Skip to main content
Log in

The effects of time-dependent dissipation on the basins of attraction for the pendulum with oscillating support

  • Original Paper
  • Published:
Nonlinear Dynamics Aims and scope Submit manuscript

Abstract

We consider a pendulum with vertically oscillating support and time-dependent damping coefficient which varies until reaching a finite final value. Although it is the final value which determines which attractors eventually exist, the sizes of the corresponding basins of attraction are found to depend strongly on the full evolution of the dissipation. In particular, we investigate numerically how dissipation monotonically varying in time changes the sizes of the basins of attraction. It turns out that, in order to predict the behaviour of the system, it is essential to understand how the sizes of the basins of attraction for constant dissipation depend on the damping coefficient. For values of the parameters where the systems can be considered as a perturbation of the simple pendulum, which is integrable, we characterise analytically the conditions under which the attractors exist and study numerically how the sizes of their basins of attraction depend on the damping coefficient. Away from the perturbation regime, a numerical study of the attractors and the corresponding basins of attraction for different constant values of the damping coefficient produces a much more involved scenario: changing the magnitude of the dissipation causes some attractors to disappear either leaving no trace or producing new attractors by bifurcation, such as period doubling and saddle-node bifurcation. Finally, we pass to the case of an initially non-constant damping coefficient, both increasing and decreasing to some finite final value, and we numerically observe the resulting effects on the sizes of the basins of attraction: when the damping coefficient varies slowly from a finite initial value to a different final value, without changing the set of attractors, the slower the variation the closer the sizes of the basins of attraction are to those they have for constant damping coefficient fixed at the initial value. Furthermore, if during the variation of the damping coefficient attractors appear or disappear, remarkable additional phenomena may occur. For instance, it can happen that, in the limit of very large variation time, a fixed point asymptotically attracts the entire phase space, up to a zero-measure set, even though no attractor with such a property exists for any value of the damping coefficient between the extreme values.

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
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26

Similar content being viewed by others

References

  1. Bartuccelli, M.V., Berretti, A., Deane, J.H.B., Gentile, G., Gourley, S.A.: Selection rules for periodic orbits and scaling laws for a driven damped quartic oscillator. Nonlinear Anal. Real World Appl. 9(5), 1966–1988 (2008)

    Article  MATH  MathSciNet  Google Scholar 

  2. Bartuccelli, M.V., Deane, J.H.B., Gentile, G.: Globally and locally attractive solutions for quasi-periodically forced systems. J. Math. Anal. Appl. 328(1), 699–714 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  3. Bartuccelli, M.V., Deane, J.H.B., Gentile, G.: Attractiveness of periodic orbits in parametrically forced systems with time-increasing friction. J. Math. Phys. 53 (2012), no.10, 102703, p. 27.

  4. Bartuccelli, M.V., Deane, J.H.B., Gentile, G., Gourley, S.A.: Global attraction to the origin in a parametrically-driven nonlinear oscillator. Appl. Math. Comput. 153(1), 1–11 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  5. Bartuccelli, M.V., Gentile, G., Georgiou, K.V.: On the dynamics of a vertically driven damped planar pendulum. Proc. R. Soc. Lond. A 457(2016), 3007–3022 (2001)

    Article  MATH  MathSciNet  Google Scholar 

  6. Bartuccelli, M.V., Gentile, G., Georgiou, K.V.: On the stability of the upside-down pendulum with damping. Proc. R. Soc. Lond. A 458(2018), 255–269 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  7. Bartuccelli, M.V., Gentile, G., Georgiou, K.V.: KAM theory, Lindstedt series and the stability of the upside-down pendulum. Discrete Contin. Dyn. Syst. 9(2), 413–426 (2003)

    MATH  MathSciNet  Google Scholar 

  8. Bowman, F.: Introduction to Elliptic Functions with Applications. English Universities Press, London (1953)

    MATH  Google Scholar 

  9. Brizard, A.J.: Jacobi zeta function and action-angle coordinates for the pendulum. Commun. Nonlinear Sci. Numer. Simul. 18, 511–518 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  10. Brown, K.: Topics in calculus and probability, Self-published on Lulu (2012), 64–77. Also available at: www.mathpages.com

  11. Byrd, P.F., Friedman, M.D.: Handbook of Elliptic Integrals for Engineers and Scientists, 2nd edn. Springer, New York-Heidelberg (1971)

    Book  MATH  Google Scholar 

  12. Cannon, J.R., Miller, K.: Some problems in numerical analytic continuation. J. Soc. Indust. Appl. Math. Ser. B 2(1), 87–98 (1965)

    Article  MATH  MathSciNet  Google Scholar 

  13. Chow, ShN, Hale, J.K.: Methods of Bifurcation Theory, Grundlehren der Mathematischen Wissenschaften 251. Springer, New York-Berlin (1982)

    Google Scholar 

  14. Fasano, A., Marmi, S.: Analytical Mechanics. Oxford University Press, Oxford (2006)

    MATH  Google Scholar 

  15. Feudel, U., Grebogi, C.: Why are chaotic attractors rare in multistable systems, Phys. Rev. E 91 (2003), 134102, p. 4.

