 94(5), 1 (2016). https://doi.org/10.1103/PhysRevE.94.052218

    Article  Google Scholar 

  98. Hörmander, L.R.: Hypoelliptic second order differential equations. Acta Math. 119(1), 147 (1968). https://doi.org/10.1007/BF02392081

    MathSciNet  MATH  Google Scholar 

  99. Hairer, M.: On Malliavin’s proof of Hörmander’s theorem. Bull. Sci. Math. 135(6–7), 650 (2011). https://doi.org/10.1016/j.bulsci.2011.07.007

    Article  MathSciNet  MATH  Google Scholar 

  100. Eisner, T., Farkas, B., Haase, M., Nagel, R.: Operator Theoretic Aspects of Ergodic Theory. Springer International Publishing, Cham (2015). https://doi.org/10.1007/978-3-319-16898-2

    Book  MATH  Google Scholar 

  101. Gallavotti, G.: Nonequilibrium and Irreversibility. Springer, Cham (2014). https://doi.org/10.1007/978-3-319-06758-2

    Book  MATH  Google Scholar 

  102. Misra, B., Prigogine, I., Courbage, M.: From deterministic dynamics to probabilistic descriptions. Physica A 98(1–2), 1 (1979). https://doi.org/10.1016/0378-4371(79)90163-8

    Article  ADS  MathSciNet  MATH  Google Scholar 

  103. Gaspard, P.: Chaos, Scattering and Statistical Mechanics. Cambridge University Press, Cambridge (1998)

    Book  MATH  Google Scholar 

  104. Garbaczewski, P., Olkiewicz, R. (eds.): Dynamics of Dissipation. Springer, Berlin (2002)

    MATH  Google Scholar 

  105. Keller, G., Liverani, C., Roma, T.U.D.: Stability of the spectrum for transfer operators. Tech. Rep. Scuola Norm. Sup. Pisa, Pisa (1998)

  106. Baladi, V., Holschneider, M.: Approximation of nonessential spectrum of transfer operators. Nonlinearity 12(December), 525 (1999). https://doi.org/10.1088/0951-7715/12/3/006

    Article  ADS  MathSciNet  MATH  Google Scholar 

  107. Froyland, G.: On Ulam approximation of the isolated spectrum and eigenfunctions of hyperbolic maps. Discret. Contin. Dyn. Syst. 17(3), 671 (2007). https://doi.org/10.3934/dcds.2007.17.671

    Article  MathSciNet  MATH  Google Scholar 

  108. Lang, S.: Real and Functional Analysis. Springer, New York (1993)

    Book  MATH  Google Scholar 

  109. Crommelin, D., Vanden-Eijnden, E.: Data-based inference of generators for Markov jump processes using convex optimization. Multiscale Model. Simul. 7(4), 1 (2009)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

AT would like to thank Sebastian Schubert for very helpful discussions on Lyapunov exponents and vectors and for providing his numerical code to estimate them. AT would also like to thank Andrey Gritsun for valuable discussions on the role of periodic orbits in chaotic systems as well as their numerical continuation. AT and VL acknowledge support from the DFG Project MERCI. VL would like to thank C. Liverani, T. Kuna, and V. Baladi for various inspiring conversations. VL acknowledges the support of the Horizon2020 Project CRESCENDO and of the DFG SFB/Transregio project TRR181. HD likes to acknowledge the support of the Netherlands Center for Earth System Science (NESSC) funded by the Netherlands Foundation for Scientific Research (NWO). Part of this work was carried out on the Dutch national e-infrastructure (Cartesius) with the support of SURF Cooperative under the project SH284.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexis Tantet.

Appendices

Appendix A: Transfer Operators, Resonances and Decay of Correlations

Let us shortly review the spectral theory of transfer operators relevant for this study. A more detailed exposition can be found in [43] and references therein.

1.1 Ergodicity, Mixing and Correlations

A key concept in ergodic theory is that of an invariant measure for the flow \(\varPhi _t\), that is, a probability measure \(\mu \) such that

$$\begin{aligned} \mu (\varPhi _t^{-1} A) = \mu (A), \quad \mathrm {for~any~} A \in \mathcal {B}. \end{aligned}$$
(A.1)

In other words, a measure is invariant if, according to this measure, the probability of a state to be in some set does not change as this state is propagated by the flow. It follows then that the average with respect to the invariant measure \(\mu \) of any integrable observable g is also invariant with time, i.e.

$$\begin{aligned} \int _X g(\varPhi _t x) \mu (d x) = \int _X g(x) \mu (d x) := \langle g\rangle _\mu , \end{aligned}$$
(A.2)

so that the invariant measure gives a statistical steady-state.

