Skip to main content
Log in

Reduced-order modeling of geometrically nonlinear structures. Part I: A low-order elimination technique

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

Abstract

A general methodology for reduced-order modeling of geometrically nonlinear structures is proposed. This approach is built upon equivalent elimination of low-order nonlinear terms by employing a key concept termed ‘passive patterns’, defined to be essential dynamic features of nonlinear structures, produced by, namely slaved to, the active mode via low-order nonlinear effects. Thus, besides the active/dominant structural mode, passive patterns are regarded as secondary ‘energy-containing’ features. Their asymptotic construction procedure is explicitly presented in a weakly nonlinear framework. It is pointed out that both active mode and passive patterns are non-trivial dynamic features of nonlinear structures, and reduction errors of routine Galerkin truncation method are due to incomplete characterization of these passive patterns. Through elimination of their low-order sources (like nonlinearity), the proposed technique fully captures the passive patterns in an equivalent manner and thus leads to refined reduced-order models (ROMs) of the nonlinear structures, while its connection with other existing nonlinear reduction methods is detailed in Part II (Guo and Rega in Nonlinear Dyn, 2023) via an expanded theoretical correspondence.

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 availability

All data in this study will be made available on reasonable request.

References

  1. Guo, T.D., Rega, G.: Reduced order modelling of geometrically nonlinear structures. Part 2: Correspondence and unified perspectives on different reduction techniques. Nonlinear Dyn. (2023) https://doi.org/10.1007/s11071-023-08745-8

    Article  Google Scholar 

  2. Rega, G.: Nonlinear dynamics in mechanics and engineering: 40 years of developments and Ali H. Nayfeh’s legacy, Nonlinear Dyn. 99, 11–34 (2020)

    MATH  Google Scholar 

  3. Steindl, A., Troger, H.: Methods for dimension reduction and their application in nonlinear dynamics. Int. J. Solids Struct. 38, 2131–2147 (2001)

    MATH  Google Scholar 

  4. Rega, G., Troger, H.: Dimension reduction of dynamical systems: methods, models, applications. Nonlinear Dyn. 41, 1–15 (2005)

    MathSciNet  MATH  Google Scholar 

  5. C. Touzé, A. Vizzaccaro, O. Thomas, Model order reduction methods for geometrically nonlinear structures: a review of nonlinear techniques. Nonlinear Dyn. 105, 1–50 (2021).

    Google Scholar 

  6. Lu, K., Jin, Y., Chen, Y., Yang, Y., Hou, L., Zhang, Z., Li, Z., Fu, C.: Review for order reduction based on proper orthogonal decomposition and outlooks of applications in mechanical systems. Mech. Syst. Signal Process. 123, 264–297 (2019)

    Google Scholar 

  7. C.E. Mazzilli, P.B. Gonçalves, G.R. Franzini, Reduced-order modelling based on non-linear modes, Int. J. Mech. Sci. 106915 (2021)

  8. Shaw, S.W., Pierre, C.: Modal analysis-based reduced-order models for nonlinear structures: an invariant manifold approach, The. Shock and Vibration Digest 31, 3–16 (1999)

    Google Scholar 

  9. Nayfeh, A.H.: Reduced-order models of weakly nonlinear spatially continuous systems. Nonlinear Dyn. 16, 105–125 (1998)

    MathSciNet  MATH  Google Scholar 

  10. Touzé, C., Thomas, O., Chaigne, A.: Hardening/softening behaviour in non-linear oscillations of structural systems using non-linear normal modes. J. Sound. Vib. 273, 77–101 (2004)

    Google Scholar 

  11. Amabili, M.: Reduced-order models for nonlinear vibrations, based on natural modes: the case of the circular cylindrical shell. Philos. Trans. R. Soc. A: Math. Phys. Eng. Sci. 371, 20120474 (2013)

    MathSciNet  MATH  Google Scholar 

  12. Haller, G., Ponsioen, S.: Nonlinear normal modes and spectral submanifolds: existence, uniqueness and use in model reduction. Nonlinear Dyn. 86, 1–42 (2016)

