Skip to main content
Log in

Influences of inherent geometrical nonlinearity of high-static-low-dynamic-stiffness resonator on flexural wave attenuation performance of metamaterial beam

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

Abstract

Metamaterials with high-static-low-dynamic stiffness (HSLDS) resonators possess the ability in low/ultra-low frequency flexural wave attenuation. However, the inherent geometrical nonlinearity of the HSLDS resonators was used to be neglected for simplicity, leading to severe restriction of its operating range. In this paper, the intrinsic geometric nonlinearity of the HSLDS resonator is considered to extend its applications. The nonlinear dynamic model of a metamaterial beam coupled with nonlinear HSLDS resonators is established based on the Galerkin method. An efficient frequency-domain analysis strategy, which consists of the harmonic balance method (HBM), arc-length continuation, and Hill’s method, is then adopted to compute periodic solution branches. Results show that once the excitation amplitude reaches some extent and the nonlinear stiffness of the resonator plays a dominant role due to higher response, only the nonlinear model can capture the dynamics of the coupling system. Furthermore, in the case of large responses, double peaks in frequency response bend right, and the second peak is significantly suppressed. Meanwhile, the location and wave attenuation performance of the bandgap remains unchanged. The results indicate that the nonlinear metamaterial beam performs better in the range out of bandgap. Hence, the nonlinear metamaterial is more robust against mistuning. Due to the above merit, the operating range of HSLDS resonators and the bearable excitation amplitude in nonlinear metamaterial beam are enhanced four times compared to linear metamaterial beam. Thus, the metamaterial beam with nonlinear HSLDS resonators has better environmental adaptability and broader application prospects.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Data availability

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

References

  1. Dong, Y., Itoh, T.: Metamaterial-based antennas. Proc. IEEE 100(7), 2271–2285 (2012)

    Google Scholar 

  2. Park, J., Youn, J.R., Song, Y.S.: hydrodynamic metamaterial cloak for drag-free flow. Phys. Rev. Lett. 123(7), 074502 (2019)

    Google Scholar 

  3. Park, J.J., et al.: Acoustic superlens using membrane-based metamaterials. Appl. Phys. Lett. 106(5), 051901 (2015)

    Google Scholar 

  4. Iemma, U.: Theoretical and numerical modeling of acoustic metamaterials for aeroacoustic applications. Aerospace 3(2), 15 (2016)

    Google Scholar 

  5. Hashemi, M.R., Cakmakyapan, S., Jarrahi, M.: Reconfigurable metamaterials for terahertz wave manipulation. Rep. Prog. Phys. 80(9), 094501 (2017)

    Google Scholar 

  6. Zheng, Y., et al.: A piezo-metastructure with bistable circuit shunts for adaptive nonreciprocal wave transmission. Smart Mater. Struct. 28(4), 045005 (2019)

    Google Scholar 

  7. Emerson, T.A., Manimala, J.M.: Passive-adaptive mechanical wave manipulation using nonlinear metamaterial plates. Acta Mech. 231(11), 4665–4681 (2020)

    Google Scholar 

  8. Elmadih, W., et al.: Three-dimensional resonating metamaterials for low-frequency vibration attenuation. Sci. Rep. 9(1), 11503 (2019)

    Google Scholar 

  9. Xu, X., et al.: Tailoring vibration suppression bands with hierarchical metamaterials containing local resonators. J. Sound Vib.. 442, 237–248 (2019)

    Google Scholar 

  10. Sigalas, M.M., Economou, E.N.: Elastic and acoustic wave band structure. J. Sound Vib. 158(2), 377–382 (1992)

    Google Scholar 

  11. Liu, Z., et al.: Locally resonant sonic materials. Science 289(5485), 1734–1736 (2000)

    Google Scholar 

  12. Pai, P.F., Peng, H., Jiang, S.: Acoustic metamaterial beams based on multi-frequency vibration absorbers. Int. J. Mech. Sci. 79, 195–205 (2014)