A flow \(\varPhi _t\) with an invariant measure \(\mu \) has interesting statistical properties when \(\mu \) is ergodic, that is, when the sets A which are invariant, i.e. \(\varPhi _t^{-1} A = A\), are either of measure 0 or 1.Footnote 11 Then, by the celebrated individual ergodic theorem of Birkhoff [24, Chap. 7.3], the average of any \(\mu \)-integrable observable g is such that

$$\begin{aligned} \langle g\rangle _\mu = \lim _{T \rightarrow \infty } \frac{1}{T} \int _0^T g(\varPhi _t x) dt := \bar{g}, \quad \mathrm {for ~} \mu \mathrm {-almost~every~} x. \end{aligned}$$
(A.3)

Thus, when \(\mu \) is ergodic, the time mean is independent of the initial state x except for a set of null measure.

Remark 2

There may exist many ergodic measures. This is actually the case for the Lorenz flow for \(\rho _{\mathrm {Homo}}< \rho < \rho _A\), where, as seen in Sect. 2, three attractors coexist. Each attractor then supports at least one invariant measure [65, Chap. 4]. In this case, the equality (A.3) between ensemble averages and time averages may hold only for initial states belonging to the attractor supporting the measure, while time averages for two initial conditions in different basins of attraction will not coincide in general. More useful for experiments is then the eventual physical property of the measure. The latter ensures that the equality (A.3) between ensemble averages and time averages holds not only for initial states in a set of positive measure \(\mu \), but also for states in any set of positive Lebesgue measure m in the basin of attraction of a given attractor [12].

A particularly important quantity in ergodic theory is the correlation function,

$$\begin{aligned} C_{f,g}(t) := \int _X f(x) g(\varPhi _t x) \mu (dx) - \langle f\rangle _\mu \langle g\rangle _\mu , \quad t \ge 0, \end{aligned}$$
(A.4)

between any observables \(f, g \in L^2_\mu (X)\). It gives a measure of the relationship between the two observables as one evolves with time. The decay of correlation functions with time is a macroscopic manifestation of chaos. Indeed, this decay is equivalent to the mixing property (see e.g. [24, Chap. 4] and [65, Chap. 4]),

$$\begin{aligned} \lim _{t\rightarrow \infty } \mu (A \cap \varPhi _t^{-1} B) = \mu (A) \mu (B), \quad \mathrm {for~any~} A, B \in \mathcal {B}, \end{aligned}$$
(A.5)

of the invariant measure \(\mu \), which is stronger than ergodicity. In other words, the probability for some state to be in any set B after some time t is independent of the probability of this state to be in any set A initially. Any information about the initial state of an ensemble is thus gradually forgotten.

Remark 3

For a physical measure \(\mu \) supported by an attractor \(\varLambda \), the correlation function can be estimated by the time mean

$$\begin{aligned} C_{f,g}(t) = \lim _{T \rightarrow \infty } \frac{1}{T} \int _0^T (f(x) - \bar{f})(g(\varPhi _t x) - \bar{g}) ~ dt, \end{aligned}$$
(A.6)

from a single time initialized in the basin of attraction of \(\varLambda \). For this reason, we will see in Sect. 3.3.2 that complete information on \(\mathcal {P}^\mu _t, t \ge 0\) can be obtained from from observations on the corresponding attractor alone. Unfortunately, for the forced and dissipative systems considered here, volumes contract on average [101, Chap. 2.8] so that the invariant measure \(\mu \) is supported by an attractor with zero Lebesgue measure m. It follows that, as opposed to the semigroup \(\mathcal {P}^m_t, t \ge 0\), no information on the dynamics away from the attractor is carried by \(\mathcal {P}^\mu _t\).