    MathSciNet  MATH  Google Scholar 

  13. Haller, G., Ponsioen, S.: Exact model reduction by a slow–fast decomposition of nonlinear mechanical systems. Nonlinear Dyn. 90, 617–647 (2017)

    MathSciNet  MATH  Google Scholar 

  14. Jain, S., Tiso, P., Haller, G.: Exact nonlinear model reduction for a von Kármán beam: Slow-fast decomposition and spectral submanifolds. J. Sound. Vib. 423, 195–211 (2018)

    Google Scholar 

  15. P. Kokotović, H.K. Khalil, J. O'Reilly, Singular perturbation methods in control: analysis and design. SIAM (1999).

  16. Nicolaidou, E., Hill, T.L., Neild, S.A.: Indirect reduced-order modelling: using nonlinear manifolds to conserve kinetic energy. Proceedings of the Royal Society A 476, 20200589 (2020)

    MathSciNet  MATH  Google Scholar 

  17. Vizzaccaro, A., Shen, Y., Salles, L., Blahoš, J., Touzé, C.: Direct computation of nonlinear mapping via normal form for reduced-order models of finite element nonlinear structures. Comput. Methods Appl. Mech. Eng. 384, 113957 (2021)

    MathSciNet  MATH  Google Scholar 

  18. Jain, S., Haller, G.: How to compute invariant manifolds and their reduced dynamics in high-dimensional finite-element models? Nonlinear Dyn. 107, 1417–1450 (2022)

    Google Scholar 

  19. Vizzaccaro, A., Opreni, A., Salles, L., Frangi, A., Touzé, C.: High order direct parametrisation of invariant manifolds for model order reduction of finite element structures: application to large amplitude vibrations and uncovering of a folding point, Nonlinear Dyn. 1–47 (2022)

  20. Kerschen, G., Golinval, J.C., Vakakis, A.F., Bergman, L.A.: The method of proper orthogonal decomposition for dynamical characterization and order reduction of mechanical systems: an overview. Nonlinear Dyn. 41, 147–169 (2005)

    MathSciNet  MATH  Google Scholar 

  21. Amabili, M., Sarkar, A., Païdoussis, M.: Chaotic vibrations of circular cylindrical shells: Galerkin versus reduced-order models via the proper orthogonal decomposition method. J. Sound. Vib. 290, 736–762 (2006)

    Google Scholar 

  22. Cenedese, M., Axås, J., Yang, H., Eriten, M., Haller, G.: Data-driven nonlinear model reduction to spectral submanifolds in mechanical systems. Philos. Trans. R. Soc. A 380, 20210194 (2022)

    MathSciNet  Google Scholar 

  23. Lacarbonara, W.: Direct treatment and discretizations of non-linear spatially continuous systems. J. Sound. Vib. 221, 849–866 (1999)

    MathSciNet  MATH  Google Scholar 

  24. Nayfeh, A.H., Lacarbonara, W.: On the discretization of distributed-parameter systems with quadratic and cubic nonlinearities. Nonlinear Dyn. 13, 203–220 (1997)

    MathSciNet  MATH  Google Scholar 

  25. Shaw, S.W., Pierre, C.: Normal modes for non-linear vibratory systems. J. Sound. Vib. 164, 85–124 (1993)

    MATH  Google Scholar 

  26. Nayfeh, A.H.: On direct methods for constructing nonlinear normal modes of continuous systems. J. Vib. Control 1, 389–430 (1995)

    MathSciNet  MATH  Google Scholar 

  27. Vakakis, A.F., Manevitch, L.I., Mikhlin, Y.V.: Normal Modes and Localization in Nonlinear Systems. Springer, Netherlands (2001)

    MATH  Google Scholar 

  28. Touzé, C., Amabili, M.: Nonlinear normal modes for damped geometrically nonlinear systems: application to reduced-order modelling of harmonically forced structures. J. Sound. Vib. 298, 958–981 (2006)

    Google Scholar 

  29. Kerschen, G., Peeters, M., Golinval, J.-C., Vakakis, A.F.: Nonlinear normal modes, Part I: A useful framework for the structural dynamicist. Mech. Syst. Signal Process. 23, 170–194 (2009)

