Skip to main content
Log in

Self-excited oscillation produced by a phase shift: linear and nonlinear instabilities

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

Abstract

Self-excited oscillation is a method to make a resonator vibrate at its natural frequency. This is widely used in vibration sensors such as mass sensors, atomic force microscopes and stiffness sensors. Self-excited oscillation is mainly produced by positive velocity feedback and time-delayed displacement feedback including phase-shifted displacement feedback. In this paper, we consider phase-shifted displacement feedback using a phase shifter. We perform nonlinear analysis to clarify the finite steady-state amplitude of the self-excited oscillation by considering nonlinear damping without a Laplace transform in the frequency domain, and formulate the effect of the phase shifter using an ordinary differential equation. We apply the method of multiple scales to the third-order ordinary differential equations expressing the coupling between the resonator and the phase shifter. This analytically reveals the parameter range of the phase shifter that produces self-excited oscillation in the resonator and the steady-state amplitude depending on the phase shift. We conduct an experiment using a cantilever as the resonator and produce self-excited oscillation via the feedback signal based on a phase shifter in a digital computer. The theoretically predicted characteristics of the self-excited oscillation agree well with the experimental ones.

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

Similar content being viewed by others

Data availibility

The datasets of the current study are available from the corresponding author on reasonable request.

References

  1. Ono, T., Li, X., Miyashita, H.: Mass sensing of adsorbed molecules in sub-picogram sample with ultrathin silicon resonator. Rev. Sci. Instrum. 74, 1240 (2003)

    Article  Google Scholar 

  2. Lee, Y., Lim, G., Moon, W.: A self-excited micro cantilever biosensor actuated by PZT using the mass micro balancing technique. Sens. Actuators A. 130–131, 105–110 (2006)

    Article  Google Scholar 

  3. Okajima, T., Sekiguchi, H., Arakawa, H., Ikai, A.: Self-oscillation technique for AFM in liquids. Appl. Surf. Sci. 210, 68 (2003)

    Article  Google Scholar 

  4. Yabuno, H., Kaneko, H., Kuroda, M., Kobayashi, T.: Van der Pol type self-excited micro-cantilever probe of atomic force microscopy. Nonlin. Dyn. 54, 137–149 (2008)

    Article  Google Scholar 

  5. Sun, X., Yuan, W., Qiao, D., Sun, M., Ren, S.: Design and analysis of a new tuning fork structure for resonant pressure sensor. Micromachines 7(9), 148 (2016)

    Article  Google Scholar 

  6. Greenleaf, J., Fatemi, M., Insana, M.: Selected methods for imaging elastic properties of biological tissues. Annu. Rev. Biomed. 5, 57 (2003)

    Article  Google Scholar 

  7. Sacks, M.: Biaxial mechanical evaluation of planar biological materials. J. Elast. 61, 199 (2000)

    Article  Google Scholar 

  8. Vashisht, R.K., Peng, Q.: Fractional calculus-based energy efficient active chatter control of milling process using small size electromagnetic actuators. ASME. J. Vib. Acoust. 143(1), 011005 (2021)

    Article  Google Scholar 

  9. Kasai, Y., Yabuno, H., Ishine, T., Yamamoto, T., Matsumoto, S.: Mass sensing using a virtual cantilever virtually coupled with a real cantilever. Appl. Phys. Lett. 115, 063103 (2019)

    Article  Google Scholar 

  10. Yang, J., Yabuno, H., Yanagisawa, N., Yamamoto, Y., Matsumoto, S.: Measurement of added mass for an object oscillating in viscous fluids using nonlinear self-excited oscillations. Nonlin. Dyn. 102, 1987–1996 (2020)

    Article  Google Scholar 

  11. Malas, A., Chatterjee, S.: Modal self-excitation in a class of mechanical systems by nonlinear displacement feedback. J. Vib. Control. 24(4), 784–796 (2018)