  16. Feudel, U., Grebogi, C., Hunt, B.R., Yorke, J.A.: Map with more than 100 coexisting low-period periodic attractors. Phys. Rev. E 54, 71–81 (1996)

    Article  MathSciNet  Google Scholar 

  17. Gallicchio, E., Egorov, S.A., Berne, B.J.: On the application of numerical analytic continuation methods to the study of quantum mechanical vibrational relaxation processes. J. Chem. Phys. 109(18), 7745–7755 (1998)

    Google Scholar 

  18. Gentile, G., Bartuccelli, M., Deane, J.: Bifurcation curves of subharmonic solutions and Melnikov theory under degeneracies. Rev. Math. Phys. 19(3), 307–348 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  19. Gentile, G., Gallavotti, G.: Degenerate elliptic resonances. Commun. Math. Phys. 257(2), 319–362 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  20. Ghil, M., Wolansky, G.: Non-Hamiltonian perturbations of integrable systems and resonance trapping. SIAM J. Appl. Math. 52(4), 1148–1171 (1992)

    Article  MATH  MathSciNet  Google Scholar 

  21. Guckenheimer, J., Holmes, Ph: Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields, Revised and Corrected Reprint of the 1983 original, Applied Mathematical Sciences 42. Springer, New York (1990)

    Google Scholar 

  22. Jordan, D.W., Smith, P.: Nonlinear Ordinary Differential Equations, An Introduction for Scientists and Engineers, 4th edn. Oxford University Press, Oxford (2007)

    MATH  Google Scholar 

  23. Kodaira, K.: Complex Analysis. Cambridge University Press, Cambridge (2007)

    Book  MATH  Google Scholar 

  24. Krasovsky, N.N.: Problems of the Theory of Stability of Motion. Stanford University Press, Stanford (1963)

    Google Scholar 

  25. Landau, L.D., Lifshitz, E.M.: Mechanics, 3rd edn. Pergamon Press, Oxford (1976)

    Google Scholar 

  26. Lawden, D.F.: Elliptic Functions and Applications. Springer, New York (1989)

    Book  MATH  Google Scholar 

  27. Magnus, W., Winkler, S.: Hill’s Equation, 2nd edn. Dover Publications, New York (2004)

    MATH  Google Scholar 

  28. Murray, C.D., Dermott, S.F.: Solar System Dynamics. Cambridge University Press, Cambridge (2001)

    Google Scholar 

  29. Myers, R.H., Myers, S.L., Walpole, R.E.: Probability and Statistics for Engineers and Scientists, 6th edn. Prentice Hall, London (1998)

    Google Scholar 

  30. Nagle, R.K., Saff, E.B., Snider, A.D.: Fundamentals of Differential Equations, Eight edn. Addison-Wesley, Boston (2012)

    Google Scholar 

  31. Palis, J.: A global view of dynamics and a conjecture on the denseness of finitude of attractors. Astérisque 261, 339–351 (2000)

    Google Scholar 

  32. Percival, I., Richards, D.: Introduction to Dynamics, 1st edn. Cambridge University Press, Cambridge (1994)

  33. Rodrigues, Ch.S., de Moura, A.P.S., Grebogi, C.: Emerging attractors and the transition from dissipative to conservative dynamics. Phys. Rev. E 80 (2009), no. 2, 026205, p. 4.

  34. Stewart, I., Tall, D.: Complex Analysis. Cambridge University Press, Cambridge (1983)

  35. Watson, G.N., Whittaker, E.T.: A Course of Modern Analysis, 4th edn. Cambridge University Press, Cambridge (1947)

    Google Scholar 

  36. Wright, J.A., Deane, J.H.B., Bartuccelli, M., Gentile, G.: Analytic continuation applied to the problem of the pendulum with vertically oscillating support. (in preparation)

Download references

Acknowledgments

The Adams–Bashforth–Moulton method used was MATLAB’s ODE113. We thank Jonathan Deane for helpful conversations on analytic continuation and support with coding in C. This research was completed as part of an EPSRC-funded PhD.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michele Bartuccelli.

Appendices

Appendix 1: Global attraction to the two fixed points

To compute the conditions for attraction to the origin, we use the method outlined in [4]; see also [2]. We define \(f(\tau )\) as in (1.3) and require \(f(\tau ) > 0\): the consequences of this restriction are that the method can only be applied to the downward pointing pendulum when \(\alpha > \beta \). Then, we apply the Liouville transformation

$$\begin{aligned} \tilde{\tau } = \int \limits _0^\tau \sqrt{f(s)}\mathrm{d}s \end{aligned}$$
(7.1)

and write our Eq. (1.3) in terms of the new time \(\tilde{\tau }\) as

$$\begin{aligned} \theta _{\tilde{\tau }\tilde{\tau }} + \left( \frac{\tilde{f}(\tilde{\tau })_{\tilde{\tau }}}{2\tilde{f}(\tilde{\tau })}+ \frac{\gamma }{\sqrt{\tilde{f}(\tilde{\tau })}}\right) \theta _{\tilde{\tau }} + \sin \theta = 0, \end{aligned}$$
(7.2)

where the subscript \(\tilde{\tau }\) represents derivative with respect to the new time \(\tilde{\tau }\) and \(\tilde{f}(\tilde{\tau }) := f(\tau )\). This can be represented as the two-dimensional system on \({\mathbb {T}}\times {\mathbb {R}}\), by setting \(x(\tilde{\tau })=\theta (\tilde{\tau })\) and writing

$$\begin{aligned} x_{\tilde{\tau }} \!=\! y, \qquad y_{\tilde{\tau }} \!=\! \!-\!\frac{y}{\sqrt{\tilde{f}}} \left( \frac{\tilde{f}_{\tilde{\tau }}}{2\sqrt{\tilde{f}}} \!+\! \gamma \right) \!-\! \sin {x},\qquad \end{aligned}$$
(7.3)

for which we have the energy \(E(x,y) = 1 -\cos {x} + y^2/2\). By setting \(H(\tilde{\tau })=E(x(\tilde{\tau }),y(\tilde{\tau }))\), one finds

$$\begin{aligned} H_{\tilde{\tau }} = -\frac{y^2}{\sqrt{\tilde{f}}}\left( \frac{\tilde{f}_{\tilde{\tau }}}{2\sqrt{\tilde{f}}} + \gamma \right) , \end{aligned}$$
(7.4)

thus \(H_{\tilde{\tau }} \le 0\), i.e. \(x,\,y\) are bounded given that \(\gamma \) satisfies

$$\begin{aligned} \gamma > - \min _{\tilde{\tau } \ge 0} \frac{\tilde{f}_{\tilde{\tau }}}{2\sqrt{\tilde{f}}} = - \min _{\tau \ge 0} \frac{f'}{2f}. \end{aligned}$$
(7.5)