    Google Scholar 

  30. Ponsioen, S., Pedergnana, T., Haller, G.: Automated computation of autonomous spectral submanifolds for nonlinear modal analysis. J. Sound. Vib. 420, 269–295 (2018)

    Google Scholar 

  31. Breunung, T., Haller, G.: Explicit backbone curves from spectral submanifolds of forced-damped nonlinear mechanical systems. Proc. R. Soc. A: Math. Phys. Eng. Sci. 474, 20180083 (2018)

    MathSciNet  MATH  Google Scholar 

  32. S. Jain, T. Thurnher, M. Li, G. Haller, SSMTool 2.3: Computation of invariant manifolds & their reduced dynamics in high-dimensional mechanics problems. https://github.com/haller-group/SSMTool-2.3. Accessed 2023-5-20.

  33. Lacarbonara, W., Camillacci, R.: Nonlinear normal modes of structural systems via asymptotic approach. Int. J. Solids Struct. 41, 5565–5594 (2004)

    MATH  Google Scholar 

  34. Lacarbonara, W., Rega, G., Nayfeh, A.H.: Resonant non-linear normal modes. Part I: Analytical treatment for structural one-dimensional systems. Int. J. Nonlinear Mech. 38, 851–872 (2003)

    MATH  Google Scholar 

  35. Carr, J.: Applications of Centre Manifold Theory. Springer, Berlin (1991)

    MATH  Google Scholar 

  36. Lenci, S., Rega, G.: Dimension reduction of homoclinic orbits of buckled beams via the non-linear normal modes technique. Int. J. Non Linear Mech. 42, 515–528 (2007)

    MATH  Google Scholar 

  37. Touzé, C.: Normal form theory and nonlinear normal modes: theoretical settings and applications. In: Kerschen, G. (ed.) 'Modal Analysis of Nonlinear Mechanical Systems' , CISM Courses and Lectures 555. Springer, pp. 75–160 (2014)

  38. Nayfeh, A.H.: The Method of Normal Forms. Wiley, New York (2011)

  39. Murdock, J.: Normal Forms and Unfoldings for Local Dynamical Systems. Springer, Berlin (2006)

  40. Guckenheimer, J., Holmes, P.: Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Springer, Berlin (2013)

  41. Guo, T.D., Rega, G.: Direct and discretized perturbations revisited: a new error source interpretation, with application to moving boundary problem. Eur. J. Mech. A/Solids 81, 103936 (2020)

    MathSciNet  MATH  Google Scholar 

  42. Guo, T.D., Rega, G., Kang, H.J.: General perturbation correction: full-decomposition and physics-based elimination of non-secular terms. Int. J. Mech. Sci. 216, 106966 (2022)

    Google Scholar 

  43. Holmes, P., Lumley, J.L., Berkooz, G., Rowley, C.W.: Turbulence, Coherent Structures, Dynamical Systems and Symmetry, 2nd edn. Cambridge University Press, Cambridge (2012)

  44. Touzé, C., Thomas, O., Huberdeau, A.: Asymptotic non-linear normal modes for large-amplitude vibrations of continuous structures. Comput. Struct. 82, 2671–2682 (2004)

    Google Scholar 

  45. Guo, T.D., Kang, H.J., Wang, L.H., Zhao, Y.Y.: An inclined cable excited by a non-ideal massive moving deck: theoretical formulation and nonlinear responses. Nonlinear Dyn. 95, 749–767 (2019)

    Google Scholar 

  46. Lacarbonara, W., Yabuno, H.: Refined models of elastic beams undergoing large in-plane motions: Theory and experiment. Int. J. Solids Struct. 43, 5066–5084 (2006)

    MATH  Google Scholar 

  47. Kerschen, G. (Ed.), Modal Analysis of Nonlinear Mechanical Systems, CISM Courses and Lectures 555. Springer (2014)

  48. Cirillo, G.I., Mauroy, A., Renson, L., Kerschen, G., Sepulchre, R.: A spectral characterization of nonlinear normal modes. J. Sound. Vib. 377, 284–301 (2016)