    Article  MathSciNet  Google Scholar 

  12. Malas, A., Chatterjee, S.: Modal self-excitation by nonlinear acceleration feedback in a class of mechanical systems. J. Sound Vib. 376, 1–17 (2016)

    Article  Google Scholar 

  13. Patel, B., Kundu, P.K., Chatterjee, S.: Nonlinear feedback anti-control of limit cycle and chaos in a mechanical oscillator: theory and experiment. Nonlin. Dyn. 104, 3223–3246 (2021)

    Article  Google Scholar 

  14. Chatterjee, S.: Self-excited oscillation under nonlinear feedback with time-delay. J. Sound Vib. 330(9), 1860–1876 (2011)

    Article  Google Scholar 

  15. Maccari, A.: Vibration amplitude control for a van der Pol-Duffing oscillator with time delay. J. Sound Vib. 317(1–2), 20–29 (2008)

    Article  Google Scholar 

  16. Hou, A., Guo, S.: Stability and Hopf bifurcation in van der Pol oscillators with state-dependent delayed feedback. Nonlin. Dyn 79, 2407–2419 (2015)

    Article  MathSciNet  Google Scholar 

  17. Sarkar, A., Mondal, J., Chatterjee, S.: Controlling self-excited vibration using positive position feedback with time-delay. J Braz. Soc. Mech. Sci. Eng. 42, 464 (2020)

    Article  Google Scholar 

  18. Rodriguez, T., Garcia, R.: Theory of Q control in atomic force microscopy. Appl. Phys. Lett. 82, 4821 (2003)

    Article  Google Scholar 

  19. Mourol, J., Tiribilli, B., Paoletti, P.: Nonlinear behaviour of self-excited microcantilevers in viscous fluids. J. Micromech. Microeng. 27, 095008 (2017)

    Article  Google Scholar 

  20. Mourol, J., Tiribilli, B., Paoletti, P.: A versatile mass-sensing platform with tunable nonlinear self-excited microcantilevers. IEEE Trans. Nanotechnol. 17(4), 751–762 (2018)

    Article  Google Scholar 

  21. Dowling, A.: Nonlinear self-excited oscillations of a ducted flame. J. Fluid Mech. 346, 271–290 (1996)

    Article  MathSciNet  Google Scholar 

  22. Li, J., Li, X., Hua, H.: Active nonlinear saturation-based control for suppressing the free vibration of a self-excited plant. Commun. Nonlin. Sci. Numer. Simul. 15(4), 1071–1079 (2010)

    Article  MathSciNet  Google Scholar 

  23. Lin, Y., Yabuno, H., Yamamoto, Y., Liu, X., Matsumoto, S.: Highly sensitive AFM using self-excited weakly coupled cantilevers. Appl. Phys. Lett. 115(13), 133105 (2019)

    Article  Google Scholar 

  24. Nakamura, T., Yabuno, H., Yano, M.: Amplitude control of self-excited weakly coupled cantilevers for mass sensing using nonlinear velocity feedback control. Nonlin. Dyn. 99, 85–97 (2020)

    Article  Google Scholar 

  25. Habib, G., Rega, G., Stepan, G.: Nonlinear bifurcation analysis of a single-dof model of a robotic arm subject to digital position control. ASME. J. Comput. Nonlin. Dyn. 8(1), 011009 (2013)

    Article  Google Scholar 

  26. Nayfeh, A., Mook, T.: Nonlinear Oscillations, p. 105. Wiley, New York (1995)

    Book  Google Scholar 

  27. van Der Pol, B.: On “relaxation oscillations.” Phil. Mag. J. Sci. 2, 978–992 (1926)

  28. Nayfeh, A., Mook, T.: Nonlinear Oscillations, p. 56. Wiley, New York (1995)

    Book  Google Scholar 

  29. Meirovitch, L.: Elements of Vibration Analysis, p30, McGrraw-Hill (1975)

Download references

Acknowledgements

This work was supported by a Grant-in-Aid for Scientific Research B (Grant No. 16H02318) from the Japan Society for the Promotion of Science (JSPS).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Linjun An.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