    Google Scholar 

  13. Chen, Y.Y., Huang, G.L., Sun, C.T.: Band gap control in an active elastic metamaterial with negative capacitance piezoelectric shunting. J. Vib. Acoust. 136(6) 061008 (2014)

    Google Scholar 

  14. Frandsen, N.M.M., et al.: Inertial amplification of continuous structures: Large band gaps from small masses. J. Appl. Phys. 119(12), 124902 (2016)

    Google Scholar 

  15. Ma, J., et al.: Dynamic analysis of periodic vibration suppressors with multiple secondary oscillators. J. Sound Vib. 424, 94–111 (2018)

    Google Scholar 

  16. Droz, C., et al.: Improving sound transmission loss at ring frequency of a curved panel using tunable 3D-printed small-scale resonators. J. Acoust. Soc. America 145(1), 72–78 (2019)

    Google Scholar 

  17. Meng, H., et al.: Rainbow metamaterials for broadband multi-frequency vibration attenuation: numerical analysis and experimental validation. J. Sound Vib. 465, 115005 (2020)

    Google Scholar 

  18. Oudich, M., et al.: A sonic band gap based on the locally resonant phononic plates with stubs. New J. Phys. 12(8), 083049 (2010)

    Google Scholar 

  19. Oudich, M., et al.: Experimental evidence of locally resonant sonic band gap in two-dimensional phononic stubbed plates. Phys. Rev. B 84(16), 165136 (2011)

    Google Scholar 

  20. Carrella, A., Brennan, M.J., Waters, T.P.: Static analysis of a passive vibration isolator with quasi-zero-stiffness characteristic. J. Sound Vib. 301(3), 678–689 (2007)

    Google Scholar 

  21. Carrella, A., et al.: Force and displacement transmissibility of a nonlinear isolator with high-static-low-dynamic-stiffness. Int. J. Mech. Sci. 55(1), 22–29 (2012)

    Google Scholar 

  22. Liu, X., Huang, X., Hua, H.: On the characteristics of a quasi-zero stiffness isolator using Euler buckled beam as negative stiffness corrector. J. Sound Vib. 332(14), 3359–3376 (2013)

    Google Scholar 

  23. Huang, X., et al.: Vibration isolation characteristics of a nonlinear isolator using Euler buckled beam as negative stiffness corrector: a theoretical and experimental study. J. Sound Vib. 333(4), 1132–1148 (2014)

    Google Scholar 

  24. Fulcher, B.A., et al., Analytical and experimental investigation of buckled beams as negative stiffness elements for passive vibration and shock isolation systems. J. Vib. Acoust. 136(3) 031009 (2014)

    Google Scholar 

  25. Dong, G., et al.: Simulated and experimental studies on a high-static-low-dynamic stiffness isolator using magnetic negative stiffness spring. Mech. Syst. Signal Process. 86, 188–203 (2017)

    Google Scholar 

  26. Zheng, Y., et al.: A Stewart isolator with high-static-low-dynamic stiffness struts based on negative stiffness magnetic springs. J. Sound Vib. 422, 390–408 (2018)

    Google Scholar 

  27. Zhou, J., et al.: Local resonator with high-static-low-dynamic stiffness for lowering band gaps of flexural wave in beams. J. Appl. Phys. 121(4), 044902 (2017)

    MathSciNet  Google Scholar 

  28. Wang, K., et al.: Lower band gaps of longitudinal wave in a one-dimensional periodic rod by exploiting geometrical nonlinearity. Mech. Syst. Signal Process. 124, 664–678 (2019)

    Google Scholar 

  29. Cai, C., et al.: Design and numerical validation of quasi-zero-stiffness metamaterials for very low-frequency band gaps. Compos. Struct. 236, 111862 (2020)

    Google Scholar 

  30. Peng, F., et al.: Low-frequency vibration suppression of metastructure beam with high-static–low-dynamic stiffness resonators employing magnetic spring. J. Vib. Control 30(1–2), 237–249