    Google Scholar 

  49. Mezić, I.: Spectral properties of dynamical systems, model reduction and decompositions. Nonlinear Dyn. 41, 309–325 (2005)

    MathSciNet  MATH  Google Scholar 

  50. Bagheri, S.: Koopman-mode decomposition of the cylinder wake. J. Fluid Mech. 726, 596–623 (2013)

    MathSciNet  MATH  Google Scholar 

  51. Noack, B.R.: From snapshots to modal expansions–bridging low residuals and pure frequencies. J. Fluid Mech. 802, 1–4 (2016)

    MathSciNet  MATH  Google Scholar 

  52. Taira, K., Brunton, S.L., Dawson, S.T., Rowley, C.W., Colonius, T., McKeon, B.J., Schmidt, O.T., Gordeyev, S., Theofilis, V., Ukeiley, L.S.: Modal analysis of fluid flows: an overview. AIAA J. 55, 4013–4041 (2017)

    Google Scholar 

  53. Rowley, C.W., Mezić, I., Bagheri, S., Schlatter, P., Henningson, D.: Spectral analysis of nonlinear flows. J. Fluid Mech. 641, 115–127 (2009)

    MathSciNet  MATH  Google Scholar 

  54. Chow, S.-N., Li, C., Wang, D.: Normal Forms and Bifurcation of Planar Vector Fields. Cambridge University Press, Cambridge (1994)

  55. Nayfeh, A., Nayfeh, S.: On nonlinear modes of continuous systems. J. Vib. Acoust. 116, 129–136 (1994)

    Google Scholar 

  56. Neild, S.A., Wagg, D.: Applying the method of normal forms to second-order nonlinear vibration problems. Proc. R Soc. A 467, 1141–1163 (2011)

    MathSciNet  MATH  Google Scholar 

  57. Elliott, A., Cammarano, A., Neild, S., Hill, T., Wagg, D.: Comparing the direct normal form and multiple scales methods through frequency detuning. Nonlinear Dyn. 94, 2919–2935 (2018)

    MATH  Google Scholar 

Download references

Acknowledgements

This study was initiated during the first author’s stay at Sapienza University of Rome as a visiting professor, when T.D Guo appreciates helpful discussions with Prof. Walter Lacarbonara and Prof. Francesco Romeo in Sapienza. Further, the authors appreciate constructive comments and suggestions of Prof. Cyril Touzé and also anonymous reviewers. This study is supported by National Science Foundation of China under Grant Nos. 12372007, 11872176, 11972151 and also by Guangxi Science & Technology Base and Talent Project under grant no. 2020AC19209.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tieding Guo.

Ethics declarations

Conflict of interest

The authors have no conflicts of interest relevant to the current article.

Additional information

Publisher's Note

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

Appendices

Appendix A

For the active modal coordinate \(p_{m} \left( t \right) = A_{m} \left( t \right)e^{{{\text{i}}\,\omega_{m} \,t}} + cc.\), we have

$$ \begin{aligned} \frac{{{\text{d}}p_{m} }}{{{\text{d}}t}} & = {\text{i}}\omega_{m} A_{m} e^{{{\text{i}}\,\omega_{m} \,t}} + \frac{{{\text{d}}A_{m} }}{{{\text{d}}t}} \cdot e^{{{\text{i}}\,\omega_{m} \,t}} + cc., \\ \frac{{{\text{d}}^{2} p_{m} }}{{{\text{d}}t^{2} }} & = - \omega_{m}^{2} p_{m} + \left( {2{\text{i}}\omega_{m} \frac{{{\text{d}}A_{m} }}{{{\text{d}}t}} \cdot e^{{{\text{i}}\,\omega_{m} \,t}} + cc.} \right) = - \omega_{m}^{2} p_{m} + \varepsilon g_{\left( 1 \right)} + \varepsilon^{2} g_{\left( 2 \right)} { + } \cdots \\ \end{aligned} $$
(53)

with \(O\left( {{{{\text{d}}A_{m} } \mathord{\left/ {\vphantom {{{\text{d}}A_{m} } {{\text{d}}t}}} \right. \kern-0pt} {{\text{d}}t}}} \right) \ll 1\).

