Skip to main content
Log in

Nonlinear dynamics of the quasi-zero-stiffness SD oscillator based upon the local and global bifurcation analyses

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

Abstract

In this paper, the complicated nonlinear dynamics of the harmonically forced quasi-zero-stiffness SD (smooth and discontinuous) oscillator is investigated via direct numerical simulations. This oscillator considered that the gravity is composed of a lumped mass connected with a vertical spring of positive stiffness and a pair of horizontally compressed springs providing negative stiffness, which can achieve the quasi-zero stiffness widely used in vibration isolation. The local and global bifurcation analyses are implemented to reveal the complex dynamic phenomena of this system. The double-parameter bifurcation diagrams are constructed to demonstrate the overall topological structures for the distribution of various responses in parameter spaces. Using the Floquet theory and parameter continuation method, the local bifurcation patterns of periodic solutions are obtained. Moreover, the global bifurcation mechanisms for the crises of chaos and metamorphoses of basin boundaries are examined by analysing the attractors and attraction basins, exploring the evolutions of invariant manifolds and constructing the basin cells. Meanwhile, additional nonlinear dynamic phenomena and characteristics closely related to the bifurcations are discussed including the resonant tongues, jump phenomena, amplitude–frequency responses, chaotic seas, transient chaos, chaotic saddles, and also their generation mechanisms are presented.

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

Similar content being viewed by others

References

  1. McFarlanda, D.M., Bergmana, L.A., Vakakis, A.F.: Experimental study of non-linear energy pumping occurring at a single fast frequency. Int. J. Non-Linear Mech. 40(6), 891–899 (2005)

    Article  Google Scholar 

  2. Ibrahim, R.A.: Recent advances in nonlinear passive vibration isolators. J. Sound Vib. 314, 371–452 (2008)

    Article  Google Scholar 

  3. Lee, Y.S., Vakakis, A.F., Bergman, L.A., McFarland, D.M., Kerschen, G., Nucera, F., Tsakirtzis, S., Panagopoulos, P.N.: Passive non-linear targeted energy transfer and its applications to vibration absorption: a review. Inst. Mech. Eng. K J. Multi-Body Dyn. 222, 77–134 (2008)

    Google Scholar 

  4. Liu, C., Jing, X., Daley, S., Li, F.: Recent advances in micro-vibration isolation. Mech. Syst. Signal Process. 56–57, 55–80 (2015)

    Article  Google Scholar 

  5. Carrella, A., Brennan, M.J., Waters, T.P.: Static analysis of a passive vibration isolator with quasi-zero-stiffness characteristic. J. Sound Vib. 301, 678–689 (2007)

    Article  Google Scholar 

  6. Kovacic, I., Brennan, M.J., Waters, T.P.: A study of a nonlinear vibration isolator with a quasi-zero stiffness characteristic. J. Sound Vib. 315(3), 700–711 (2008)

    Article  Google Scholar 

  7. Carrella, A., Brennan, M.J., Waters Jr., T.P., Lopes, V.: Force and displacement transmissibility of a nonlinear isolator with high-static-low-dynamic stiffness. Int. J. Mech. Sci. 55(1), 22–29 (2012)

    Article  Google Scholar 

  8. Hao, Z., Cao, Q.: The isolation characteristics of an archetypal dynamical model with stable-quasi-zero-stiffness. J. Sound Vib. 340, 61–79 (2015)

    Article  Google Scholar 

  9. Zhou, J.X., Wang, X.L., Xu, D.L., Bishop, S.: Nonlinear dynamic characteristics of a quasi-zero stiffness vibration isolator with cam-roller-spring mechanisms. J. Sound Vib. 346, 53–69 (2015)

    Article  Google Scholar 

  10. Xu, J., Sun, X.T.: A multi-directional vibration isolator based on quasi-zero-stiffness structure and time-delayed active control. Int. J. Mech. Sci. 100, 126–135 (2015)

    Article  Google Scholar 

  11. Hao, Z.F., Cao, Q.J., Wiercigroch, M.: Two-sided damping constraint control strategy for high-performance vibration isolation and end-stop impact protection. Nonlinear Dyn. (2016) doi:10.1007/s11071-016-2685-5

  12. Alevras, P., Brown, I., Yurchenko, D.: Experimental investigation of a rotating parametric pendulum. Nonlinear Dyn. 81, 201–213 (2015)

    Article  Google Scholar 

  13. Harne, R.L., Thota, M., Wang, K.W.: Bistable energy harvesting enhancement with an auxiliary linear oscillator. Smart Mater. Struct. 22(12), 125028 (2013)

