Skip to main content

Theory and Applications of the Orthogonal Fast Lyapunov Indicator (OFLI and OFLI2) Methods

  • Chapter
  • First Online:
Chaos Detection and Predictability

Part of the book series: Lecture Notes in Physics ((LNP,volume 915))

Abstract

During the last decades the Nonlinear Dynamics field has produced a large number of numerical techniques oriented to the analysis of the behavior of the orbits in different systems. These methods are mainly focused to distinguish chaotic from regular behavior. Among the variational methods, based into the variational equations, we discuss in this paper the so-called Orthogonal Fast Lyapunov Indicator (OFLI and OFLI2) methods that are variants of the FLI method but designed to obtain also some information about the periodic orbits of the systems. We review the OFLI and OFLI2 methods and we show several computational aspects related with avoiding the appearance of spurious structures, with their use in the analysis of regular/chaotic behaviors, but also with the analysis of periodic orbits and regular regions, and with the efficient computation of the solution of the variational equations by means of Taylor series methods. Finally, the methods are shown in several Hamiltonian problems, as well as in several classical dissipative systems, as the Lorenz and Rössler models.

Dedicated to the Memory of Eugenio Barrio (1934–2014)

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    TIDES: a Taylor series Integrator for Differential EquationS (GNU free software). Webpage: http://cody.unizar.es/software.html and http://sourceforge.net/projects/tidesodes/.

References

  1. Galias, Z., Zgliczyński, P.: Computer assisted proof of chaos in the Lorenz equations. Physica D 115, 165–188 (1998)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  2. Tucker, W.: A rigorous ODE solver and Smale’s 14th problem. Found. Comput. Math. 2, 53–117 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  3. Skokos, C.: The Lyapunov characteristic exponents and their computation. In: Souchay, J.J., Dvorak, R. (eds.) Dynamics of Small Solar System Bodies and Exoplanets. Lecture Notes in Physics, vol. 790, pp. 63–135. Springer, Berlin, Heidelberg (2010)

    Chapter  Google Scholar 

  4. Laskar, J.: Frequency analysis for multi-dimensional systems. Global dynamics and diffusion. Physica D 67, 257–281 (1993)

    MATH  MathSciNet  Google Scholar 

  5. Laskar, J.: Frequency analysis of a dynamical system. Celest. Mech. Dyn. Astron. 56, 191–196 (1993)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  6. Cincotta, P.M., Simó, C.: Simple tools to study global dynamics in non-axisymmetric galactic potentials - I. Astron. Astrophys. Suppl. 147, 205–228 (2000)

    Article  ADS  Google Scholar 

  7. Cincotta, P.M., Giordano, C.M., Simó, C.: Phase space structure of multi-dimensional systems by means of the mean exponential growth factor of nearby orbits. Physica D 182, 151–178 (2003)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  8. Froeschlé, C., Lega, E.: On the structure of symplectic mappings. The fast Lyapunov indicator: a very sensitivity tool. Celest. Mech. Dyn. Astron. 78, 167–195 (2000)

    ADS  MATH  Google Scholar 

  9. Guzzo, M., Lega, E., Froeschlé, C.: On the numerical detection of the effective stability of chaotic motions in quasi-integrable systems. Physica D 163, 1–25 (2002)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  10. Skokos, C., Antonopoulos, C., Bountis, T.C., Vrahatis, M.N.: Detecting order and chaos in Hamiltonian systems by the SALI method. J. Phys. A 37, 6269–6284 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  11. Gottwald, G.A., Melbourne, I.: A new test for chaos in deterministic systems. Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 460, 603–611 (2004)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  12. Gottwald, G.A., Melbourne, I.: Testing for chaos in deterministic systems with noise. Physica D 212, 100–110 (2005)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  13. Barrio, R., Shilnikov, A.: Parameter-sweeping techniques for temporal dynamics of neuronal systems: case study of Hindmarsh-Rose model. J. Math. Neurosci. 1(6), 20 (2011)

    MATH  MathSciNet  Google Scholar 

  14. Barrio, R., Shilnikov, A., Shilnikov, L.: Kneadings, symbolic dynamics and painting Lorenz chaos. Int. J. Bifurcat. Chaos 22, 1230016, 24 (2012)

    Google Scholar 

  15. Darriba, L.A., Maffione, N.P., Cincotta, P.M., Giordano, C.M.: Comparative study of variational chaos indicators and ODEs’ numerical integrators. Int. J. Bifurcat. Chaos 22, 1230033, 33 (2012)