    MathSciNet  Google Scholar 

  31. Wu, Q., et al.: Low-frequency multi-mode vibration suppression of a metastructure beam with two-stage high-static-low-dynamic stiffness oscillators. Acta Mech. 230(12), 4341–4356 (2019)

    MathSciNet  Google Scholar 

  32. Cveticanin, L., Mester, G.: Theory of acoustic metamaterials and metamaterial beams: an overview. Acta Polytechnica Hungarica 13(7), 43–62 (2016)

    Google Scholar 

  33. Cveticanin, L., Cveticanin, D.: Application of the acoustic metamaterial in engineering: an overview. Rom. J. Mech. 2(1), 29–36 (2017)

    Google Scholar 

  34. Cveticanin, L., Zukovic, M.: Negative effective mass in acoustic metamaterial with nonlinear mass-in-mass subsystems. Commun. Nonlinear Sci. Numer. Simul. 51, 89–104 (2017)

    MathSciNet  Google Scholar 

  35. Cveticanin, L., Zukovic, M., Cveticanin, D.: On the elastic metamaterial with negative effective mass. J. Sound Vib. 436, 295–309 (2018)

    Google Scholar 

  36. Fang, X., et al.: Broadband and tunable one-dimensional strongly nonlinear acoustic metamaterials: theoretical study. Phys. Rev. E 94(5), 052206 (2016)

    Google Scholar 

  37. Fang, X., et al.: Wave propagation in a nonlinear acoustic metamaterial beam considering third harmonic generation. New J. Phys. 20(12), 123028 (2018)

    Google Scholar 

  38. Casalotti, A., El-Borgi, S., Lacarbonara, W.: Metamaterial beam with embedded nonlinear vibration absorbers. Int. J. Non-Linear Mech. 98, 32–42 (2018)

    Google Scholar 

  39. Xie, L., et al.: Bifurcation tracking by harmonic balance method for performance tuning of nonlinear dynamical systems. Mech. Syst. Signal Process. 88, 445–461 (2017)

    Google Scholar 

  40. Wu, J., Hong, L., Jiang, J.: A robust and efficient stability analysis of periodic solutions based on harmonic balance method and Floquet-Hill formulation. Mech. Syst. Signal Process. 173, 109057 (2022)

    Google Scholar 

  41. Wu, J., Hong, L., Jiang, J.: A comparative study on multi- and variable-coefficient harmonic balance methods for quasi-periodic solutions. Mech. Syst. Signal Process. 187, 109929 (2023)

    Google Scholar 

  42. Cameron, T.M., Griffin, J.H.: An alternating frequency/time domain method for calculating the steady-state response of nonlinear dynamic systems. J. Appl. Mech. 56(1), 149–154 (1989)

    MathSciNet  Google Scholar 

  43. Moore, G.: Floquet theory as a computational tool. SIAM J. Numer. Anal. 42(6), 2522–2568 (2005)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 11872290), NSAF (Grant No. U1430129), and the National Natural Science Foundation of China (Grant No. 12072247).

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 11872290), NSAF (Grant No. U1430129), and the National Natural Science Foundation of China (Grant No. 12072247).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shilin Xie.

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.

Appendix

Appendix

1.1 Harmonic balance method

Introduce the derivative operator:

$$ \nabla = {\text{diag}}(0,{\kern 1pt} {\kern 1pt} {\kern 1pt} \nabla_{1} , \ldots ,{\kern 1pt} {\kern 1pt} {\kern 1pt} \nabla_{j} , \ldots ,{\kern 1pt} {\kern 1pt} {\kern 1pt} \nabla_{H} ){\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\text{with}}{\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} \nabla_{j} = j\left[ {\begin{array}{*{20}c} 0 & 1 \\ { - 1} & 0 \\ \end{array} } \right] $$
(A1)