    Article  Google Scholar 

  14. Jiang, W.A., Chen, L.Q.: Snap-through piezoelectric energy harvesting. J. Sound Vib. 333, 4314–4325 (2014)

    Article  Google Scholar 

  15. Asai, Y., Kimura, K., Asai, T., Masui, T., Omori, T., Kainuma, R.: Integrated mechanical and material design of quasi-zero-stiffness vibration isolator with super elastic Cu-Al-Mn shape memory alloy bars. J. Sound Vib. 358, 74–83 (2015)

    Article  Google Scholar 

  16. Cao, Q.J., Wiercigroch, M., Pavlovskaia, E.E.: An archetypal oscillator for smooth and discontinuous dynamics. Phys. Rev. E 74, 046218 (2006)

    Article  MathSciNet  Google Scholar 

  17. Thompson, J.M.T., Hunt, G.W.: A General Theory of Elastic Stability. Wiley, London (1973)

    MATH  Google Scholar 

  18. Molyneux, W.G.: The support of an aircraft for ground resonance tests: a survey of available methods. Aircr. Eng. Aerosp. Technol. 30, 160–166 (1958)

    Article  Google Scholar 

  19. Cao, Q.J., Wiercigroch, M., Pavlovskaia, E.E., Thompson, J.M.T., Grebogi, C.: Piecewise linear approach to an archetypal oscillator for smooth and discontinuous dynamics. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 366, 635–652 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  20. Cao, Q.J., Wiercigroch, M., Pavlovskaia, E.E.: The limit case response of the archetypal oscillator for smooth and discontinuous dynamics. Int. J. Non-Linear Mech. 43, 462–473 (2008)

    Article  Google Scholar 

  21. Tian, R.L., Cao, Q.J., Yang, S.P.: The codimension-two bifurcation for the recent proposed SD oscillator. Nonlinear Dyn. 59(1–2), 19–27 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  22. Léger, A., Pratt, E., Cao, Q.: A fully nonlinear oscillator with contact and friction. Nonlinear Dyn. 70(1), 511–522 (2012)

    Article  MathSciNet  Google Scholar 

  23. Shen, J., Li, Y., Du, Z.: Subharmonic and grazing bifurcations for a simple bilinear oscillator. Int. J. Non-Linear Mech. 60, 70–82 (2014)

    Article  Google Scholar 

  24. Yue, X., Xu, W., Wang, L.: Stochastic bifurcations in the SD (smooth and discontinuous) oscillator under bounded noise excitation. Sci. China-Technol. 56(5), 1010–1016 (2013)

    Google Scholar 

  25. Zhang, Y., Luo, G., Cao, Q., Lin, M.: Wada basin dynamics of a shallow arch oscillator with more than 20 coexisting low-period periodic attractors. Int. J. Non-Linear Mech. 58, 151–161 (2014)

    Article  Google Scholar 

  26. Zhang, Y., Lu, L.F.: Basin boundaries with nested structure in a shallow arch oscillator. Nonlinear Dyn. 77, 1121–1132 (2014)

    Article  MathSciNet  Google Scholar 

  27. Zhang, Y., Zhang, H.: Metamorphoses of basin boundaries with complex topology in an archetypal oscillator. Nonlinear Dyn. 79(4), 2309–2323 (2015)

    Article  MathSciNet  Google Scholar 

  28. Zhang, Y., Zhang, H., Gao, W.: Multiple Wada basins with common boundaries in nonlinear driven oscillators. Nonlinear Dyn. 79(4), 2667–2674 (2015)

    Article  MathSciNet  Google Scholar 

  29. Cao, Q.J., Wang, D., Chen, Y.S., Wiercigroch, M.: Irrational elliptic functions and the analytical solutions of SD oscillator. J. Theor. Appl. Mech. 50(3), 701–715 (2012)

    Google Scholar 

  30. Santhosh, B., Padmanabhan, C., Narayanan, S.: Numeric-analytic solutions of the smooth and discontinuous oscillator. Int. J. Mech. Sci. 84, 102–119 (2014)

    Article  Google Scholar 

  31. Chen, H.B., Xie, J.H.: Harmonic and subharmonic solutions of the SD oscillator. Nonlinear Dyn. (2016). doi:10.1007/s11071-016-2659-7

  32. Chen, E.L., Cao, Q.J., Feng, M., Tian, R.L.: The preliminary investigation on design and experimental research of the nonlinear characteristics of SD oscillator. Chin. J. Theor. Appl. Mech. 44(3), 584–590 (2012)