    Google Scholar 

  16. Fouchard, M., Lega, E., Froeschlé, C., Froeschlé, C.: On the relationship between fast Lyapunov indicator and periodic orbits for continuous flows. Celest. Mech. Dyn. Astron. 83, 205–222 (2002)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  17. Barrio, R.: Sensitivity tools vs. Poincaré sections. Chaos Solitons Fractals 25, 711–726 (2005)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  18. Barrio, R.: Painting chaos: a gallery of sensitivity plots of classical problems. Int. J. Bifurcat. Chaos 16, 2777–2798 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  19. Barrio, R., Borczyk, W., Breiter, S.: Spurious structures and chaos indicators. Chaos Solitons Fractals 40(4), 1697–1714 (2007)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  20. Barrio, R., Blesa, F., Lara, M.: VSVO formulation of the Taylor method for the numerical solution of ODEs. Comput. Math. Appl. 50, 93–111 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  21. Barrio, R.: Sensitivity analysis of ODE’s/DAE’s using the Taylor series method. SIAM J. Sci. Comput. 27, 1929–1947 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  22. Hénon, M., Heiles, C.: The applicability of the third integral of motion: some numerical experiments. Astron. J. 69, 73–79 (1964)

    Article  ADS  MathSciNet  Google Scholar 

  23. Geist, K., Parlitz, U., Lauterborn, W.: Comparison of different methods for computing Lyapunov exponents. Prog. Theor. Phys. 83, 875–893 (1990)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  24. Dieci, L., Van Vleck, E.S.: Lyapunov and other spectra: a survey. In: Collected Lectures on the Preservation of Stability Under Discretization (Fort Collins, CO, 2001), pp. 197–218. SIAM, Philadelphia, PA (2002)

    Google Scholar 

  25. Thiffeault, J.L., Boozer, A.H.: Geometrical constraints on finite-time Lyapunov exponents in two and three dimensions. Chaos 11, 16–28 (2001)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  26. Osedelec, V.I.: A multiplicative ergodic theorem. Lyapunov characteristic numbers for dynamical systems. Trans. Moscow Math. Soc. 19, 197–231 (1968)

    Google Scholar 

  27. Legras, B., Vautard, R.: A guide to Liapunov vectors. In: Palmer, T. (ed.) Predictability. ECWF Seminar. vol. I, pp. 135–146. ECMWF, Reading (1996)

    Google Scholar 

  28. Haken, H.: At least one Lyapunov exponent vanishes if the trajectory of an attractor does not contain a fixed point. Phys. Lett. A 94, 71–72 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  29. Meyer, H.D.: Theory of the Liapunov exponents of Hamiltonian systems and a numerical study on the transition from regular to irregular classical motion. J. Chem. Phys. 84, 3147–3161 (1986)

    Article  ADS  MathSciNet  Google Scholar 

  30. Yamaguchi, Y.Y., Iwai, T.: Geometric approach to Lyapunov analysis in Hamiltonian dynamics. Phys. Rev. E (3) 64, 066206, 16 (2001)

    Google Scholar 

  31. Grond, F., Diebner, H.H., Sahle, S., Mathias, A., Fischer, S., Rossler, O.E.: A robust, locally interpretable algorithm for Lyapunov exponents. Chaos Solitons Fractals 16, 841–852 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  32. Guzzo, M., Lega, E.: The numerical detection of the Arnold web and its use for long-term diffusion studies in conservative and weakly dissipative systems. Chaos 23, 023124 (2013)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  33. Barrio, R., Blesa, F.: Systematic search of symmetric periodic orbits in 2DOF Hamiltonian systems. Chaos Solitons Fractals 41, 560–582 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  34. Abad, A., Barrio, R., Dena, A.: Computing periodic orbits with arbitrary precision. Phys. Rev. E 84, 016701 (2011)

    Article  ADS  Google Scholar 

  35. Barrio, R., Blesa, F., Serrano, S.: Fractal structures in the Hénon-Heiles Hamiltonian. Europhys. Lett. 82, 10003 (2008)