Moreover, we have that for all \(\tilde{\tau } > 0\)

$$\begin{aligned} H(\tilde{\tau }) + \int \limits _0^{\tilde{\tau }} \frac{y^2}{\sqrt{\tilde{f}}} \left( \frac{\tilde{f}'}{2\sqrt{\tilde{f}}} + \gamma \right) \mathrm{d}s = H(0), \end{aligned}$$
(7.6)

so that, as \(\tilde{\tau } \rightarrow \infty \), using the properties above we can arrive at

$$\begin{aligned} \min _{s \ge 0} \left[ \frac{1}{\sqrt{\tilde{f}}}\left( \frac{\tilde{f}_{\tilde{\tau }}}{2\sqrt{\tilde{f}}} + \gamma \right) \right] \int \limits _0^\infty y^2(s) \mathrm{d}s < \infty . \end{aligned}$$
(7.7)

Hence, \(y \rightarrow 0\) as time tends to infinity. There are two regions of phase space to consider. Any level curve of \(H\) strictly inside the separatrix of the unperturbed pendulum is the boundary of a positively invariant set \(D\) containing the origin: since \(S = \{(x(\tilde{\tau }),y(\tilde{\tau })) : H_{\tilde{\tau }} = 0\} \cap D\) consists purely of the origin, we can apply the local Barbashin–Krasovsky–La Salle theorem [24] to conclude that every trajectory that begins strictly inside the separatrix will converge to the origin as \(\tilde{\tau }\rightarrow +\infty \).

Outside of the separatrix, we may use Eq. (7.4), which shows the energy to be strictly decreasing while \(y \ne 0\), provided \(\gamma \) is chosen large enough, coupled with \(y \rightarrow 0\) as time tends to infinity. The result is that all trajectories tend to the invariant points on the \(x\)-axis as time tends to infinity. One of two cases must occur: either the trajectory moves inside the separatrix or it does not. In the first instance, we have already shown that the limiting solution is the origin. In the latter, there is only one possibility. As all points on the \(x\)-axis are contained within the separatrix other than the unstable fixed point, the trajectory must move onto such a fixed point and hence belongs to its stable manifold, which is a zero-measure set. Therefore, we conclude that a full-measure set of initial conditions are attracted by the origin. Reverting back to the original system with time \(\tau \), we conclude that for that system too the basin of attraction of the origin has full measure, provided \(\beta <\alpha \) and \(\gamma \) satisfies (7.5).

Appendix 2: Action-angle variables

In this section, we detail the calculation of the action-angle variables for the simple pendulum in time \(\tau \). More details on calculating action-angle variables can be found in [9, 14, 32]. The simple pendulum has equation of motion given by

$$\begin{aligned} \theta '' + \alpha \sin {\theta } = 0, \end{aligned}$$
(8.1)

where the dashes represent derivative with respect to the scaled time \(\tau \). The Hamiltonian for the simple pendulum in this notation is

$$\begin{aligned} E = H(\theta ,\theta ') = \frac{1}{2} (\theta ')^2 - \alpha \cos {\theta } \end{aligned}$$
(8.2)

or, in terms of the usual notation for Hamiltonian dynamics,

$$\begin{aligned} E = H(p,q) = \frac{1}{2} p^2 - \alpha \cos {q}, \end{aligned}$$
(8.3)

where \(q = \theta \) and \(p= q'=\theta '\). Rearranging this for \(p\) we obtain \(p=\pm p(E,q)\), with

$$\begin{aligned} p(E,q) \!=\! \sqrt{2(E \!+\! \alpha \cos {q})} \!=\! \sqrt{2\alpha (E_{0} \!+\! \cos {q})},\nonumber \\ \end{aligned}$$
(8.4)

where \(E_{0} = E/\alpha \). It is clear that there are two types of dynamics, oscillatory dynamics when \(E_{0} < 1\) and rotational dynamics when \(E_{0} > 1\), separated at a separatrix when \(E_{0} = 1\), for which no action-angle variables exist.

1.1 Librations

We first consider the case \(E_{0} < 1\). The action variable is

$$\begin{aligned} I&= \frac{1}{2\pi } \oint p \mathrm{d}q = \frac{2}{\pi }\sqrt{2\alpha } \int \limits _{0}^{q_1} \sqrt{E_{0} + \cos {q}} \; \mathrm{d}q\nonumber \\&= \frac{8}{\pi }\sqrt{\alpha }\Bigl [(k_1^2 - 1)\mathbf{K}(k_1) + \mathbf{E}(k_1)\Bigr ], \end{aligned}$$
(8.5)

where \(k_1^2 = (E_{0}+1)/2\) and \(q_{1}=\arccos (-E_{0})\). The functions \(\mathbf{K}(k)\) and \(\mathbf{E}(k)\) are the complete elliptic integrals of the first and second kinds, respectively.