    Google Scholar 

  33. Lan, C., Yang, S., Wu, Y.: Design and experiment of a compact quasi-zero-stiffness isolator capable of a wide range of loads. J. Sound Vib. 333(20), 4843–4858 (2014)

    Article  Google Scholar 

  34. Vakakis, A.F., Gendelman, O.V., Bergman, L.A., McFarland, D.M., Kerschen, G., Lee, Y.S.: Nonlinear Targeted Energy Transfer in Mechanical and Structural Systems. Springer, Dordrecht (2008)

    MATH  Google Scholar 

  35. Yang, K., Harne, R.L., Wang, K.W., Huang, H.: Dynamic stabilization of a bistable suspension system attached to a flexible host structure for operational safety enhancement. J. Sound Vib. 333, 6651–6661 (2014)

    Article  Google Scholar 

  36. AL-Shudeifat, M.A.: Highly efficient nonlinear energy sink. Nonlinear Dyn. 76, 1905–1920 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  37. Cohen, N., Bucher, I.: On the dynamics and optimization of a non-smooth bistable oscillator-Application to energy harvesting. J. Sound Vib. 333, 4653–4667 (2014)

    Article  Google Scholar 

  38. Ueda, Y.: Survey of regular and chaotic phenomena in the forced Duffing Oscillator. Chaos Solitons Fractals 1(3), 199–231 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  39. Massera, J.L.: The number of subharmonic solutions of non-linear differential equations of the second order. Ann. Math. 50(1), 118–126 (1949)

    Article  MathSciNet  MATH  Google Scholar 

  40. Levinson, N.: Transformation theory of non-Linear differential equations of the second order. Ann. Math. 45(4), 723–737 (1944)

    Article  MathSciNet  MATH  Google Scholar 

  41. Stewart, H.B., Thompson, J.M.T., Ueda, Y., Lansbury, A.N.: Optimal escape from potential wells: patterns of regular and chaotic bifurcation. Phys. D 85, 259–295 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  42. Hao, Z., Cao, Q.: A novel dynamical model for GVT nonlinear supporting system with stable-quasi-zero-stiffness. J. Theor. Appl. Mech. 52(1), 199–213 (2014)

    Google Scholar 

  43. Xu, X., Wiercigroch, M., Cartmell, M.P.: Rotating orbits of a parametrically-excited pendulum. Chaos Solitons Fractals 23, 1537–1548 (2005)

    Article  MATH  Google Scholar 

  44. Ueda, Y., Yoshida, S., Stewart, H.B., Thompson, J.M.T.: Basin explosions and escape phenomena in the twin-well Duffing oscillator: compound global bifurcations organizing behaviour. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 332, 169–186 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  45. Ueda, Y.: Random phenomena resulting from nonlinearity in the system described by Duffing’s equation. Int. J. Non-Linear Mech. 20(5/6), 481–491 (1985)

    Article  Google Scholar 

  46. Doedel, E.J., Champneys, A.R., Decola, F., Fairgrieve, T., Kuznetsov, Y., Paffenroth, R., Sandstede, B., Wang, X.J., Zhang, C.H.: Auto-07P: Continuation and Bifurcation Software for Ordinary Differential Equations. Computer Science, Concordia University, Montreal (2011)

    Google Scholar 

  47. Chávez, J.P., Pavlovskaia, E., Wiercigroch, M.: Bifurcation analysis of a piecewise-linear impact oscillator with drift. Nonlinear Dyn. 77, 213–227 (2014)

    Article  MathSciNet  Google Scholar 

  48. Nayfeh, A.H., Balachandran, B.: Applied Nonlinear Dynamics: Analytical, Computational, and Experimental Methods. Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim (2004)

    MATH  Google Scholar 

  49. Nusse, H.E., Yorke, J.A.: Dynamics: Numerical Explorations. Applied Mathematical Sciences, vol. 101. Springer, New York (1998)

    Book  Google Scholar 

  50. Kuznetsov, Y.A.: Elements of Applied Bifurcation Theory. Springer, New York (1998)

    MATH  Google Scholar 

  51. Nayfeh, A.H.: Introduction to Perturbation Techniques. Wiley, New York (1981)

    MATH  Google Scholar 

  52. Sander, E., Yorke, J.A.: A period-doubling cascade precedes chaos for planar maps. Chaos 23, 033113 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  53. Stewart, H.B., Ueda, Y., Grebogi, C., Yorke, J.A.: Double crises in two-parameter dynamical systems. Phys. Rev. Lett. 75(13), 2478–2481 (1995)