    Article  Google Scholar 

  36. Barrio, R., Blesa, F., Serrano, S.: Bifurcations and chaos in Hamiltonian systems. Int. J. Bifurcat. Chaos 20, 1293–1319 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  37. Barrio, R., Blesa, F., Serrano, S.: Qualitative analysis of the (N + 1)-body ring problem. Chaos Solitons Fractals 36, 1067–1088 (2008)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  38. Barrio, R., Blesa, F., Elipe, A.: On the use of chaos indicators in rigid-body motion. J. Astronaut. Sci. 54, 359–368 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  39. Barrio, R., Blesa, F., Serrano, S.: Periodic, escape and chaotic orbits in the Copenhagen and the (n + 1)-body ring problems. Commun. Nonlinear Sci. Numer. Simul. 14, 2229–2238 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  40. Blesa, F., Piasecki, S., Dena, Á., Barrio, R.: Connecting symmetric and asymmetric families of periodic orbits in squared symmetric Hamiltonians. Int. J. Mod. Phys. C 23, 1250014, 22 (2012)

    Google Scholar 

  41. Blesa, F., Seoane, J.M., Barrio, R., Sanjuán, M.A.F.: To escape or not to escape, that is the question—perturbing the Hénon-Heiles Hamiltonian. Int. J. Bifurcat. Chaos 22, 1230010, 9 (2012)

    Google Scholar 

  42. Euler, L.: Theoria motuum lunae (E418) Petrop. Parisin. Lond. (1772)

    Google Scholar 

  43. Jacobi, C.G.J.: Comp. Rend. 3, 59–61 (1836)

    Google Scholar 

  44. Andrle, P.: A third integral of motion in a system with a potential of the fourth degree. Bull. Astron. Inst. Czechoslov. 17, 169–175 (1966)

    ADS  MATH  Google Scholar 

  45. Armbruster, D., Guckenheimer, J., Kim, S.h.: Chaotic dynamics in systems with square symmetry. Phys. Lett. A 140, 416–420 (1989)

    MathSciNet  Google Scholar 

  46. Rucklidge, A.: Global bifurcations in the Takens-Bogdanov normal form with D4 symmetry near the O(2) limit. Phys. Lett. A 284, 99–111 (2001)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  47. Proctor, M.R.E., Weiss, N.O.: Magnetoconvection. Rep. Prog. Phys. 45, 1317 (1982)

    Article  ADS  Google Scholar 

  48. Barrio, R., Blesa, F., Serrano, S.: Bifurcations and safe regions in open Hamiltonians. New J. Phys. 11, 053004 (2009)

    Article  ADS  Google Scholar 

  49. Barrio, R., Serrano, S.: A three-parametric study of the Lorenz model. Physica D 229, 43–51 (2007)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  50. Barrio, R., Serrano, S.: Bounds for the chaotic region in the Lorenz model. Physica D 238, 1615–1624 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  51. Barrio, R., Blesa, F., Serrano, S.: Behavior patterns in multiparametric dynamical systems: Lorenz model. Int. J. Bifurcat. Chaos 22, 1230019, 14 (2012)

    Google Scholar 

  52. Barrio, R., Serrano, S.: Qualitative analysis of the Rössler equations: bifurcations of limit cycles and chaotic attractors. Physica D 238, 1087–1100 (2009)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  53. Barrio, R., Blesa, F., Serrano, S., Shilnikov, A.: Global organization of spiral structures in biparameter space of dissipative systems with Shilnikov saddle-foci. Phys. Rev. E 84, 035201 (2011)

    Article  ADS  Google Scholar 

  54. Barrio, R., Blesa, F., Serrano, S.: Topological changes in periodicity hubs of dissipative systems. Phys. Rev. Lett. 108, 214102 (2012)

    Article  ADS  Google Scholar 

  55. Serrano, S., Barrio, R., Dena, A., Rodríguez, M.: Crisis curves in nonlinear business cycles. Commun. Nonlinear Sci. Numer. Simul. 17, 788–794 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  56. Lorenz, E.: Deterministic nonperiodic flow. J. Atmos. Sci. 20, 130–141 (1963)

    Article  ADS  Google Scholar 

  57. Doedel, E.J., Krauskopf, B., Osinga, H.M.: Global bifurcations of the Lorenz manifold. Nonlinearity 19, 2947–2972 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  58. Dullin, H.R., Schmidt, S., Richter, P.H., Grossmann, S.K.: Extended phase diagram of the Lorenz model. Int. J. Bifurcat. Chaos 17, 3013–3033 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  59. Guckenheimer, J., Holmes, P.: Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Applied Mathematical Sciences, vol. 42. Springer, New York (1990)