The angle variable \(\varphi \) can be found as follows

$$\begin{aligned} \varphi ' = \frac{\partial H}{\partial I } = \frac{\mathrm{d}E}{\mathrm{d}I} = \left( \frac{\mathrm{d}I}{\mathrm{d}E}\right) ^{-1}, \end{aligned}$$
(8.6)

so that

$$\begin{aligned} \frac{\mathrm{d}I}{\mathrm{d}E}&= \frac{\mathrm{d}}{\mathrm{d}E} \frac{2}{\pi } \int \limits _{0}^{q_1} \sqrt{E + \alpha \cos {q}} \; \mathrm{d}q \nonumber \\&= \frac{2}{\pi \sqrt{\alpha }} \mathbf{K}(k_1). \end{aligned}$$
(8.7)

Hence, we have

$$\begin{aligned} \varphi (\tau ) = \frac{\pi }{2\mathbf{K}(k_1)} \sqrt{\alpha }(\tau - \tau _0). \end{aligned}$$
(8.8)

Take \(s = \sin {(q/2)}\); then, using Eq. (8.2), it is easy to show that

$$\begin{aligned} (s')^2 = \frac{g}{l}(1-s^2)\left( k_1^2 - s^2\right) . \end{aligned}$$
(8.9)

Integrating using the Jacobi elliptic functions

$$\begin{aligned} s(\tau ) = k_1 {{\mathrm{sn}}}{\left( \sqrt{\frac{g}{l}}(\tau - \tau _0),k_1\right) }, \end{aligned}$$
(8.10)

the expression can then be rearranged to achieve the following result:

$$\begin{aligned} q&= 2\arcsin {\left[ k_1 {{\mathrm{sn}}}{\left( \frac{2\mathbf{K}(k_1)}{\pi }\varphi ,k_1\right) }\right] },\nonumber \\ p&= 2 k_1 \sqrt{\alpha } {{\mathrm{cn}}}{\left( \frac{2\mathbf{K}(k_1)}{\pi }\varphi ,k_1\right) }, \end{aligned}$$
(8.11)

which coincide with Eqs. (2.4). Using (11.3) in Appendix 3, one obtains from (8.5)

$$\begin{aligned} \frac{\partial I}{\partial k_{1}} = \frac{8}{\pi } k_{1} \mathbf{K}(k_{1}) \, \sqrt{\alpha } , \end{aligned}$$
(8.12)

a relation which has been used to derive (2.12).

1.2 Rotations

In the case of rotational dynamics, we have

$$\begin{aligned}&I = \frac{1}{2\pi } \int \limits _{0}^{2\pi } p \, q = \frac{1}{2\pi }\sqrt{\alpha } \int \limits _{0}^{2\pi } \sqrt{E_{0} + \cos {q}} \ \mathrm{d}q\nonumber \\&\quad = \frac{4}{k_2\pi }\sqrt{\alpha } \, \mathbf{E}(k_2), \end{aligned}$$
(8.13)

where this time we let \(k_2^2 = 2/(E_{0} + 1) = 1/k_1^2\). The angle variable \(\varphi \) can similarly be found using (8.6), where \(\mathrm{d}I/\mathrm{d}E\) can be similarly calculated as

$$\begin{aligned} \frac{\mathrm{d}I}{\mathrm{d}E}&= \frac{\mathrm{d}}{\mathrm{d}E} \frac{1}{2\pi } \int \limits _0^{2\pi } \sqrt{E +\alpha \cos {q}} \;\mathrm{d}q \nonumber \\&= \frac{k_2}{\pi \sqrt{\alpha }} \mathbf{K}(k_2), \end{aligned}$$
(8.14)

which hence gives

$$\begin{aligned} \varphi (\tau ) = \frac{\pi }{\mathbf{K}(k_2)}\sqrt{\alpha }\frac{(\tau - \tau _0)}{k_2}. \end{aligned}$$
(8.15)

Using (8.9) and the definition of \(k_2\), for the rotating solutions we find that

$$\begin{aligned} s(\tau ) = {{\mathrm{sn}}}{\left( \sqrt{\alpha } \frac{(\tau - \tau _0)}{k_2},k_2\right) } , \end{aligned}$$
(8.16)

and similarly, by simple rearrangement, we find that

$$\begin{aligned} q&= 2 \arcsin {\left[ {{\mathrm{sn}}}{\left( \frac{\mathbf{K}(k_2)}{\pi }\varphi ,k_2\right) }\right] },\nonumber \\ p&= \frac{2}{k_2} \sqrt{\alpha } \, {{\mathrm{dn}}}{\left( \frac{\mathbf{K}(k_2)}{\pi }\varphi ,k_2\right) }, \end{aligned}$$
(8.17)

which again yields Eqs. (2.6). Using (11.3) in Appendix 3, one obtains from (8.13)

$$\begin{aligned} \frac{\partial I}{\partial k_{2}} = - \frac{4}{\pi k_{2}^{2}} \mathbf{K}(k_{2}) \, \sqrt{\alpha } \end{aligned}$$
(8.18)

which has been used to derive (2.23).

Appendix 3: Jacobian determinant

Here, we compute the entries of the Jacobian matrix \(J\) of the transformation to action-angle variables, which will be used in the next Appendix. As a by-product, we check that \(J\) determinant equal to 1, that is

$$\begin{aligned} \frac{\partial q}{\partial \varphi }\frac{\partial p}{\partial I} - \frac{\partial q}{\partial I}\frac{\partial p}{\partial \varphi } =1. \end{aligned}$$
(9.1)