    Article  Google Scholar 

  54. Grebogi, C., Ott, E.: Crises, sudden changes in chaotic attractors, and transient chaos. Phys. D 7(1–2), 181–200 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  55. Kennedy, J., York, J.A.: Basins of Wada. Phys. D 51(1–3), 213–225 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  56. Aguirre, J., Sanjuán, M.A.F.: Unpredictable behavior in the Duffing oscillator: Wada basins. Phys. D 171(1–2), 41–51 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  57. Nusse, H.E., Yorke, J.A.: Wada basin boundaries and basin cells. Phys. D 90(3), 242–261 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  58. Nusse, H.E., Yorke, J.A.: Fractal basin boundaries generated by basin cells and the geometry of mixing chaotic flows. Phys. Rev. Lett. 84, 626–629 (2000)

    Article  Google Scholar 

  59. Nusse, H.E., Yorke, J.A.: Bifurcations of basins of attraction from the view point of prime ends. Topol. Appl. 154(13), 2567–2579 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  60. McDonald, S.W., Grebogi, C., Otta, E., Yorke, J.A.: Fractal basin boundaries. Phys. D 17(2), 125–153 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  61. Smale, S.: Differentiable dynamical systems. Bull. Am. Math. Soc. 73, 118–126 (1967)

    Article  MathSciNet  Google Scholar 

  62. Guckenheimer, J., Holmes, P.: Nonlinear Oscillation, Dynamical System and Bifurcation of Vector Fields. Springer, New York (1999)

    Google Scholar 

  63. Ueda, Y.: Basin-filling Peano omega-branches and structural stability of a chaotic attractor. Nonlinear Theory Appl. IEICE 5(3), 252–258 (2014)

    Article  Google Scholar 

  64. Grebogi, C., Ott, E., York, Y.A.: Metamorphoses of basin boundaries in nonlinear dynamical systems. Phys. Rev. Lett. 56, 1011–1014 (1986)

    Article  MathSciNet  Google Scholar 

  65. Aguirre, J., Viana, R.L., Sanjuán, M.A.F.: Fractal structures in nonlinear dynamics. Rev. Mod. Phys. 81, 333–386 (2009)

    Article  Google Scholar 

  66. Souza, S.L.T.D., Caldas, I., Viana, R.L., Balthazar, J.M.: Sudden changes in chaotic attractors and transient basins in a model for rattling in gearboxes. Chaos Solitons Fractals 21, 763–772 (2004)

    Article  MATH  Google Scholar 

  67. Lai, Y.C., Tél, T.: Transient Chaos: Complex Dynamics on Finite-Time Scales. Applied Mathematical Sciences, vol. 163. Springer, New York (2010)

    Google Scholar 

  68. Grebogi, C., McDonald, S.W., Otta, E., Yorke, J.A.: Final state sensitivity: an obstruction to predictability. Phys. Lett. A 99(9), 125–153 (1985)

    MathSciNet  Google Scholar 

  69. Nusse, H.E., York, J.A.: A procedure for finding numerical trajectories on chaotic saddles. Phys. D 36(1–2), 137–156 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  70. Nusse, H.E., York, J.A.: A procedure for finding accessible trajectories on basin boundaries. Nonlinearity 4, 1183–1212 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  71. Hayashi, C., Ueda, Y., Kawakami, H.: Transformation theory as applied to the solutions of non-linear differential equations of the second order. Int. J. Non-Linear Mech. 4(3), 235–255 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  72. Coddington, E.A., Levinson, N.: Theory of Ordinary Differential Equations. McGraw-Hill, New York (1955)

    MATH  Google Scholar 

  73. Hsu, C.S.: Global analysis by cell mapping. Int. J. Bifurc. Chaos 2(4), 727–771 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  74. Hong, L., Xu, J.X.: Chaotic saddles in wada basin boundaries and their bifurcations by the generalized cell-mapping digraph (GCMD) method. Nonlinear Dyn. 32(4), 371–385 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  75. You, Z., Kostelich, E., Yorke, J.A.: Calculating stable and unstable manifolds. Int. J. Bifurc. Chaos 1, 605–624 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  76. Parker, T.S., Chua, L.O.: Practical Numerical Algorithms for Chaotic Systems. Springer, New York (1989)

    Book  MATH  Google Scholar 

Download references

Acknowledgments

The first two authors would like to acknowledge the financial supports from the NSFC (Grant Nos. 11372082,11572096) and National Basic Research Program (973 Program) of China (Grant No. 2015CB057405). ZF Hao is also indebted to the scholarship for international visiting program of Harbin Institute of Technology and the hospitality of the University of Aberdeen, where professor Yoshisuke Ueda provided many valuable suggestions in improving this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Qingjie Cao.