1.2 Spectral Theory of Ergodicity and Mixing

One can see from the definition (A.4) that the correlation function is fully determined by the family of transfer (Koopman) operators \(\mathcal {P}_t^\mu \) (\(\mathcal {U}_t^\mu \)) on \(L^2_\mu (X)\), i.e. with respect to \(\mu \). In fact, classical results relate the ergodic properties of measure-preserving dynamical systems to the behaviour of the semigroups [24,25,26]. Note first, that the invariance of the measure \(\mu \) together with the invertibility of the flow ensure that the semigroups are isometries and constitute a unitary group on \(L^2_\mu (X)\). As a consequence, the spectrum of the operators is contained in the unit circle \(|z| = 1, z \in \mathbb {C}\) and \(\mathcal {P}_t^\mu \) and \(\mathcal {U}_t^\mu \) have the same eigenvalues and eigenfunctions, for any real t. Moreover, the following results hold.

Theorem 1

If the measure-preserving dynamical system \((X, \mathcal {B}, (\varPhi _t)_{t\in \mathbb {R}}, \mu )\) is

  1. (i)

    ergodic, then 1 is a simple eigenvalue of \(\mathcal {P}_t^\mu , t\in \mathbb {R}\), and conversely.

  2. (ii)

    mixing, then the only one eigenvalue is one.

The first item of this theorem can be understood from the fact that ergodic systems have only one stationary density with respect to the invariant measure. The second is due to the fact that eigenfunctions associated with eigenvalues on the unit circle do not decay and thus prevent mixing for general observables.

The rate of decay of correlations, or mixing, characterizes the (weak) convergence of ensembles to a statistical steady-state and has been the subject of intense research these past decades. Once again, the latter can be studied from the spectral properties of the semigroups [31]. However, the eigenvalues responsible for such decay, the Ruelle–Pollicott resonances [29, 30], lie inside the unit disk and are thus not accessible from operators on regular functional spaces (here, the resonances refer to the eigenvalues of the finite time operators rather than there infinitesimal generator). Anisotropic Banach spaces of distributions capturing the dynamics of contraction and expansion of the system and on which the semigroups are contracting should instead be considered [31,32,33,34,35,36].Footnote 12

A certain degree of hyperbolicity is necessary to adapt these spaces to the dynamics, and the robustness to perturbations of the spectrum or the convergence of numerical algorithms is not guarantied in general (see [33, 105] and [106, 107], respectively, for results in this direction). In the present study, it is found, however, that the discretization scheme used in Sect. 3.3 yields reasonably good approximations of the resonances.

The spectral properties of the semigroups with respect to an invariant measure \(\mu \) thus allows for the study of the ergodic properties of this measure. On the other hand, for the study of the global stability of this set, operators acting on a neighborhood of positive Lebesgue measure of a chaotic set should be considered.

1.3 Spectral Theory of Global Stability

In Sect. 1, we have advocated on heuristic grounds that the global stability of invariant sets could be studied from the evolution of densities and observables. In this case, however, the transfer (Koopman) operators \(\mathcal {P}_t^m\) (\(\mathcal {U}_t^m\)) with respect to the Lebesgue measure m should be considered. Indeed, this should allow one not only to study the mixing dynamics of an invariant measure, but also the contraction to/escape from such a measure. This has allows for new developments in the theory of the global stability of stationary points and periodic orbits from the behavior of \(\mathcal {P}_t^m\) [41] and \(\mathcal {U}_t^m\) [42]. This approach should also be amenable to chaotic invariant sets. However, the same difficulties as mentioned in the previous “Spectral Theory of Ergodicity and Mixing” section of Appendix A, regarding the functional analytic framework appropriate to capture the resonances inside the unit disk, are encountered. In order to better interpret the numerical results obtained in Sect. 4, let us however give a few comments regarding the eigenfunctions of the Koopman operators \(\mathcal {U}^X_t\) and \(\mathcal {U}^\varLambda _t\) on the space C(X) of continuous functions on the compact metric space (Xd) (e.g. \(X=\mathbb {R}^n\) with d the distance induced by the Euclidean norm) and its restriction to the support \(\varLambda \) of some invariant measure \(\mu \), respectively. Since (Xd) is a normal space and \(\varLambda \) is a closed subset of X, the Tietze–Urysohn Extension Theorem (e.g. [108]) ensures that for any continuous function \(g^\varLambda \) on \(\varLambda \), there exists a continuous function \(g^X\) on X whose restriction to \(\varLambda \) is \(g^\varLambda \). The restriction \(\mathcal {U}^\varLambda _t\) of \(\mathcal {U}^X_t\) on \(C(\varLambda )\) is thus defined such that \(\mathcal {U}^\varLambda _t g^\varLambda = \mathcal {U}^X_t g^X\).