For further details on the proof of (9.1), we refer the reader to [9], where the calculations are given in great detail. The derivative with respect to \(\varphi \) is straightforward in both the libration and rotation case, however, the dependence of \(p\) and \(q\) on the action \(I\) is less obvious. That said, the dependence of \(p\) and \(q\) on \(k_1\) in the oscillating case and \(k_2\) in the rotating case is clear and we know the relationship between \(I\) and \(k\) in both cases; hence, the derivative of the Jacobi elliptic functions can be calculated using that

$$\begin{aligned}&\frac{\partial }{\partial I} \!= \!\frac{\partial k}{\partial I}\frac{\partial }{\partial k} \!+\! \frac{\partial u}{\partial I}\frac{\partial }{\partial u} \!=\! \frac{\partial k}{\partial I}\left( \frac{\partial }{\partial k} \!+\! \frac{\partial u}{\partial k}\frac{\partial }{\partial u}\right) ,\qquad \end{aligned}$$
(9.2)

where \(u\) is the first argument of the functions, i.e. \({{\mathrm{sn}}}(u,k)\), etc. Then, for the oscillations, we have

$$\begin{aligned} \frac{\partial q}{\partial I}&= \frac{\pi }{4k_1\mathbf{K}(k_1)\sqrt{\alpha }}\left[ \frac{{{\mathrm{sn}}}(\cdot )}{{{\mathrm{dn}}}(\cdot )}+ \frac{2\mathbf{E}(k_1)\varphi {{\mathrm{cn}}}(\cdot )}{\pi k_{1}'^2}\right. \nonumber \\&\left. + \frac{k_1^2{{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}^2(\cdot )}{k_1'^2{{\mathrm{dn}}}(\cdot )} - \frac{\mathbf{E}(\cdot ){{\mathrm{cn}}}(\cdot )}{k_1'^2} \right] , \nonumber \\ \frac{\partial p}{\partial I}&= \frac{\pi }{4k_1\mathbf{K}(k_1)}\left[ {{\mathrm{cn}}}(\cdot ) - \frac{2\mathbf{E}(k_1)\varphi {{\mathrm{sn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{\pi k_{1}'^2}\right. \nonumber \\&\left. - \frac{k_1^2{{\mathrm{sn}}}^2(\cdot ){{\mathrm{cn}}}(\cdot )}{k_1'^2} + \frac{\mathbf{E}(\cdot ){{\mathrm{sn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{k_1'^2}\right] ,\nonumber \\ \frac{\partial q}{\partial \varphi }&= \frac{4k_1\mathbf{K}(k_1){{\mathrm{cn}}}(\cdot )}{\pi } ,\nonumber \\ \frac{\partial p}{\partial \varphi }&= -\sqrt{\alpha } \, \frac{4k_1\mathbf{K}(k_1){{\mathrm{sn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{\pi }, \end{aligned}$$
(9.3)

where \((\cdot ) = \left( \frac{2\mathbf{K}(k_1)\varphi }{\pi },k_1\right) \) and \(k_1' = \sqrt{1 - k_1^2}\). From the above, it is easy to check that Eq. (9.1) is satisfied. Similarly, for the rotations, we have

$$\begin{aligned} \frac{\partial q}{\partial I}&= - \frac{\pi k_2^2}{2 \mathbf{K}(k_2) \, \sqrt{\alpha }} \left[ \frac{\varphi \, \mathbf{E}(k_{2})\,{{\mathrm{dn}}}(\cdot )}{\pi k_{2}k_{2}'^2}\right. \nonumber \\&\left. + \frac{k_2{{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot )}{k_2'^2} - \frac{\mathbf{E}(\cdot ){{\mathrm{dn}}}(\cdot )}{k_2k_2'^2}\right] , \nonumber \\ \frac{\partial p}{\partial I}&= \frac{\pi k_2^2}{2 \mathbf{K}(k_2)} \left[ \frac{{{\mathrm{dn}}}(\cdot )}{k_2^2} + \frac{\varphi \, \mathbf{E}(k_{2})\,{{\mathrm{sn}}}(\cdot ) \, {{\mathrm{cn}}}(\cdot )}{\pi k_{2}'^2}\right. \nonumber \\&\left. + \frac{{{\mathrm{sn}}}^2(\cdot ){{\mathrm{dn}}}(\cdot )}{k_2'^2} - \frac{\mathbf{E}(\cdot ){{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot )}{k_2'^2} \right] , \nonumber \\ \frac{\partial q}{\partial \varphi }&= \frac{2\mathbf{K}(k_2){{\mathrm{dn}}}(\cdot )}{\pi },\nonumber \\ \frac{\partial p}{\partial \varphi }&= -\sqrt{\alpha } \, \frac{2k_2\mathbf{K}(k_2){{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot )}{\pi } , \end{aligned}$$
(9.4)

where \((\cdot ) = \left( \frac{\mathbf{K}(k_2)}{\pi }\varphi ,k_2\right) \) and \(k_2' = \sqrt{1 - k_2^2}\). It is once again easily checked from the above that (9.1) is satisfied.

Appendix 4: Equations of motion for the perturbed system

By (9.1), one has

$$\begin{aligned} \begin{pmatrix} \partial \varphi /\partial q &{} \partial \varphi / \partial p \\ \partial I /\partial q &{} \partial I / \partial p \end{pmatrix} = \begin{pmatrix} \partial p / \partial I &{} - \partial q / \partial I \\ - \partial p/ \partial \varphi &{} \partial q / \partial \varphi . \end{pmatrix} \end{aligned}$$
(10.1)

We rewrite the Eq. (1.3) in the action-angle coordinates introduced in Appendix 2 as follows.