    Google Scholar 

  60. Shil′nikov, A.L., Shil′nikov, L.P., Turaev, D.V.: Normal forms and Lorenz attractors. Int. J. Bifurcat. Chaos 3, 1123–1139 (1993)

    Google Scholar 

  61. Viana, M.: What’s new on Lorenz strange attractors? Math. Intelligencer 22, 6–19 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  62. Saltzman, B.: Finite amplitude free convection as an initial value problem - 1. J. Atmos. Sci. 19, 329–341 (1962)

    Article  ADS  Google Scholar 

  63. Rössler, O.E.: An equation for continuous chaos. Phys. Lett. A 57, 397–398 (1976)

    Article  ADS  Google Scholar 

  64. Gaspard, P., Kapral, R., Nicolis, G.: Bifurcation phenomena near homoclinic systems: a two-parameter analysis. J. Stat. Phys. 35, 697–727 (1984)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  65. Gallas, J.A.C.: The structure of infinite periodic and chaotic hub cascades in phase diagrams of simple autonomous flows. Int. J. Bifurcat. Chaos 20, 197–211 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  66. Guckenheimer, J., Meloon, B.: Computing periodic orbits and their bifurcations with automatic differentiation. SIAM J. Sci. Comput. 22, 951–985 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  67. Corliss, G., Chang, Y.F.: Solving ordinary differential equations using Taylor series. ACM Trans. Math. Softw. 8, 114–144 (1982)

    Article  MATH  MathSciNet  Google Scholar 

  68. Barrio, R., Rodríguez, M., Blesa, F.: Computer-assisted proof of skeletons of periodic orbits. Comput. Phys. Commun. 183, 80–85 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  69. Simó, C.: Dynamical systems, numerical experiments and super-computing. Mem. Real Acad. Cienc. Artes Barcelona 61, 3–36 (2003)

    MathSciNet  Google Scholar 

  70. Dena, Á., Rodríguez, M., Serrano, S., Barrio, R.: High-precision continuation of periodic orbits. Abstr. Appl. Anal. 12 (2012). Art. ID 716024

    Google Scholar 

  71. Barrio, R.: Performance of the Taylor series method for ODEs/DAEs. Appl. Math. Comput. 163, 525–545 (2005)

    MATH  MathSciNet  Google Scholar 

  72. Abad, A., Barrio, R., Blesa, F., Rodríguez, M.: Algorithm 924: TIDES, a Taylor series integrator for differential equationS. ACM Trans. Math. Softw. 39, 5:1–5:28 (2012)

    Google Scholar 

  73. Jorba, À., Zou, M.: A software package for the numerical integration of ODEs by means of high-order Taylor methods. Exp. Math. 14, 99–117 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  74. Hairer, E., Nørsett, S.P., Wanner, G.: Solving Ordinary Differential Equations. I. Springer Series in Computational Mathematics, vol. 8. Springer, Berlin (1993)

    Google Scholar 

  75. Fousse, L., Hanrot, G., Lefèvre, V., Pélissier, P., Zimmermann, P.: MPFR: a multiple-precision binary floating-point library with correct rounding. ACM Trans. Math. Softw. 33, 13:1–13:15 (2007)

    Google Scholar 

  76. Abad, A., Barrio, R., Blesa, F., Rodríguez, M.: TIDES tutorial: integrating ODEs by using the Taylor series method. Monografías de la Academia de Ciencias de la Universidad de Zaragoza 36, 1–116 (2011)

    MATH  Google Scholar 

  77. Hairer, E., McLachlan, R.I., Razakarivony, A.: Achieving Brouwer’s law with implicit Runge-Kutta methods. BIT 48, 231–243 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  78. Brouwer, D.: On the accumulation of errors in numerical integration. Astron. J. 30, 149–153 (1937)

    Article  ADS  Google Scholar 

  79. Simó, C.: Global dynamics and fast indicators. In: Global Analysis of Dynamical Systems, pp. 373–389. Institute of Physics, Bristol (2001)

    Google Scholar 

  80. Sharp, P.W.: N-body simulations: the performance of some integrators. ACM Trans. Math. Softw. 32, 375–395 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  81. Grazier, K.R., Newman, W.I., Hyman, J.M., Sharp, P.W.: Long simulations of the Solar System: Brouwer’s law and chaos. ANZIAM J. 46, C1086–C1103 (2004/05)