Appendix

Appendix

A general second-order non-autonomous dynamical system can be written as

$$\begin{aligned} \left\{ \begin{array}{l} \dot{x} = F \left( {t,x,y} \right) , \\ \dot{y} = G \left( {t,x,y} \right) , \\ \end{array} \right. \end{aligned}$$
(8)

where F(txy), G(txy) are periodic in time t with the period T and are assumed to be analytic in xy and t to guarantee the existence and uniqueness of the solution for a specific initial condition. Clearly, the discussed system in this paper is a special case of the above-mentioned dynamical system. To facilitate quick understanding of a great deal of results in this paper, some theoretical preliminaries are introduced briefly in the following:

A-1 A solution of the dynamical system (8) describes the motion starting from an arbitrary point \((x_{0}, y_{0})\) of the phase plane (xy) when \(t=t_{0}\). The solution can be written as

$$\begin{aligned} x=x(t; t_{0}, x_{0}, y_{0}), y=y(t; t_{0}, x_{0}, y_{0}), \end{aligned}$$
(9)

whose projection into the phase plane (xy) is called a phase trajectory. To study the characteristics of phase trajectories in the phase plane (xy) , a new definition of the rotation number \((\rho =m/n, m, n\in \mathbb {N})\) is given, that is, the bigger one between the average numbers of the points of the maximum in time history curves of (tx) and (ty) with respect to a period \(t=nT\), which is in good agreement with the original definition of rotation number [38] but allows one to get it from the phase portrait directly. Moreover, all the frequency spectrum components of phase trajectories can be obtained by average power spectrum, which can demonstrate the change of the rotation numbers as the parameter varies. Specially, in this paper, the solution, which possesses a rotation number m / n and hence is notated by \(P_{m/n}\), is referred to be harmonic resonance if \(m=n=1\), subharmonic resonance of n order if \(m=1, n>1\), super-harmonic resonance of m order if \(m>1, n=1\), and m / n super-subharmonic resonance if \(m>1, n>1\). In addition, for quasi-periodic or chaotic motion, the frequency components are no longer so simple but with \(\rho \) of irrational number and continuous frequency spectrum components.

A-2 For system (8), a Poincaré section can be obtained by stroboscopically taking points \(P_{n}\) at the instant \(t=t_{0}+nT\), where n is an integer and \(n=0\) corresponds to the starting point \(p_{0}=(x_{0}, y_{0})\). Moreover, the initial phase \(\theta \) of the external excitation \(F{\mathrm{cos}}(\varOmega t+\theta )\) in system (2) can alter the location of the stroboscopic points in the Poincaré section, but cannot change the topological structure of an attractor that is composed of stroboscopic points. Accordingly, without loss of generality, in this paper, \(\theta =0\) is taken. The solution series \(\{p_{n},n\in \mathbb {Z}\}\) is classified into two cases: \(\{p_{n},n=0,1,2,\ldots \}\), which is named a positive half-orbit that expresses the motion starting from \(p_{0}\) as time proceeds, and \(\{p_{n},n=0,-1,-2,\ldots \}\), which is called a negative half-orbit that denotes the behaviour when the time evolves inversely [38]. A set of points can be taken from the positive orbit to express the steady-state motion, that is, n tends to infinity (or say, n tends to a sufficiently large number) and meanwhile the transient state is abandoned. Hence, a discrete dynamical system on \(\mathbb {R}^2\) space can be defined and used in studying the system (8), that is, the Poincaré map

$$\begin{aligned} T^n: {x_0} \longmapsto {x_n},\quad n\in \mathbb {Z}. \end{aligned}$$
(10)

It is ready to study the steady-state behaviours of system (8) through observing the points in the Poincaré section: a set of n points represents for a period-n solution, and a set of infinity points means a quasi-period or chaotic motion. The chaotic behaviours can be verified by some well-developed mathematical methods, such as Lyapunov exponent, box-counting dimension, fractal dimension, correlation dimension and averaging power spectrum. Moreover, the chaotic motion has the property that two phase trajectories starting from two very close initial conditions can diverge to each other exponentially. In addition, it has been proved that in the dissipative non-autonomous dynamical system of the second order, there is no quasi-periodic solution [38, 40] .