The velocity and acceleration can be expressed by Fourier coefficient vectors, which are given by Eq. (17):

$$ \begin{gathered} \dot{x}(t) = \omega ({\varvec{T}}(\omega t) \otimes {\varvec{I}}_{n} )(\nabla \otimes {\varvec{I}}_{n} ){\varvec{X}} = \omega [({\varvec{T}}(\omega t)\nabla ) \otimes {\varvec{I}}_{n} ]{\varvec{X}} \hfill \\ \ddot{x}(t) = \omega^{2} ({\varvec{T}}(\omega t) \otimes {\varvec{I}}_{n} )(\nabla^{2} \otimes {\varvec{I}}_{n} ){\varvec{X}} = \omega^{2} [({\varvec{T}}(\omega t)\nabla^{2} ) \otimes {\varvec{I}}_{n} ]{\varvec{X}} \hfill \\ \end{gathered} $$
(A2)

Substituting Eq. (16), Eq. (A1) and Eq. (A2) into Eq. (15), and balancing the harmonic terms by utilizing the orthogonality property of trigonometric basis, the nonlinear system Eq. (15) of differential equations in the time domain is transformed into nonlinear algebraic equations in the frequency domain:

$$ {\varvec{R}}({\varvec{X}},\omega ) = {\varvec{Z}}(\omega ){\varvec{X}} + {\varvec{F}}_{nl} ({\varvec{X}}) - {\varvec{P}} = {\mathbf{0}} $$
(A3)

where \({\varvec{Z}}(\omega )\) is the dynamic stiffness matrix, which is block-diagonal:

$$ \begin{gathered} {\varvec{Z}}(\omega ) = \omega^{2} \nabla^{2} \otimes {\varvec{M}} + \omega \nabla \otimes {\varvec{C}} + {\varvec{I}}_{2H + 1} \otimes {\varvec{K}} = {\text{diag}}({\varvec{K}},{\kern 1pt} {\kern 1pt} {\kern 1pt} {\varvec{Z}}_{1} , \ldots ,{\kern 1pt} {\kern 1pt} {\kern 1pt} {\varvec{Z}}_{j} , \ldots ,{\kern 1pt} {\kern 1pt} {\kern 1pt} {\varvec{Z}}_{H} ) \hfill \\ {\varvec{Z}}_{j} = \left[ {\begin{array}{*{20}c} {{\varvec{K}} - j^{2} \omega^{2} {\varvec{M}}} & {\omega {\varvec{C}}} \\ { - \omega {\varvec{C}}} & {{\varvec{K}} - j^{2} \omega^{2} {\varvec{M}}} \\ \end{array} } \right] \hfill \\ \end{gathered} $$
(A4)

With the above derivation of equations of motion, the problem is converted into finding the Fourier coefficients of Eq. (15), which can be accomplished by the New-Raphson procedure. Note that the nonlinear term \({\varvec{f}}_{nl}\) and \({\varvec{F}}_{nl}\) cannot be directly computed in the frequency domain. The alternating frequency-time(AFT) scheme [42], which computes the nonlinear term in the time domain and then switches back to the frequency domain, is coupled with standard HBM.

1.2 Tangent prediction

With the calculated solution point \(({\varvec{X}}_{i} ;\omega_{i} )\) on the hand, the predicted point \((\overline{\user2{X}}_{i + 1} ;\overline{\omega }_{i + 1} )\) can be obtained using the first-order Taylor expansion of Eq. (15) at point \(({\varvec{X}}_{i} ;\omega_{i} )\). After neglecting high-order terms and introducing a constrain equation, a tangent vector \({\varvec{t}} = (\Delta {\varvec{X}};\Delta \omega )\) can be obtained from