    Google Scholar 

  82. Grazier, K.R., Newman, W.I., Hyman, J.M., Sharp, P.W., Goldstein, D.J.: Achieving Brouwer’s law with high-order Störmer multistep methods. ANZIAM J. 46, C786–C804 (2004/05)

    Google Scholar 

  83. Higham, N.J.: Accuracy and Stability of Numerical Algorithms, 2nd edn. SIAM, Philadelphia, PA (2002)

    Book  MATH  Google Scholar 

Download references

Acknowledgements

The author thanks his colleagues and friends Dr. Fernando Blesa and Dr. Sergio Serrano for many interesting discussions and common work on this subject. The author thanks the referees for their help in improving the paper. The author has been supported during this research by the Spanish Research Grant MTM2012-31883, MTM2015-64095-P and by Gobierno de Aragon and Fondo Social Europeo.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Roberto Barrio .

Editor information

Editors and Affiliations

Appendix: Adaptive ODE Integrator—Taylor Series Method

Appendix: Adaptive ODE Integrator—Taylor Series Method

A basic feature in using any chaos indicator is to obtain numerically solutions of ODE systems and variational equations. In Dynamical Systems there are also more requirements, for example, in the process of determination of periodic orbits we obviously have to integrate the differential system, normally for a short time, with very high precision, especially for highly unstable periodic orbits. Moreover, in the study of the bifurcations and stability of periodic orbits we also have to integrate the first order variational equations using as initial conditions the identity matrix as also occurs in the use of chaos indicators. To reach this goal we may, obviously, use any numerical ODE method like, for example, a Runge–Kutta method. The last few years, in the computational dynamics community [66] one of the preferred methods is the Taylor series method.

The Taylor method is one of the oldest numerical methods for solving ordinary differential equations but it is scarcely used in the numerical analysis community. Its formulation is quite simple [67]. Let us consider the initial value problem \(\dot{\mathbf{y}} = \mathbf{f}(t,\,\mathbf{y})\). Now, the value of the solution at t i (that is, y(t i )) is approximated by y i from the nth degree Taylor series of y(t) at t = t i (the function f has to be a smooth function). So, denoting \(h_{i} = t_{i} - t_{i-1}\),

$$\displaystyle{\begin{array}{rcl} \mathbf{y}(t_{0})& =:&\mathbf{y}_{0}, \\ \mathbf{y}(t_{i})& \simeq &\mathbf{y}_{i-1} + \mathbf{f}(t_{i-1},\mathbf{y}_{i-1})\,h_{i} + \frac{1} {2!}\,\frac{d\mathbf{f}(t_{i-1},\mathbf{y}_{i-1})} {dt} \,h_{i}^{2}+\ldots \\ & &\ldots + \frac{1} {n!}\,\frac{d^{n-1}\mathbf{f}(t_{i-1},\mathbf{y}_{i-1})} {dt^{n-1}} \,h_{i}^{n}\,\,\, =:\,\,\, \mathbf{y}_{ i}.\end{array} }$$

Therefore, the problem is reduced to the determination of the Taylor coefficients \(\{1/(\,j + 1)!\,d^{j}\mathbf{f}/dt^{j}\}\). This may be done quite efficiently by means of the automatic differentiation (AD) techniques (for more details see [20]). Note that the Taylor method has several good features; one of them is that it gives directly a dense output in the form of a power series being therefore quite useful when an event location criteria may be used (as in the computation of Poincaré sections), it can be formulated as an interval method giving guaranteed integration methods (used, by instance, in the computer assisted proof of chaos [1] and skeletons of periodic orbits [68]), Taylor methods may manage directly high order differential equations just taking into account that the Taylor coefficients for the solution and its derivatives are evidently related, Taylor methods of degree n are also of order n and so Taylor methods of high degree give us numerical methods of high order (therefore, they are very useful for high-precision solution of ODEs, as needed, for example, in some fine studies in dynamical systems [69] and in the computation of unstable periodic orbits [34, 70]).