A-3 It is significant to obtain all the periodic solutions in a nonlinear dynamical system, which can be searched with the quasi-Newton method based upon the Poincaré map [48, 49]. The stability of a period-n orbit, which is represented by n points (each of which is termed as a period-n point) in the Poincaré section, can be determined in terms of the Floquet theory [48], which will help understand the complex bifurcation of periodic solutions. The period-n point can be treated as a fixed point with a period nT, so, without loss of generality, we only need to consider the stability of a fixed or periodic point. Assume that \(p_{0}(x_{0}, y_{0})\) is a periodic point with the period T and \(p(x_{0}+u, y_{0}+v)\) is a neighbouring point of \(p_{0}\). For small uv, it can be expanded into Taylor power series with the high-order terms neglected as follows

$$\begin{aligned} \left( \begin{array}{l} {\dot{u}} \\ {\dot{v}} \\ \end{array} \right) = \left( {\begin{array}{cc} {\frac{{\partial F}}{{\partial x}}} &{}\quad {\frac{{\partial F}}{{\partial y}}} \\ {\frac{{\partial G}}{{\partial x}}} &{}\quad {\frac{{\partial G}}{{\partial y}}} \\ \end{array}} \right) _{\left| {\left( {x,y} \right) = \left( {x_0 ,y_0 } \right) } \right. } \left( \begin{array}{l} u \\ v \\ \end{array} \right) . \end{aligned}$$
(11)

Integrating (11) from \(t_0\) to \(t_0+T\) with initial conditions \(u(t_0)=1, v(t_0)=0\) and \(u(t_0)=0, v(t_0)=1\), the matrix of the fundamental solution is obtained as

$$\begin{aligned} A(t) = \left| {\begin{array}{*{20}c} {u_1 (t)} &{}\quad {u_2 (t)} \\ {v_1 (t)} &{}\quad {v_2 (t)} \\ \end{array}} \right| , \end{aligned}$$
(12)

whose eigenvalues \(\lambda _{i} (i=1,2)\) (Floquet multiplier, which is notated as \(M_{F}\)) can characterize the stability of the corresponding orbit of the periodic point in the Poincaré section. According to the Floquet’s theorem, the equation (11) has a solution with relations

$$\begin{aligned} u(t + T) = \lambda u(t), \quad v(t + T) = \lambda v(t), \end{aligned}$$
(13)

while the group of solution can be expressed with fundamental solutions as

$$\begin{aligned} u(t)=s_1{u_1(t)}+s_2{u_2(t)},\ v(t)=s_1{v_1(t)}+s_2{v_2(t)},\nonumber \\ \end{aligned}$$
(14)

which satisfy

$$\begin{aligned} \frac{{s_1 }}{{s_2 }} = \frac{{u_2 (T)}}{{\lambda - u_1 (T) }} = \frac{{ \lambda -v_2 (T) }}{{v_1 (T)}}. \end{aligned}$$
(15)

The periodic point \(p_0\) is regarded as simple if both \(|\lambda _{i}| (i=1,2\), and let \(|\lambda _{2}|\le |\lambda _{1}|\)) are unequal to 1, else it is called multiple. The simple periodic point has been classified by Levinson et al. [40] as follows: completely stable or sink if \(|\lambda _{1}|<1, |\lambda _{2}|<1\), completely unstable or source if \(|\lambda _{1}|>1, |\lambda _{2}|>1\), regular saddle or directly unstable if \(0<\lambda _{2}<1<\lambda _{1}\) and flip saddle or inversely unstable if \(\lambda _{1}<-1<\lambda _{2}<0\). This classification is complete for the system (8), because \(\lambda _{1} \lambda _{2}>0\) always holds [71, 72]. For the multiple case of periodic points, if only one of \(\lambda _{i}\) \((i=1,2)\) is 1, a small change of the system parameter can separate such periodic points and thus this case requires no special treatments. Additionally, the bifurcation scenarios of periodic solutions can be described with a unity circle in the complex plane: a saddle-node bifurcation or a symmetry-breaking bifurcation will occur if \(\lambda _{1}\) passes the circle via 1, a period-doubling or a period-halving bifurcation (the two cases are collectively called period-doubling bifurcations in this paper) will occur if \(\lambda _{1}\) goes out of the circle via \(-1\), a Neimark-Sacker bifurcation will happen when \(|\lambda _{i}| (i=1,2)\) come across the circle with a pair of conjunctive complex number. It is worthy to note that Levinson and Massera [39, 40] had provided the formula of numbers of periodic points for system (8), and proved that there are no quasi-periodic and completely unstable solutions in such a system with a positive damping coefficient. If we denote the total number of period-n solutions, the total number of completely stable and completely unstable period-n solutions, the number of directly unstable and the number of inversely unstable period-n solutions as N(n), C(n), D(n), I(n), respectively, then the following relations hold [3840]:

  1. (a)

    if \(n=1\),

    $$\begin{aligned} C(1)+I(1)=D(1)+1, N(1)=2D(1)+1;\nonumber \\ \end{aligned}$$
    (16)
  2. (b)

    if \(n=2,4,6,\ldots \),

    $$\begin{aligned} C(n)+I(n)= & {} D(n)+2I(n/2), \nonumber \\&N(n)=2[D(n)+I(n/2)]; \end{aligned}$$
    (17)
  3. (c)

    if \(n=3,5,7,\ldots \),

    $$\begin{aligned} C(n)+I(n)=D(n), N(n)=2D(n). \end{aligned}$$
    (18)

A-4 Some fundamental principles and definitions for the global bifurcations are explained briefly in the following. Firstly, the attractors contain the periodic, quasi-periodic, and chaotic ones, which can be characterized by the intersections between the phase trajectory and the Poincaré section for a periodic external forcing system. Secondly, multiple attractors often coexist with respect to different initial conditions for the same system parameters, so the collection of initial points which approach to an attractor is termed the basin of attraction of this attractor. Thirdly, there are several methods to obtain the basins of attraction, e.g. the manifold method [40, 44] and the cell-mapping method by Hsu [73, 74]. The former is utilized to explain the relations between the global bifurcations and the related dynamic phenomena in this paper. The boundary of basin of attraction is comprised of the stable manifolds of certain saddle, which is a set of initial points whose trajectories do not approach to an attractor but a saddle on the basin boundary. The pair of Floquet multipliers of the saddle can determine the directions [their slopes are \(s_1/s_2\), where \(\lambda _1\) and \(\lambda _2\) are given in (15)] of its stable and unstable manifolds (which are also termed \(\omega \)-branch and \(\alpha \)-branch, respectively) in the Poincaré section. It is worthy to note that every point on the manifold will be mapped to a point on the same manifold if the system is integrated from the point over a period of the saddle. Hence, a start point \(q_0\) on the line, which is in the direction of stable and unstable eigenvectors that are through the saddle, can be chosen to calculate the manifolds with the following procedure. Here only take the algorithm for the stable manifold as an example, and the unstable manifold can be obtained with the same routine. The start point \(q_0\) is chosen on the line that is in the direction of eigenvector of the saddle and through the saddle, which also should satisfy that the return point \(p_0\) of its backward time integration over a time period of the saddle is within a specified small distance \(\epsilon \) to the eigenvector. Next, the new start point \(q_1\) is chosen on the line between \(q_0\) and \(p_0\) till its integration return point \(p_1\) lies within the distance \(\epsilon \) of \(p_0\) on the manifold. By analogy, a sequence of integration return points of proper start points, \(p_2, p_3, p_4, \ldots \), can be found, which composes the stable manifold. The accuracy of calculation for the manifolds has been proved in [75]. Moreover, the return point of the point on a manifold must lie on the manifold, so an intersection of the stable and unstable invariant manifolds implies an infinite number of intersections between the two manifolds. The intersections of stable and unstable manifolds of a saddle point, or intersections of manifolds of the same group of highly periodic saddles, are called homoclinic points, while the intersections between manifolds issued from different saddles are called heteroclinic points. Homoclinic and hereroclinic points may induce a chaotic attractor [76], and similarly both homoclinic and hereroclinic points can also result in a fractal boundary of basin of attraction [64]. Additionally, it should be pointed out that two different stable manifolds never intersect each other: because if not so, it will imply an infinite number of intersections between them, and then the saddle is one of the intersection points, which apparently does not hold. Likewise, the unstable manifolds are also in the same case.