$$ \left[ {\begin{array}{*{20}c} {\left. {{\varvec{R}}_{{\varvec{X}}} } \right|_{{({\varvec{X}}_{i + 1} ;\omega_{i + 1} )}} } & {\left. {{\varvec{R}}_{\omega } } \right|_{{({\varvec{X}}_{i + 1} ;\omega_{i + 1} )}} } \\ {{\varvec{t}}_{i - 1}^{T} } & {} \\ \end{array} } \right]\left[ {\begin{array}{*{20}c} {\Delta {\varvec{X}}_{i} } \\ {\Delta \omega_{i} } \\ \end{array} } \right] = \left[ {\begin{array}{*{20}c} {\mathbf{0}} \\ 1 \\ \end{array} } \right] $$
(A5)

where \({\varvec{R}}_{{\varvec{X}}} = {{\partial {\varvec{R}}} \mathord{\left/ {\vphantom {{\partial {\varvec{R}}} {\partial {\varvec{X}}}}} \right. \kern-0pt} {\partial {\varvec{X}}}}\) is the Jacobian matrix and is given by:

$$ \begin{gathered} {\varvec{R}}_{\omega } = {\varvec{Z}}_{\omega } {\varvec{X}} \hfill \\ \;{\varvec{Z}}_{\omega } = \frac{{\partial {\varvec{Z}}}}{\partial \omega } = 2\omega \nabla^{2} \otimes {\varvec{M}} + \nabla \otimes {\varvec{C}} \hfill \\ \end{gathered} $$
(A6)

Then, a normalization condition is added to unify the tangent vector with respect to the parameter \(\omega\):

$$ {\varvec{t}}_{i} = \left[ {\frac{{\Delta {\varvec{X}}_{i} }}{{\left| {\Delta \omega_{i} } \right|}};\frac{{\Delta \omega_{i} }}{{\left| {\Delta \omega_{i} } \right|}}} \right] $$
(A7)

Eventually, the predicted point can be given at a distance \(\Delta s_{i}\) along the direction \({\varvec{t}}_{i}\):

$$ \left[ {\begin{array}{*{20}c} {\overline{\user2{X}}_{i + 1} } \\ {\overline{\omega }_{i + 1} } \\ \end{array} } \right] = \left[ {\begin{array}{*{20}c} {{\varvec{X}}_{i} } \\ {\omega_{i} } \\ \end{array} } \right] + \Delta s_{i} {\varvec{t}}_{i} $$
(A8)

where \(\Delta s_{i}\) represents the arc length of \(i - {\text{th}}\) solution point. Note that a step control decides whether a continuation algorithm works well, affecting the continuation’s correctness and robustness. In Wu’s study [40], an improved adaptive method for arc-length control is proposed and adopted in our work.

1.3 Orthogonal corrections

In general, the predicted point is not a solution of Eq. (15), and the next solution point can be then computed by using the New-Raphson algorithm iteratively in the direction orthogonal to tangent \({\varvec{t}}\) until the criteria \(\left\| {{\varvec{R}}_{i + 1}^{k} } \right\| \le \varepsilon\) satisfies, where \(\varepsilon\) is a user-defined convergence criterion.

$$ \begin{gathered} \left[ {\begin{array}{*{20}c} {\left. {{\varvec{R}}_{{\varvec{X}}} } \right|_{{({\varvec{X}}_{i + 1}^{k} ;\omega_{i + 1}^{k} )}} } & {\left. {{\varvec{R}}_{\omega } } \right|_{{(\;{\varvec{X}}_{i + 1}^{k} ;\omega_{i + 1}^{k} )}} } \\ {{\varvec{t}}_{i}^{T} } & {} \\ \end{array} } \right]\left[ {\begin{array}{*{20}c} {\delta {\varvec{X}}_{i + 1}^{k + 1} } \\ {\delta \omega_{i + 1}^{k + 1} } \\ \end{array} } \right] = \left[ {\begin{array}{*{20}c} { - {\varvec{R}}_{i + 1}^{k} } \\ 0 \\ \end{array} } \right] \hfill \\ \left[ {\begin{array}{*{20}c} {{\varvec{X}}_{i + 1}^{k + 1} } \\ {\omega_{i + 1}^{k + 1} } \\ \end{array} } \right] = \left[ {\begin{array}{*{20}c} {{\varvec{X}}_{i + 1}^{k} } \\ {\omega_{i + 1}^{k} } \\ \end{array} } \right] + \left[ {\begin{array}{*{20}c} {\delta {\varvec{X}}_{i + 1}^{k + 1} } \\ {\omega_{i + 1}^{k + 1} } \\ \end{array} } \right]{\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} k = 0,{\kern 1pt} {\kern 1pt} {\kern 1pt} 1,{\kern 1pt} {\kern 1pt} {\kern 1pt} 2,{\kern 1pt} {\kern 1pt} {\kern 1pt} \ldots . \hfill \\ \end{gathered} $$
(A9)