1.3.1 Multiple Attractors

We first discuss the coexistence of multiple attractors with disjoint basins of attraction, (such as in the Lorenz flow for \(\rho _{\mathrm {Homo}}< \rho < \rho _A\)), in terms of multiplicity of the eigenvalue 1. When the state space X is forward invariant, i.e. such that \(X \subset \varPhi _t^{-1} X\) then one has that

$$\begin{aligned} \mathcal {U}^X_t \mathbf {1}_X = \mathbf {1}_{\phi _t^{-1} X \cap X} = \mathbf {1}_{X}, \quad t \ge 0. \end{aligned}$$
(A.7)

Thus, the forward invariance of X implies that the function \(\mathbf {1}_X\), constant on X, is an eigenfunction of the Koopman semigroup associated with the eigenvalue 1. This is for example the case when the flow has a single globally asymptotically stable attractor on X. If instead, there is a single repeller and \(\varPhi _t^{-1} X \subset X\), then \(\mathcal {U}^X_t \mathbf {1}_X = \mathbf {1}_{\phi _t^{-1} X \cap X} = \mathbf {1}_{\phi _t^{-1} X}\) so that \(\mathbf {1}_X\) is no longer an eigenfunction associated with 1.

When several attractors \(\varLambda _1, \ldots , \varLambda _l\) coexist, each basin of attraction \(B(\varLambda _i)\) is forward invariant and \(\mathcal {U}^X_t \mathbf {1}_{B(\varLambda _i)} = \mathbf {1}_{B(\varLambda _i)}, t \ge 0\). Thus, to each attractor \(\varLambda _i\) corresponds a Koopman eigenfunction \(\mathbf {1}_{B(\varLambda _i)}\) associated with a generator eigenvalue zero. However, only l of the \(l+1\) functions \(\mathbf {1}_X, \mathbf {1}_{B(\varLambda _1)}, \ldots , \mathbf {1}_{B(\varLambda _l)}\) are independent, so that there are only l eigenvalues 1. For example, when two attractors coexist, a possibility is to have as independent eigenfunctions \(\mathbf {1}_X\) and \(\mathbf {1}_{B(\varLambda _1)} - \mathbf {1}_{B(\varLambda _2)}\). Adding suitably defined (e.g. Gaussian) noise leads to a unique invariant measure; reducing the intensity of the noise to zero leads in the limit to selecting a special invariant measure constructed as linear combination of all the invariant measure of the deterministic case. Note, however, that the numerical approximations of the basins of attraction may be difficult when the geometry of the boundary is convoluted. The fact that the numerical discretization of the state space leads effectively to introducing some noise in the system might explain why also in the deterministic case one can miss the presence of various coexisting and independent invariant measures.

1.3.2 Correspondence Between the Eigenvalues of \(\mathcal {U}_t^\varLambda \) and \(\mathcal {U}_t^X\)

Let us discuss the correspondence between the eigenvalues of \(\mathcal {U}_t^X\) and \(\mathcal {U}_t^\varLambda \) for a hyperbolic attractor \(\varLambda \). Following [45], we refer to the eigenvalues of \(\mathcal {U}_t^X\) as the unstable resonances and those of the \(\mathcal {U}_t^\varLambda \) which do not correspond to the latter as the stable resonances.

What we show below is strictly applicable only for eigenvalues on the unit circle, as discussed in Remark 4. Nonetheless, we believe that it may be practically relevant in the region near the unit circle, and, using more advanced mathematical tools, could be extended for the unit disk, away form the essential spectrum.

We consider the particular case where the invariant measure \(\mu \) is supported by a uniformly hyperbolic attractor \(\varLambda \subset X\) for the continuously differentiable flow \(\varPhi _t, t \in \mathbb {R}\) on (Xd). For convenience, we will assume that X is the basin of attraction of \(\varLambda \). The global stable manifold \(W_s\) through the point \(x \in X\) can then be characterized topologically by

$$\begin{aligned} W_s(x) = \{y \in X: \quad d(\varPhi _t(x), \varPhi _t(y)) \rightarrow 0, \quad t \rightarrow \infty \}. \end{aligned}$$
(A.8)

It follows directly that \(W_s\) is invariant, i.e. \(\varPhi _t W_s(x) \subset W_s(x)\) and \(\cup _{x \in \varLambda } W_s(x) = X\), since X is the basin of attraction of \(\varLambda \). Moreover, as a consequence of the stable manifold theorem [65, Chap. 6], \(x \mapsto W_s(x)\) is continuous.