Just as a short look at the practical implementation of the Taylor series method we remark that in the literature there are efficient variable-stepsize variable-order (VSVO) formulations. For example, in [20, 71, 72] the variable-stepsize formulation is based on the error estimator using the last two coefficients and gives the following stepsize prediction

$$\displaystyle{h_{i+1} = \mathtt{fac}\cdot \min \left \{\,\left ( \frac{\mathtt{Tol}} {\|\{ \frac{1} {(n-1)!}\,\mathbf{f}^{(n-2)}(t_{i})\|_{\infty }}\right )^{ \frac{1} {n-1} },\,\left ( \frac{\mathtt{Tol}} {\| \frac{1} {n!}\,\mathbf{f}^{(n-1)}(t_{i})\|_{\infty }}\right )^{ \frac{1} {n} }\,\right \}}$$

where fac is a safety factor and Tol the user error tolerance. A very simple order selection that only depends on the user error tolerance is given [73] by the formula \(n(\mathtt{Tol}) = -\frac{1} {2}\,\ln \mathtt{Tol}\). See [20, 71] for a more extensive analysis and comparison with variable-stepsize variable-order formulations of the Taylor method. In Fig. 3.15 we present some comparisons on the Hénon–Heiles problem with initial conditions (x 0, y 0, X 0, Y 0) = (0, 0. 52, 0. 371956090598519, 0) and E = 0. 157494996 in the time interval [0, 200] using the Taylor method (software TIDES [72]) and the well established codes dop853 and odex developed by Hairer and Wanner [74]. These codes are based on an explicit Runge–Kutta of order 8(5,3) given by Dormand and Prince with stepsize control and dense output and the extrapolation method, respectively. All the methods are compared only in double and quadruple precision using the Lahey LF 95 compiler (fortran) because the dop853 cannot be directly used in multiple precision. The multiple-precision tests are done using C++ and the GMP and MPFR [75] multiple precision packages. From Fig. 3.15 we note that for low precision the dop853 code is a bit faster but when the precision demands are increased the Taylor method is by far the fastest, being for very high precision the only reliable method. Moreover, we can appreciate the different slope of the variable order method (Taylor method) and the fixed order one (dop853), being clear that for high precision the variable order schemes become the more competitive because they are more versatile.

Fig. 3.15
figure 15

Comparison of the CPU time in seconds vs. relative error in the numerical integration of a KAM orbit of the Hénon–Heiles problem using two well established codes, dop853 (explicit Runge–Kutta method) and odex (extrapolation method), and the Taylor series method (TIDES code) with variable-order and variable-stepsize in double-precision (DP), quadruple-precision (QP) and multiple-precision (MP)

For the computation of the OFLI and OFLI2 we are interested not only in the differential equations but also in the variational equations. In order to avoid their explicit generation we have devised [21] an alternative that permits us to obtain the solution of the variational equations without computing them explicitly. Therefore, we have to obtain a numerical solution of y(t) and \(\mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{y}(t)\), being \(\mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{y}(t)\) the Lie derivative of the solution y(t) with respect to the vector δ y(t 0) (that is, in this case the directional derivative). Note that the partial derivatives of the solution with respect to the initial conditions are given by

$$\displaystyle{\varPi =\big (\,\mathcal{L}_{\mathbf{e}_{1}}\mathbf{y}(t)\,\mid \,\mathcal{L}_{\mathbf{e}_{2}}\mathbf{y}(t)\,\mid \,\ldots \,\mid \,\mathcal{L}_{\mathbf{e}_{n}}\mathbf{y}(t)\,\big)}$$

with (e 1, e 2, , e n ) the canonical base of \(\mathbb{R}^{n}\).

The Taylor series method computes the Taylor series of the solution of the differential equation and the Taylor series of the partial derivatives of the solution

$$\displaystyle{\begin{array}{rcl} \delta \mathbf{y}(t_{i})& =& \frac{\partial \mathbf{y}(t_{i})} {\partial \mathbf{y}(t_{0})} \cdot \delta \mathbf{y}(t_{0}) = \mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{y}(t_{i}) \\ & \simeq &\mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{y}(t_{i-1}) + \mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{f}(t_{i-1})\,h_{i} + \frac{1} {2!}\,\mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{f}^{(2)}(t_{ i-1})\,h_{i}^{2} \\ & & +\ldots + \frac{1} {n!}\,\mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{f}^{(n-1)}(t_{ i-1})\,h_{i}^{n}. \end{array} }$$

We now may compute the coefficients \(1/(\,j + 1)!\,\mathcal{L}_{\delta \mathbf{y}(t_{0})}\mathbf{f}^{(\,j)}(t_{i-1})\) by rules of automatic differentiation of the elementary functions (±, ×, ∕, \(\ln\), \(\sin\), ) obtained in [21]. Automatic differentiation gives a recursive procedure to obtain the numerical value of the reiterated derivatives of the elementary functions at a given point. We present here, as example, the rules for the sum, product by a constant, product, division and real power of functions (see [21] for the complete list of rules of any elementary operation):