1.1 Librations

In this section, we want to write (1.3) in terms of the action-angle introduced in Appendix 2. By taking into account the forcing term \(-\beta \cos \tau \cos \theta \) in (1.3), one finds

$$\begin{aligned}&I' \!=\! \frac{\partial I}{\partial q}q' \!+\! \frac{\partial I}{\partial p}p' \!=\! - \frac{\partial p}{\partial \varphi }q' \!+\! \frac{\partial q}{\partial \varphi }p' \nonumber \\&\quad = \frac{8 \beta k_1^2 \mathbf{K}(k_1)}{\pi }\cos (\tau \!-\! \tau _0){{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot ), \nonumber \\&\varphi ' \!=\! \frac{\partial \varphi }{\partial q}q' \!+\! \frac{\partial \varphi }{\partial p}p' \!=\! \frac{\partial p}{\partial I}q' \!-\! \frac{\partial q}{\partial I}p' \nonumber \\&\quad = \frac{\pi \sqrt{\alpha }}{2\mathbf{K}(k_1)} \!-\! \frac{\pi \beta }{2\mathbf{K}(k_1)\,\sqrt{\alpha }}\nonumber \\&\left[ {{\mathrm{sn}}}^2(\cdot ) \!+\! \frac{2\mathbf{E}(k_1)\varphi {{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{\pi k_{1}'^2}+ \left. \frac{k_1^2{{\mathrm{sn}}}^2(\cdot ){{\mathrm{cn}}}^2(\cdot )}{1\!-\!k_1^2} \right. \right. \nonumber \\&\quad \left. - \frac{\mathbf{E}(\cdot ) {{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{1 \!-\! k_1^2}\right] \cos (\tau \!-\! \tau _0) , \end{aligned}$$
(10.2)

where we have used the properties of the Jacobi elliptic functions in Appendix 5. As in Appendix 3, we are shortening \((\cdot ) = \left( \frac{2\mathbf{K}(k_1)\varphi }{\pi },k_1\right) \).

We then wish to add the dissipative term \(\gamma \theta '\). This results in the following equations:

$$\begin{aligned} I'&= \frac{8 \beta k_1^2 \mathbf{K}(k_1)}{\pi } \cos (\tau - \tau _0){{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )\nonumber \\&-\frac{8 \gamma k_1^2 \sqrt{\alpha } \, \mathbf{K}(k_1)}{\pi }{{\mathrm{cn}}}^2(\cdot ), \nonumber \\ \varphi '&= \frac{\pi \sqrt{\alpha }}{2\mathbf{K}(k_1)}- \frac{\pi \beta }{2\mathbf{K}(k_1)\,\sqrt{\alpha }}\left[ {{\mathrm{sn}}}^2(\cdot )\right. \nonumber \\&\left. +\, \frac{2\mathbf{E}(k_1)\varphi {{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{\pi (1-k_{1}^2)}+ \frac{k_1^2{{\mathrm{sn}}}^2(\cdot ){{\mathrm{cn}}}^2(\cdot )}{1-k_1^2}\right. \nonumber \\&\left. -\, \frac{\mathbf{E}(\cdot ) {{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{1 - k_1^2}\right] \cos (\tau - \tau _0) \nonumber \\&+\, \frac{\gamma \pi {{\mathrm{cn}}}(\cdot )}{2\mathbf{K}(k_1)}\left[ \frac{{{\mathrm{sn}}}(\cdot )}{{{\mathrm{dn}}}(\cdot )} \right. \left. + \frac{2 \mathbf{E}(k_1)\varphi {{\mathrm{cn}}}(\cdot )}{\pi (1-k_{1}^2)}\right. \nonumber \\&\left. +\, \frac{k_1^2{{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}^2(\cdot )}{(1-k_1^2){{\mathrm{dn}}}(\cdot )} - \frac{ \mathbf{E}(\cdot ){{\mathrm{cn}}}(\cdot )}{1 - k_1^2}\right] . \end{aligned}$$
(10.3)

Using the property that, see [26], \(\mathbf{E}(u,k) \!=\! \mathbf{E}(k)u/\mathbf{K}(k)\) \( + \mathbf{Z}(u,k)\), we arrive at equations (2.9). The function \(\mathbf{Z}(u,k)\) is the Jacobi zeta function, which is periodic with period \(2\mathbf{K}(k)\) in \(u\).

1.2 Rotations

The presence of the forcing term leads to the Equations.

$$\begin{aligned} I'&= - \frac{\partial p}{\partial \varphi }q' + \frac{\partial q}{\partial \varphi }p'\nonumber \\&= \frac{4 \beta \mathbf{K}(k_2)}{\pi }\cos (\tau - \tau _0) {{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot ), \nonumber \\ \varphi '&= \frac{\partial p}{\partial I}\dot{q} - \frac{\partial q}{\partial I}\dot{p}= \frac{\pi \sqrt{\alpha }}{k_2\mathbf{K}(k_2)}+\, \frac{\pi k_2 \beta }{\sqrt{\alpha }\mathbf{K}(k_2)}\nonumber \\&\times \left[ \frac{ \mathbf{E}(k_{2}) \, \varphi \,{{\mathrm{sn}}}(\cdot )\,{{\mathrm{cn}}}(\cdot )\,{{\mathrm{dn}}}(\cdot )}{\pi (1-k_{2}^2)} \right. +\,\frac{k^2_2 {{\mathrm{sn}}}^2(\cdot ){{\mathrm{cn}}}^2(\cdot )}{1 - k_2^2}\nonumber \\&\left. -\frac{\mathbf{E}(\cdot ){{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{1 - k_2^2} \right] \cos (\tau - \tau _0). \end{aligned}$$
(10.4)