Proposition 2

Let \(\varPhi _t, t \in \mathbb {R}\) be a continuously differentiable flow on the compact space X. Let \(\varLambda \) be a uniformly hyperbolic attractor and denote by \(\mathcal {U}_t^\varLambda : C(\varLambda ) \rightarrow C(\varLambda ), t\ge 0,\) the restriction to \(C(\varLambda )\) of the Koopman operator \(\mathcal {U}^X_t: C(X) \rightarrow C(X), g \mapsto g \circ \varPhi _t\). Assume that \(\psi ^\varLambda \) in \(C(\varLambda )\) is an eigenfunction of \(\mathcal {U}_t^\varLambda \) associated with the eigenvalue \(\zeta \in \mathbb {C}\) for some \(t > 0\), i.e. \(\mathcal {U}_t^\varLambda \psi ^\varLambda = \zeta \psi ^\varLambda \). Then the function \(\psi ^X\) such that

$$\begin{aligned} \psi ^X(y) = \psi ^\varLambda (x), \quad \mathrm {whenever~} y \in W_s(x) \end{aligned}$$
(A.9)

is in C(X) and is an eigenfunction for \(\mathcal {U}_t^X\) associated with the eigenvalue \(\zeta \).

In other words, \(\psi ^X\) takes on a leaf of the global stable manifold a constant value given by that of \(\psi ^\varLambda \) on this leaf.

Proof

Let us first verify that \(\psi ^X\) is indeed in C(X). Note first that \(\psi ^X\) is defined on X, since \(W_s = X\). The continuity of \(\psi ^X\) follows from that of \(\psi ^\varLambda \) and \(x \mapsto W_s(x)\). To see this, let \(\{y_n\}\) be a sequence in X converging to y. For each \(y_n\) there is an \(x_n\) in \(\varLambda \) such that \(y_n \in W_s(x_n)\). Thus

$$\begin{aligned} \psi ^X(y_n) = \psi ^\varLambda (x_n) \end{aligned}$$
(A.10)

From the continuity of \(x \mapsto W_s(x)\), the limit x of \(x_n\) exists and is such that \(y \in W_s(x)\) and \(\psi ^X(y) = \psi ^\varLambda (x)\). From the continuity of \(\psi ^\varLambda \), it follows that

$$\begin{aligned} \lim _{n \rightarrow \infty } \psi ^X(y_n) = \lim _{n \rightarrow \infty } \psi ^\varLambda (x_n) = \psi ^\varLambda (x) = \psi ^\varLambda (y), \end{aligned}$$
(A.11)

so that \(\psi ^X\) is continuous on the metric space (Xd).

That \(\psi ^X\) is an eigenfunction follows directly from the invariance of the global stable manifold:

$$\begin{aligned} \mathcal {U}_t^X \psi ^X(y) = \psi ^X(\varPhi _t y)&= \psi ^\varLambda (x)&\mathrm {whenever~} \varPhi _t y \in W_s(x) \end{aligned}$$
(A.12)
$$\begin{aligned}&= \psi ^\varLambda (x)&\mathrm {whenever~} y \in W_s(\varPhi _t^{-1} x) \end{aligned}$$
(A.13)
$$\begin{aligned}&= \psi ^\varLambda (\varPhi _t z)&\mathrm {whenever~} y \in W_s(z) \end{aligned}$$
(A.14)
$$\begin{aligned}&= \zeta \psi ^\varLambda (z)&\mathrm {whenever~} y \in W_s(z) \end{aligned}$$
(A.15)
$$\begin{aligned}&= \zeta \psi ^X(y). \end{aligned}$$
(A.16)

\(\square \)

Thus, to each eigenfunction \(\mathcal {U}_t^\varLambda \) on \(C(\varLambda )\) corresponds an eigenfunction of \(\mathcal {U}_t^X\) on C(X) associated with the same eigenvalue, i.e. the spectrum of \(\mathcal {U}_t^\varLambda \) is a subset of the spectrum of \(\mathcal {U}_t^X\).