A-5 To understand the complicated metamorphosis of boundary of basin of attraction, some critical concepts and theorems are introduced in the following. As well known, the coexistence of multiple attractors is a common phenomenon in a nonlinear dynamical system. All the initial points that approach to an attractor compose the basin of attraction of this attractor. Obviously, there is basin boundary for a basin of attraction, B. To verify the complex property of basin boundary, a boundary point is firstly defined as such a point whose every open neighbourhood possesses a non-empty intersection with B and at least one other basin. All the boundary points of B constitute the boundary of B that is written as \(\partial \bar{B}\). The basin boundary is fractal if it contains a transversal homoclinic point. Moreover, there exist more complex properties for some basin boundaries, such as Wada point, Wada basin and Wada property [5559, 65]. Those boundaries involve at least three coexisting basins of attraction. Several definitions and theorems are important to prove them. First, a boundary point is a Wada point if every open neighbourhood of the point has a non-empty intersection with at least three basins. The basin is called Wada basin if each boundary point is a Wada point, while it is called partial Wada basin if only partially boundary points are Wada points. Furthermore, the Wada property of attraction basin, which usually implies strong complexity for basin boundary in the sense of predictability, denotes that each boundary point of one basin is also the boundary point of all other basins. The Wada property is based upon a theorem in [55] to supply a sufficient condition for its existence: assume p is a periodic point on the boundary such that (i) the unstable manifold of p intersects every basin and (ii) the stable manifold of p is dense in each of the basin boundaries. It is easy to understand but difficult to verify it numerically. Nusse and York [57] supplied several effective theorems to address this issue, of which two theorems used in this paper are listed in the end of this section. Only by applying the Theorem 1, it is still difficult to validate the existence of Wada basin, because we have to find all the accessible hyperbolic points, but at least it can determine the existence of partial Wada point, while, according to Theorem 2, we only need to construct basin cell for the verification of Wada basin. In the two theorems and their corollaries, an accessibility hypothesis is considered as a premise, that is, every accessible basin boundary point is on the stable manifold of certain hyperbolic periodic point. In fact, the Theorem 2 implies that the accessibility hypothesis is equivalent to the existence of basin cell [57]. To understand the theorems about basin cell, it is necessary to introduce accessible point and the construction of basin cell. We say a point p in \(\partial \bar{B}\) is accessible from basin B if a curve can be plotted starting from B such that p is the first boundary point of B that the curve hits. Secondly, basin cell Q of basin B is a compact trapping region, that is, a trajectory in B once enters into Q and will never leave again. Hence, a basin cell must contain an attractor. Basin cell is constructed by pieces of stable and unstable manifolds of a periodic saddle that situates on the boundary of a basin of attraction, for which some schematic diagrams are given in [57]. Here the description about basin cell is given: assume that a group of period-n saddles \(P=\{ p_{ {i}}|1\le i\le n \}\) generates a basin cell C (with boundary \(\partial C\)), we say, \(\partial C \cap W^{\mathrm{s}}(p_i), \partial C \cap W^{\mathrm{u}}(p_i)\) are the stable and unstable edges of the cell C, respectively, and the common point of a stable and an unstable manifolds is called a corner point, while a piece of unstable manifold of \(p_{ {i}}\) that links \(p_{ {i}}\) and a corner point is called a supporting arc, which does not contain any other point of C. Therefore, we call a basin cell developed by a period-n saddle as a 2n-sided cell.

Theorem 1

([57]) Let p be a \(B_1\)-accessible hyperbolic periodic point, and assume that the unstable manifold \(W^{\mathrm{u}}(p)\) intersects each \(B_k\), where \(1 \le k \le N\). Then (a) every point z on \(W^{\mathrm{s}}(p)\) is a Wada point with respect to the basins \(B_k\) (\(1 \le k \le N\)), (b) if the orbit of p is the only \(B_1\)-accessible periodic orbit, then \(\partial \bar{B}_1\) is a Wada basin boundary with respect to \(B_k\) (\(1 \le k \le N\)).

Theorem 2

([57]) Let \(P_1\) be a hyperbolic periodic orbit that generates a basin cell \(C_1\). Let \(B_k\) (\(1 \le k \le N\)) be N disjoint basins (\(N\ge 3\)), where \(B_1\) is the basin of \(C_1\). Assume that the unstable manifold \(W^{\mathrm{u}}(P_k)\) intersects each basin \(B_k\) (\(1 \le k \le N\)), then \(\partial \bar{B}_1\) is a Wada basin boundary with respect to \(B_k\) (\(1 \le k \le N\)).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hao, Z., Cao, Q. & Wiercigroch, M. Nonlinear dynamics of the quasi-zero-stiffness SD oscillator based upon the local and global bifurcation analyses. Nonlinear Dyn 87, 987–1014 (2017). https://doi.org/10.1007/s11071-016-3093-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-016-3093-6

Keywords

Navigation