Again, if we wish to add a dissipative term, we arrive at the equations

$$\begin{aligned} I'&= \frac{4 \beta \mathbf{K}(k_2)}{\pi }\cos (\tau \!-\! \tau _0) {{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )\nonumber \\&-\frac{4\gamma \sqrt{\alpha } \, \mathbf{K}(k_2)}{\pi k_2}{{\mathrm{dn}}}^2(\cdot ), \nonumber \\ \varphi '&= \frac{\pi \sqrt{\alpha }}{k_2\mathbf{K}(k_2)} \!+\! \frac{\pi k_2 \beta }{\sqrt{\alpha }\mathbf{K}(k_2)} \left[ \frac{ \mathbf{E}(k_{2}) \, \varphi \,{{\mathrm{sn}}}(\cdot )\,{{\mathrm{cn}}}(\cdot )\,{{\mathrm{dn}}}(\cdot )}{\pi (1\!-\!k_{2}^2)} \right. \nonumber \\&\left. + \frac{k^2_2 {{\mathrm{sn}}}^2(\cdot ){{\mathrm{cn}}}^2(\cdot )}{1 \!-\! k_2^2}\!-\! \frac{\mathbf{E}(\cdot ){{\mathrm{sn}}}(\cdot ){{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{1 \!-\! k_2^2} \right] \nonumber \\&\times \cos (\tau \!-\! \tau _0) \!-\! \frac{\gamma \pi }{\mathbf{K}(k_2)} \left[ \frac{ \mathbf{E}(k_{2}) \, \varphi \,{{\mathrm{dn}}}^{2} (\cdot )}{\pi (1\!-\!k_{2}^2)}\right. \nonumber \\&\left. + \frac{k_2^2{{\mathrm{sn}}}(\cdot ) {{\mathrm{cn}}}(\cdot ){{\mathrm{dn}}}(\cdot )}{1 \!-\! k_2^2} \!-\! \frac{ \mathbf{E}(\cdot ){{\mathrm{dn}}}^2(\cdot )}{1\!-\!k_2^2} \right] . \end{aligned}$$
(10.5)

Again, using that \(\mathbf{E}(u,k) = \mathbf{E}(k)u/\mathbf{K}(k) + \mathbf{Z}(u,k)\), we arrive at the equations (2.22).

Appendix 5: Useful properties of the elliptic functions

The complete integrals of the first and second kind are, respectively,

$$\begin{aligned} \mathbf{K}(k)&= \int \limits _{0}^{\pi /2} \frac{\mathrm{d}\psi }{\sqrt{1-k^2 \sin ^2 \psi }} ,\nonumber \\ \mathbf{E}(k)&= \int \limits _{0}^{\pi /2} \mathrm{d}\psi \, \sqrt{1-k^2 \sin ^2 \psi } , \end{aligned}$$
(11.1)

whereas the incomplete elliptic integral of the second kind is

$$\begin{aligned} \mathbf{E}(u,k) = \int \limits _{0}^{{{\mathrm{sn}}}(u,k)} \mathrm{d}x \frac{\sqrt{1-k^2 x^{2}}}{\sqrt{1-x^2}}. \end{aligned}$$
(11.2)

One has

$$\begin{aligned}&\!\!\!\frac{\partial \mathbf{K}(k)}{\partial k} = \frac{1}{k} \left( \frac{\mathbf{E}(k)}{1-k^2} - \mathbf{K}(k) \right) ,\nonumber \\&\!\!\!\frac{\partial \mathbf{E}(k)}{\partial k} = \frac{1}{k} \left( \mathbf{E}(k) - \mathbf{K}(k) \right) . \end{aligned}$$
(11.3)

The following properties of the Jacobi elliptic functions have been used in the previous sections. The derivatives with respect to the first arguments are

$$\begin{aligned} \frac{\partial }{\partial u} {{\mathrm{sn}}}(u,k)&= {{\mathrm{cn}}}(u,k) \, {{\mathrm{dn}}}(u,k) , \nonumber \\ \frac{\partial }{\partial u} {{\mathrm{cn}}}(u,k)&= - {{\mathrm{sn}}}(u,k) \, {{\mathrm{dn}}}(u,k) , \\ \frac{\partial }{\partial u} {{\mathrm{dn}}}(u,k)&= - k^2 {{\mathrm{sn}}}(u,k) \, {{\mathrm{cn}}}(u,k) ,\nonumber \end{aligned}$$
(11.4)

while the derivatives with respect to the elliptic modulus are

$$\begin{aligned} \frac{\partial }{\partial k} {{\mathrm{sn}}}(u,k)&= \frac{u}{k} {{\mathrm{cn}}}(u,k) \, {{\mathrm{dn}}}(u,k)\nonumber \\&+\, \frac{k}{k'^2} \, {{\mathrm{sn}}}(u,k) \, {{\mathrm{cn}}}^{2} (u,k)\nonumber \\&-\,\frac{1}{k k'^2} \mathbf{E}(u,k)\,{{\mathrm{cn}}}(u,k) \, {{\mathrm{dn}}}(u,k) ,\nonumber \\ \frac{\partial }{\partial k} {{\mathrm{cn}}}(u,k)&= - \frac{u}{k} {{\mathrm{sn}}}(u,k) \, {{\mathrm{dn}}}(u,k)\nonumber \\&-\, \frac{k}{k'^2} \, {{\mathrm{sn}}}^2(u,k) \, {{\mathrm{cn}}}(u,k)\nonumber \\&+\,\frac{1}{k k'^2} \mathbf{E}(u,k)\,{{\mathrm{sn}}}(u,k) \, {{\mathrm{dn}}}(u,k) ,\nonumber \\ \frac{\partial }{\partial k} {{\mathrm{dn}}}(u,k)&= -ku \, {{\mathrm{sn}}}(u,k) \, {{\mathrm{cn}}}(u,k)\nonumber \\&-\, \frac{k}{k'^2} \, {{\mathrm{sn}}}^{2}(u,k) \, {{\mathrm{dn}}}(u,k)\nonumber \\&+\,\frac{k}{k'^2} \mathbf{E}(u,k)\,{{\mathrm{sn}}}(u,k) \, {{\mathrm{cn}}}(u,k),\nonumber \\ \end{aligned}$$
(11.5)

where \(k'^2=1-k^2\).