Remark 4

There is, however, a major caveat to the applicability of this results. Indeed, it cannot be applied to eigenfunctions associated with eigenvalues inside the unit disk, since \(\mathcal {U}_t^\varLambda \) has a roughening effect due to the contraction on the unstable manifold manifold of \(\varLambda \), backward in time. For that purpose, spaces a distributions should be considered for \(\psi ^\varLambda \).

Remark 5

A similar result holds for the nonuniformly hyperbolic case [66]. However, the stable foliation is then only measurable so that eigenfunctions in spaces of measurable functions should be considered.

Appendix B: Robustness of the Resonances

In this appendix, we explain and give support to the choice of parameters used to obtain the results of Sect. 4 and given in Sect. 3.4.

1.1 Numerical Integration

In this study, the time step of integration is particularly important for the position of the attractor crisis. It appears that for larger time steps, the attractor crisis occurs for values of \(\rho \) smaller than \(\rho _A\). A time step of \(10^{-4}\) allows for the crisis to occur within \(1\%\) of \(\rho _A\), while keeping the numerical integration tractable.

1.2 Discretization

The choice of the grid used to approximate the transfer operators by transition matrices is the most critical step of the numerical application in this study. First, in order for probabilities to be conserved, the grid should cover a bounded set within which any trajectory remains. This condition if fulfilled by a grid covering the ball \(R_o\) (see Sect. 2). A discretization of \(R_o\) is then easily implemented when working in spherical coordinates \((r, \theta , \phi )\).

Second, due to the fine-grained geometry of the eigenvectors associated with chaotic sets, the generator eigenvalues are slow to converge with the grid resolution.Footnote 13 In this study, the focus is, however, on the generator eigenvalues close to the imaginary axis, which are expected to be more robust to perturbations than eigenvalues further from the imaginary axis. The results of a test of convergence with respect to the grid resolution \(n_d\)-by-\(n_d/2\)-by-\(n_d\) is given in Fig. 10. One can see that for \(n_d \ge 400\), the real part of the first nonzero eigenvalue is close to convergence. To get an idea of the value to which this real part would converge for higher resolutions, the dashed orange line represents a least square fit of an exponential \(a e^{b n_d},\) with \(a, b \in \mathbb {R}\), to it. The quality of this fit and the fact that the fitted curve is an exponential converging to zero suggests that this eigenvalue, for \(\rho \approx 24\), would converge to the imaginary axis if the resolution were to be further increased. Eventually, a grid resolution of 400-by-200-by-400 is chosen, allowing for the real part of the first nonzero generator eigenvalue to remain within \(2\%\) of the corresponding value obtained for \(n_d = 500\). The same grid is used to estimate \(\mathbf {P}^\mu _\tau \), for which similar numerical convergence is also observed (not shown here).

Fig. 10
figure 10

Convergence test of the real parts of the five leading generator eigenvalues calculated from \(\mathbf {P}^m_\tau \) with respect to the grid resolution \(n_d\)-by-\(n_d/2\)-by-\(n_d\), for \(\rho = 24\), \(\tau = 0.05\) and from \(3.2 \times 10^{10}\) trajectories

1.3 Transition Time

The transition time \(\tau \) for which the transition matrices \(P_\tau ^m\) and \(P_\tau ^\mu \) are estimated is a key parameter. In theory, the spectral mapping formula (3.8) allows to calculate the generator eigenvalues \(\lambda _k\) from the eigenvalues \(\zeta _k(\tau ), k \ge 0\) for the transfer operator \(\mathcal {P}_\tau ^\eta \), for any transition time \(\tau \). A first issue is, however, that, in taking the complex logarithm divided by \(\tau \) to get the \(\lambda _k\) from the \(\zeta _k(\tau )\), the imaginary part of the \(\lambda _k\) is only known modulo \(2\pi /\tau \). An arbitrary choice of the principal part would then be valid only for true \(\lambda _k\) such that \(|\mathfrak {I}(\lambda _k)| \le \pi /\tau \), so that this window shortens as \(\tau \) is increased (see [109, Sect. 2.4]).