1.4 Stability and bifurcation analysis with Hill’s method

Hill’s method is well-known for determining a solution branch’s stabilities and bifurcation behaviors in the frequency domain. Suppose that the solution \({\varvec{x}}^{*} (t)\) in the time domain is perturbed with a disturbance term \(e^{\Lambda t} {\varvec{s}}(t)\), where \(e^{\Lambda t}\) is the decay term and \({\varvec{s}}(t)\) the periodic term:

$$ {\varvec{x}}(t) = {\varvec{x}}^{*} (t) + e^{\Lambda t} {\varvec{s}}(t) $$
(A10)

By substituting Eq. (A10) into Eq. (15) and applying the Galerkin procedure, the following quadratic eigenvalue problem is obtained:

$$ (\Delta_{2} \Lambda^{2} + \Delta_{2} \Lambda + {\varvec{R}}_{{\varvec{X}}} )e^{\Lambda t} {\varvec{S}} = {\mathbf{0}} $$
(A11)

where \({\varvec{S}}\) are the complex eigenvectors and

$$ \begin{gathered} \Delta_{1} = 2\omega \nabla \otimes {\varvec{M}} + {\varvec{I}}_{2H + 1} \otimes {\varvec{C}} \hfill \\ \Delta_{2} = {\varvec{I}}_{2H + 1} \otimes {\varvec{M}} \hfill \\ \end{gathered} $$
(A12)

Then, the Hill’s eigenvalues can be obtained from

$$ ({\varvec{H}} - \Lambda {\varvec{I}}_{2L} ){\varvec{Q}} = {\mathbf{0}} $$
(A13)

where \({\varvec{H}}\) is consistent with the Hill matrix in the state space, which can be expressed as

$$ {\varvec{H}} = \left[ {\begin{array}{*{20}c} 0 & {{\varvec{I}}_{L} } \\ { - \Delta_{2}^{ - 1} {\varvec{R}}_{{\varvec{X}}} } & { - \Delta_{2}^{ - 1} \Delta_{1} } \\ \end{array} } \right] $$
(A14)

Note that \(2L = 2n \times (2H + 1)\) eigenvalues can be obtained from Eq. (41), but only \(2n\) of them have the physical meaning and correspond to the Floquet exponents, denoted by \(\lambda_{i} ,{\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} i = 1,{\kern 1pt} {\kern 1pt} {\kern 1pt} 2, \ldots ,{\kern 1pt} {\kern 1pt} {\kern 1pt} 2n\) in the following. Selecting the \(2n\) eigenvalues with the smallest imaginary part in the modulus is widely used to approximate Floquet exponents [43]. Once the approximated Floquet exponents are filtered, the stability and bifurcation behaviors of a periodic solution can thus be analyzed by observing their evolvement.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, Q., Liu, C., Su, Y. et al. Influences of inherent geometrical nonlinearity of high-static-low-dynamic-stiffness resonator on flexural wave attenuation performance of metamaterial beam. Nonlinear Dyn 112, 7831–7845 (2024). https://doi.org/10.1007/s11071-024-09519-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-024-09519-6

Keywords

Navigation