Appendices

Appendix A

We show the method to experimentally identify the nondimensional coefficients of linear and cubic nonlinear damping via the free vibration of the cantilever beam as an experimentally obtained time history in Fig. 11. Letting \(x_e\) be 0, we obtain the equation governing the free vibration as

$$\begin{aligned} \ddot{x}^*+\gamma {\dot{x}}^*+x^*+\gamma _3{\dot{x}}^{*3}=0, \end{aligned}$$
(A.1)

where the dot denotes the derivative with respect to nondimensional time \(t^*\). The parameters \(\gamma \), \(\gamma _3\) and nondimensional time \(t^*\) have the same definitions as in Sect. 2. Then, Eqs. (12), (14), (17) and (21) are simplified, respectively, as

$$\begin{aligned}&O(\epsilon ^{\frac{1}{2}}) :\quad {D_0^2}x_{1}+x_{1}=0, \end{aligned}$$
(A.2)
$$\begin{aligned}&O(\epsilon ^{\frac{3}{2}}) :\quad {D_0^2}x_{3}+x_{3}=-2D_0D_1x_{1}-{\hat{\gamma }}D_0x_{1}-{\hat{\gamma }}_3\left( D_0x_{1}\right) ^3, \end{aligned}$$
(A.3)
$$\begin{aligned}&x_{1}=A(t_1)\text{ e}^{i{t_0}}+CC, \end{aligned}$$
(A.4)
$$\begin{aligned}&D_1A+\frac{1}{2}{\hat{\gamma }}A+\frac{3}{2}{\hat{\gamma }}_3|A|^2A=0. \end{aligned}$$
(A.5)

Substituting Eqs. (18) and (A.4) into Eq. (A.5) and separating the resulting equation into the real and imaginary parts yields

figure b

We obtain the following functions of nondimensional amplitude a and \(\phi \) with respect to time \(t^*\) based on Eq. (A.6):

$$\begin{aligned}&a=\frac{1}{\sqrt{(\frac{1}{a(0)^2}+\frac{3\gamma _3}{4\gamma })\text{ e}^{\gamma t^*}-\frac{3\gamma _3}{4\gamma }}}, \end{aligned}$$
(A.8)
$$\begin{aligned}&\phi =\phi _0, \end{aligned}$$
(A.9)

where a(0) and \(\phi _0\) denote the initial amplitude and the integral constant determined by the initial condition, respectively. The solution of \(x^*\) obtained through Eqs. (8), (18), (A.4), (A.8) and (A.9) is

$$\begin{aligned} x^*=a\cos {(t^*+\phi _0)}+O(\epsilon ^{\frac{3}{2}}). \end{aligned}$$
(A.10)

Using the envelope theoretically obtained from Eq. (A.8) and the peak of each period of the experimentally obtained time history in Fig. 11, we identify \(\gamma \) and \(\gamma _3\). When the amplitude of the free vibration is small, the cubic nonlinear damping is extremely small compared with the linear damping. Therefore, we can neglect the nonlinear damping in the time history as shown in Fig. 11b. In this case, we can calculate \(\gamma \) to be 0.0022 from the average of the logarithmic decrements [29], which are obtained from the 17 peaks in Fig. 11b. However, when the amplitude of the free vibration is comparatively large as shown in Fig. 11c, there is the effect of nonlinear damping. We use the time history from 0 s to 100 s in Fig. 11c to determine the coefficient of nonlinear damping \(\gamma _3\); the number of peaks is \(n=17\). We set the value of every peak in Fig. 11c as \(p_i\) (\(i=1,2,\cdots , n\)) successively from 0 s to 100 s. Substituting the moment \(t^*\) corresponding to each \(p_i\) and \(\gamma =0.0022\) into Eq. (A.8), we can obtain the function \(a_i(\gamma _3)\) corresponding to \(p_i\). We calculate \(F(\gamma _3)=\sum _{i=1}^{n}\{p_i-a_i(\gamma _3)\}^2\) and determine \(\gamma _3\) so that \(\frac{\partial F(\gamma _3)}{\partial \gamma _3}=0\). As a result, \(\gamma \) and \(\gamma _3\) are experimentally identified as shown in Table 1. The black dashed line in Fig. 11a shows the envelope obtained from Eq. (A.8) using the values of \(\gamma \) and \(\gamma _3\) obtained above.