Second, a compromise should be found in order to estimate correctly as many eigenvalues as possible, while only approximating those for which the eigenvectors can be resolved for a given grid. Indeed, the longer the transition time \(\tau \), the smaller \(|\zeta _k(\tau )| = e^{\mathfrak {R}(\lambda _k) \tau }\) for \(\lambda _k\) with a small real part. Thus, in order to be able to estimate the eigenvalue \(\zeta _k(\tau )\) numerically, \(\tau \) should be sufficiently small for \(e^{\mathfrak {R}(\lambda _k) \tau }\) to be larger than a threshold under which numerical errors become important. This threshold depends on several factors such as the sampling, the nonnormality of the transfer operators and roundoff errors [109].

On the other hand, it is not always a good strategy to take \(\tau \) short to approximate as many \(\lambda _k\) far from the imaginary axis as possible. Indeed, eigenvalues further from the imaginary tend to be associated with eigenvectors with more and more changes of sign. For the latter to be resolved, the grid resolution should be higher and higher. However, if for a fixed grid, eigenvalues \(\zeta _k(\tau )\) associated with eigenvectors which cannot be appropriately resolved have not decayed, their imprecise approximation will also have an impact on eigenvalues closer to the imaginary axis. As a rule of thumb, the lower the grid resolution, the larger should \(\tau \) be, so as to approximate only the eigenvalues for which the eigenvectors can be properly resolved at this resolution. For a grid of 400-by-200-by-400, a transition time \(\tau \) of 0.05 time units is found to give of a good compromise.

Fig. 11
figure 11

Convergence test of the real parts of the leading generator eigenvalues of \(\mathbf {P}_\tau ^m\) (left) and \(\mathbf {P}_\tau ^\mu \) (right) with respect to the number of samples \(N_s\), for \(\rho = 24\) and \(\rho = 28\), respectively, and a grid of \(400-by-200-by-400\)

1.4 Number and Length of Trajectories

Ulam’s method relies on the estimation of transition probabilities from time series. The quality of these estimations depend on the sampling. To test the sampling, one strategy is to estimate confidence intervals (see e.g. [86, SI] and [88]). Another approach, followed here, is to directly test the convergence of the generator eigenvalues with respect to the number of samples \(N_s\). This convergence of the transition probabilities and the eigenvalues with \(N_s\) is known to occur at a rate of \(\mathcal {O}(N_s^{1/2})\) [90, 109]. When estimating \(\mathbf {P}_\tau ^m\) from many short time series, the number of samples is given by the number of trajectories. In this case, the robustness of the eigenvalues to \(N_s\) is shown in the left panel in Fig. 11. One can see that a number \(6.4 \times 10^{9}\) of trajectories is more than enough for an estimation on the chosen grid of 400-by-200-400. When estimating \(\mathbf {P}_\tau ^\mu \) from a few long time series, the number of samples is given by the number of trajectories by their lengths \(T_\mathrm {samp}\) divided by their sampling rate. Here, a sampling rate of 100 samples per time unit is used and 24 long trajectories are used (in order to distribute each on a computer thread). The robustness of the eigenvalues to \(T_\mathrm {samp}\) is shown in the right panel in Fig. 11. One can see that a length \(T_\mathrm {samp}\) of \(1 \times 10^{5}\) time units is more than enough for an estimation on the chosen grid of 400-by-200-400.

1.5 Numerical Eigenvalue Problem

We have seen that the fine-grained geometry of the eigenvectors of the Lorenz flow require a large grid resolution in order to achieve numerical convergence, if only for the leading eigenvalues. While, the estimation of the transition matrices from time series is, an embarrassingly parallel problem, which can be easily distributed on several nodes of a calculator, solving the eigenproblem for such large transition matrices is more challenging, both in terms of computations and memory. On the other hand, the sparse structure of the transition matrices allows to use iterative eigensolvers. Here, we have chosen the block Krylov Schur algorithm implemented in the Anasazi package [91] of the Trilinos library. This algorithm is a common and robust choice for such problems and the Anasazi implementation makes it straightforward to distribute it on several nodes of a calculator.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tantet, A., Lucarini, V. & Dijkstra, H.A. Resonances in a Chaotic Attractor Crisis of the Lorenz Flow. J Stat Phys 170, 584–616 (2018). https://doi.org/10.1007/s10955-017-1938-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1938-0

Keywords

Mathematics Subject Classification

Navigation