Finding the value of \(\Delta \) for rotations in Sect. 2 requires use of

$$\begin{aligned} \int \limits _0^{x_1} {{\mathrm{dn}}}^2(x,k) \mathrm{d}x \!=\! \int \limits _0^{{{\mathrm{sn}}}(x_1,k)} \frac{\sqrt{1 \!-\! k^2\hat{x}^2}}{\sqrt{1 \!-\! \hat{x}^2}} \mathrm{d}\hat{x} \!=\! \mathbf{E}(x_1,k).\nonumber \\ \end{aligned}$$
(11.6)

In the case of librations, we also require the relation \(k^2 {{\mathrm{cn}}}^2(\cdot ) + (1 - k^2) = {{\mathrm{dn}}}^2(\cdot )\).

The integral for in Eq. (2.20) is found by

(11.7)

where .

The Jacobi elliptic functions can be expanded in a Fourier series as

$$\begin{aligned}&{{\mathrm{sn}}}(u,k) \!=\! \frac{2\pi }{k\mathbf{K}(k)}\sum _{n=1}^\infty \frac{{\mathfrak {q}}^{n - 1/2}}{1 \!-\! {\mathfrak {q}}^{2n - 1}}\sin \left( \frac{(2n \!-\! 1)\pi u}{2\mathbf{K}(k)}\right) ,\nonumber \\&{{\mathrm{cn}}}(u,k)\!=\! \frac{2\pi }{k\mathbf{K}(k)}\sum _{n=1}^\infty \frac{{\mathfrak {q}}^{n - 1/2}}{1 \!+\! {\mathfrak {q}}^{2n - 1}}\cos \left( \frac{(2n \!-\! 1)\pi u}{2\mathbf{K}(k)}\right) ,\nonumber \\&{{\mathrm{dn}}}(u,k) \!=\! \frac{\pi }{2\mathbf{K}(k)}\!+\!\frac{2\pi }{\mathbf{K}(k)}\sum _{n=1}^\infty \frac{{\mathfrak {q}}^{n}}{1 \!-\! {\mathfrak {q}}^{2n}}\cos \left( \frac{2n\pi u}{2\mathbf{K}(k)}\right) \!, \end{aligned}$$
(11.8)

where \({\mathfrak {q}}\) is the nome, defined as

$$\begin{aligned} {\mathfrak {q}} = \exp \left( - \frac{ \pi \mathbf{K}(k')}{\mathbf{K}(k)}\right) , \end{aligned}$$

with \(k'=\sqrt{1-k^2}\).

In the calculation of \(\langle R^{(n)} \rangle \) for \(n \ge 2\), when the pendulum is in libration, we require the evaluation of the integrals

$$\begin{aligned} \frac{1}{T}\int \limits _0^T \frac{2\mathbf{K}(k_1)}{\sqrt{\alpha }\pi } \frac{\partial }{\partial \tau }\Bigl ({{\mathrm{cn}}}^2(\sqrt{\alpha }\tau )\Bigr ) \mathrm{d}\tau = 0, \end{aligned}$$
(11.9)

and

$$\begin{aligned}&\frac{1}{T}\int \limits _0^T \frac{2\mathbf{K}(k_1)}{\sqrt{\alpha }\pi }\frac{\partial }{\partial \tau } \Bigl ({{\mathrm{sn}}}(\sqrt{\alpha }\tau ){{\mathrm{cn}}}(\sqrt{\alpha }\tau ){{\mathrm{dn}}}(\sqrt{\alpha }\tau ) \Bigr )\cos (\tau - \tau _0) \mathrm{d}\tau \nonumber \\&\quad = \frac{1}{T}\int \limits _0^T \frac{2\mathbf{K}(k_1)}{\sqrt{\alpha }\pi }{{\mathrm{sn}}}(\sqrt{\alpha }\tau ) {{\mathrm{cn}}}(\sqrt{\alpha }\tau ){{\mathrm{dn}}}(\sqrt{\alpha }\tau )\sin (\tau - \tau _0) \mathrm{d}\tau \nonumber \\&\quad = \frac{\cos (\tau _0)}{T}\int \limits _0^T \frac{2\mathbf{K}(k_1)}{\sqrt{\alpha }\pi } {{\mathrm{sn}}}(\sqrt{\alpha }\tau ){{\mathrm{cn}}}(\sqrt{\alpha }\tau ){{\mathrm{dn}}}(\sqrt{\alpha }\tau )\sin (\tau ) \mathrm{d}\tau \nonumber \\&\qquad - \frac{\sin (\tau _0)}{T}\int \limits _0^T \frac{2\mathbf{K}(k_1)}{\sqrt{\alpha }\pi } {{\mathrm{sn}}}(\sqrt{\alpha }\tau ){{\mathrm{cn}}}(\sqrt{\alpha }\tau ){{\mathrm{dn}}}(\sqrt{\alpha }\tau )\cos (\tau ) \mathrm{d}\tau , \nonumber \\ \end{aligned}$$
(11.10)

where . The integral multiplying \(\sin (\tau _0)\) vanishes due to parity, and hence,

(11.11)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wright, J.A., Bartuccelli, M. & Gentile, G. The effects of time-dependent dissipation on the basins of attraction for the pendulum with oscillating support. Nonlinear Dyn 77, 1377–1409 (2014). https://doi.org/10.1007/s11071-014-1386-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-014-1386-1

Keywords

Mathematics Subject Classification

Navigation