Proposition

If \(f(t,\,\mathbf{y}(t)),g(t,\,\mathbf{y}(t)): (t,\,\mathbf{y}) \in \mathbb{R}^{s+1}\mapsto \mathbb{R}\) are functions of class \(\mathcal{C}^{n}\) and given a vector \(\mathbf{v} \in \mathbb{R}^{s}\) , we denote

$$\displaystyle{f^{[\,j,\,0]}:= \frac{1} {j!}\,\frac{d^{j}f(t)} {dt^{j}},\qquad f^{[\,j,\,1]}:= \frac{1} {j!}\,\mathcal{L}_{\mathbf{v}}f^{(\,j)},}$$

that is, the jth Taylor coefficient of the function f(t,  y (t)) and of its Lie derivative with respect to v , respectively. Then, we have

  1. (i)

    If h(t) = f(t) ± g(t) then h [n, i] = f [n, i] ± g [n, i] .

  2. (ii)

    If h(t) = α f(t) with \(\alpha \in \mathbb{R}\) then h [n, i] = α f [n, i] .

  3. (iii)

    If h(t) = f(t) ⋅ g(t) then

    $$\displaystyle{\begin{array}{lcl} h^{[n,\,0]} & =&\sum _{j=0}^{n}f^{[n-j,\,0]} \cdot g^{[\,j,\,0]}, \\ h^{[n,\,1]} & =&\sum _{j=0}^{n}\big(\,f^{[n-j,\,0]} \cdot g^{[\,j,\,1]} + f^{[n-j,\,1]} \cdot g^{[\,j,\,0]}\big).\end{array} }$$
  4. (iv)

    If \(h(t) = f(t)/g(t)\) then

    $$\displaystyle{\begin{array}{rcl} h^{[n,\,0]} & =& \frac{1} {g^{[0,0]}}\,\bigg(\,f^{[n,\,0]} -\sum _{ j=0}^{n-1}h^{[\,j,\,0]} \cdot f^{[n-j,\,0]}\bigg), \\ h^{[n,\,1]} & =& \frac{1} {g^{[0,0]}}\,\bigg\{f^{[n,\,1]} - h^{[n,\,0]} \cdot f^{[0,\,1]} \\ & & -\sum _{j=0}^{n-1}\big(h^{[\,j,\,0]} \cdot f^{[n-j,\,1]} + h^{[\,j,\,1]} \cdot f^{[n-j,\,0]}\big)\bigg\}.\end{array} }$$
  5. (v)

    If h(t) = f(t) α with \(\alpha \in \mathbb{R}\) and f [0,0] ≠0, then

    $$\displaystyle{\begin{array}{lcl} h^{[0,\,0]} & =&(\,f^{[0,0]}(t))^{\alpha }, \\ h^{[n,\,0]} & =& \frac{1} {nf^{[0,0]}}\,\sum _{j=0}^{n-1}\big(n\,\alpha - j(\alpha +1)\big)\,h^{[\,j,\,0]} \cdot f^{[n-j,\,0]}, \\ h^{[0,\,1]} & =& \frac{1} {f^{[0,0]}}\,\alpha \,h^{[0,\,0]} \cdot f^{[0,\,1]}, \\ h^{[n,\,1]} & =& \frac{1} {nf^{[0,0]}}\,\bigg\{ - nh^{[n,\,0]} \cdot f^{[0,\,1]} \\ & & +\sum _{ j=0}^{n-1}\big(n\,\alpha - j(\alpha +1)\big)\,\big(h^{[\,j,\,0]} \cdot f^{[n-j,\,1]} + h^{[\,j,\,1]} \cdot f^{[n-j,\,0]}\big)\bigg\}.\end{array} }$$

The use of high-precision numerical integrators in the determination of periodic orbits is justified, for instance, by the search of highly unstable periodic orbits [34].

In Fig. 3.16 we show some comparisons for the Lorenz model (3.9) in double precision, all obtained with the code TIDES using the traditional way to compute the solution of the variational equations (VAR), that is writing them explicitly, and with the use of the extended Taylor series method (ETS) and using TIDES with this capability (using the extended Automatic Differentiation rules of Proposition 3.5). In the pictures we present computational relative error vs. CPU time diagrams in seconds. The extended Taylor series method is the fastest option with a low difference, but the most important thing is that the difference in the formulation is very high. Everyone knows how cumbersome is to write variational equations of order one, two and higher!! The picture is done for computing the complete order two and just the partial derivative 2 x∂ x 0 2. Note that TIDES can compute also sensitivities with respect to parameters of a system, not only with respect to the initial conditions.