Appendix B

We investigate the sensitivity of the Hopf bifurcation points \(\alpha _{cr1}\) and \(\alpha _{cr2}\) to the uncertainties in \(\beta \) and \(\gamma \) using Eq. (26). Let the unknown exact values of \(\beta \) and \(\gamma \) be \(\beta _0\) and \(\gamma _0\). Then, the uncertainties in \(\beta \) and \(\gamma \) are defined as \(\varDelta \beta =\beta -\beta _0\) and \(\varDelta \gamma =\gamma -\gamma _0\). The Taylor expansions of \(\alpha _{cr1}\) and \(\alpha _{cr2}\) with respect to the uncertainties, \(\varDelta \beta \) and \(\varDelta \gamma \), are as follows:

$$\begin{aligned} \alpha _{cr1}(\gamma ,\beta )&= \alpha _{cr10}-\frac{\beta }{\gamma ^2}\left( 1-\frac{1}{\sqrt{1-c^2}}\right) \varDelta \gamma \nonumber \\&\qquad \!\!\! +\frac{1}{\gamma }\left( 1-\frac{1}{\sqrt{1-c^2}}\right) \varDelta \beta +O(\varDelta \gamma ^2,\varDelta \beta ^2), \end{aligned}$$
(B.1)
$$\begin{aligned} \alpha _{cr2}(\gamma ,\beta )&=\alpha _{cr20}-\frac{\beta }{\gamma ^2}\left( 1+\frac{1}{\sqrt{1-c^2}}\right) \varDelta \gamma \nonumber \\&\qquad \!\!\! +\frac{1}{\gamma }\left( 1+\frac{1}{\sqrt{1-c^2}}\right) \varDelta \beta +O(\varDelta \gamma ^2,\varDelta \beta ^2), \end{aligned}$$
(B.2)

where c=\(\gamma /\beta \), and \(\alpha _{cr10}\) and \(\alpha _{cr20}\) are two unknown exact values of \(\alpha _{cr}\) for two Hopf bifurcation points corresponding to \(\beta _{0}\) and \(\gamma _{0}\). The coefficients of \(\varDelta \beta \) and \(\varDelta \gamma \) are the sensitivities of the Hopf bifurcation points to the uncertainties in \(\beta \) and \(\gamma \). Substituting the values of \(\beta \) and \(\gamma \) in Table 1 into Eqs. (B.1) and (B.2) yields

$$\begin{aligned}&\alpha _{cr1}-\alpha _{cr10}=128\varDelta \gamma -60\varDelta \beta \end{aligned}$$
(B.3)
$$\begin{aligned}&\alpha _{cr2}-\alpha _{cr20}=2070\varDelta \gamma +969\varDelta \beta . \end{aligned}$$
(B.4)

We can see by comparing the coefficients of \(\varDelta \gamma \) and \(\varDelta \beta \) in Eqs. (B.3) and (B.4) that \(\alpha _{cr2}\) is more affected by these uncertainties than \(\alpha _{cr1}\). This indicates that the deviation between the theoretical and experimental bifurcation points at \(\alpha _{cr2}\) is larger than that at \(\alpha _{cr1}\), as observed in the experiment (Fig. 5). Therefore, there is a relatively large discrepancy between the deviations near \(\alpha _{cr2}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

An, L., Yabuno, H. Self-excited oscillation produced by a phase shift: linear and nonlinear instabilities. Nonlinear Dyn 107, 587–597 (2022). https://doi.org/10.1007/s11071-021-07060-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-021-07060-4

Keywords

Navigation