Fig. 3.16
figure 16

Computational relative error in the computation of sensitivities vs. CPU time diagrams in seconds for the Lorenz model in double-precision using TIDES code using the extended Taylor series method for the solution of the variational equations (ETS) or just the standard Taylor series method with explicit formulation of the variational equations (VAR)

To end this section we remark that the use of the Taylor series method is currently helped by the free available new state-of-the-art numerical library TIDES (Taylor Integrator of Differential EquationS) that has just been developed by Profs. Abad, Barrio, Blesa and Rodríguez [72, 76]. The reader can contact the authors to obtain the software.Footnote 1

Nowadays it is quite standard to preserve several geometric properties of the differential systems by means of “geometric integrators”. This kind of methods are specially useful when we want to solve a problem with not very high precision but with a “constant” value of the energy, for instance. The problem for very long numerical integrations is that it doesn’t matter how you perform the integration, finally the rounding errors of the computer will affect the integration, giving an increment of the error in the geometric object [77]. The optimal error in these quantities was studied first by Brouwer [78], who established that the error in energy grows at least as \(\mathcal{O}(t^{1/2})\). This error is obtained for long integrations of careful used symplectic integrators [77] or when one is able to suppress the truncation error in any numerical integrator (and also with a careful use, of course). In other circumstances we may observe a typical linear growing \(\mathcal{O}(t)\). In the case of the positions, we will have a root-mean-squared (RMS) error \(\mathcal{O}(t^{3/2})\) in the best case, and a typical error \(\mathcal{O}(t^{2})\) (as in any non-symplectic RK code).

Now, we just show how easy is to eliminate the truncation error in the Taylor series method, and so, in TIDES. The advantage is that using the error estimator of the Taylor method, as they are based in just studying some Taylor coefficients, we may use any tolerance level. A completely different situation occurs when our error estimator is based on the substraction of two similar expressions (like in some formulations of embedded Runge–Kutta pairs where the local error estimator is given by the substraction of the solution of two methods of different order) that makes impossible to use them for tolerances lower than the rounding error due to the “catastrophic digit cancelation”. So, if we fix the tolerance far below the roundoff unit of the computer we, in theory, can control the truncation error. This technique has been used previously by the group of Carles Simó [79] and by others [8082]. We have to combine this technique with a “compensated sum” formulation [83] of the time increment as we use variable stepsize strategies (in contrast with symplectic integrators that have to use fixed stepsize implementations). So, the truncation-free formulation can be described as:

(use TOL ≪ u, with u = the roundoff unit) + (“compensated sum”)

TIDES uses compensated sum in some stages of the method, so, if we want to preserve some geometric properties of the systems we just have to fix a low enough tolerance level. Obviously, this approach is computationally more expensive than other approaches and it is valid only if you also look for high precision numerical results.

In Fig. 3.17 we present the evolution of the error using TIDES with the truncated-free formulation. It is clear that this approach permits to achieve the optimal Brouwer’s law (see Fig. 3.17), like well-programmed symplectic integrators [77], but it can be used in variable-stepsize formulations being therefore a quite flexible approach.

Fig. 3.17
figure 17

Evolution of the error in the position and in the Energy of one orbit of the Hénon–Heiles system using a tolerance lower than the unit roundoff of the computer (truncation-free formulation). For the position we show the evolution of the error for the x and X variables of the Hamiltonian (3.6)

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Barrio, R. (2016). Theory and Applications of the Orthogonal Fast Lyapunov Indicator (OFLI and OFLI2) Methods. In: Skokos, C., Gottwald, G., Laskar, J. (eds) Chaos Detection and Predictability. Lecture Notes in Physics, vol 915. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-662-48410-4_3

Download citation

  • DOI: https://doi.org/10.1007/978-3-662-48410-4_3

  • Published:

  • Publisher Name: Springer, Berlin, Heidelberg

  • Print ISBN: 978-3-662-48408-1

  • Online ISBN: 978-3-662-48410-4

  • eBook Packages: Physics and AstronomyPhysics and Astronomy (R0)

Publish with us

Policies and ethics