The near-harmonic (slowly modulated) pattern coordinate \(P_{{\Omega_{k} }} \left( t \right) \ll p_{m} \left( t \right)\) is denoted by \(P_{{\Omega_{k} }} \left( t \right) = G_{{\Omega_{k} }} \left( t \right)e^{{{\text{i}}\,\Omega_{k} \,t}} + cc.\), with its derivatives given by

$$ \begin{aligned} \frac{{{\text{d}}P_{{\Omega_{k} }} }}{{{\text{d}}t}} & = {\text{i}}\Omega_{k} G_{{\Omega_{k} }} e^{{{\text{i}}\,\Omega_{k} \,t}} + \frac{{{\text{d}}G_{{\Omega_{k} }} }}{{{\text{d}}t}} \cdot e^{{{\text{i}}\,\Omega_{k} \,t}} + cc.,\quad \\ \frac{{{\text{d}}^{2} P_{{\Omega_{k} }} }}{{{\text{d}}t^{2} }} & = - \Omega_{k}^{2} P_{{\Omega_{k} }} + \left( {2{\text{i}}\Omega_{k} \frac{{{\text{d}}G_{{\Omega_{k} }} }}{{{\text{d}}t}} \cdot e^{{{\text{i}}\,\Omega_{k} \,t}} + cc.} \right) = - \Omega_{k}^{2} P_{{\Omega_{k} }} + \varepsilon h_{{\Omega_{k} \left( 1 \right)}} + \varepsilon^{2} h_{{\Omega_{k} \left( 2 \right)}} { + } \cdots \\ \end{aligned} $$
(54)

where \(O\left( {{{{\text{d}}G_{{\Omega_{k} }} } \mathord{\left/ {\vphantom {{{\text{d}}G_{{\Omega_{k} }} } {{\text{d}}t}}} \right. \kern-0pt} {{\text{d}}t}}} \right) \ll 1\) as \(P_{{\Omega_{k} }} \left( t \right)\) is slowly modulated due to nonlinear effects. Note that, although this slow modulation effect is neglected in the current work, it is important for developing more general high-order version of elimination method and will thus be taken into proper consideration in subsequent works.

Appendix B

For the second-order differentiation operator acting on \(\hat{H}_{2}^{2}\), i.e., \({{{\hat{\text{d}}}^{2} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}^{2} } {{\hat{\text{d}}}t^{2} }}} \right. \kern-0pt} {{\hat{\text{d}}}t^{2} }}:\;\;\hat{H}_{2}^{2} \to \hat{H}_{2}^{2}\), which is spanned by two monomial bases \(\{ e_{1} = p_{m}^{2} ,\;\;e_{2} = \dot{p}_{m}^{2} \}\), its matrix representation \(D^{(2)}\) is

$$ \frac{{{\hat{\text{d}}}^{2} }}{{{\hat{\text{d}}}t^{2} }}\left[ {e_{1} ,\;e_{2} } \right] = \left[ {e_{1} ,\;e_{2} } \right]\;\underbrace {{\left[ {\begin{array}{*{20}c} { - 2\omega_{m}^{2} } & {2\omega_{m}^{4} } \\ 2 & { - 2\omega_{m}^{2} } \\ \end{array} } \right]}}_{{D^{(2)} }} $$
(55)

if noting

$$ \begin{aligned} {{{\hat{\text{d}}}e_{1} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}e_{1} } {{\hat{\text{d}}}t}}} \right. \kern-0pt} {{\hat{\text{d}}}t}} & = {{{\text{d}}p_{m}^{2} } \mathord{\left/ {\vphantom {{{\text{d}}p_{m}^{2} } {{\text{d}}t}}} \right. \kern-0pt} {{\text{d}}t}} = 2p_{m} \dot{p}_{m} ,\quad {{{\hat{\text{d}}}^{2} e_{1} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}^{2} e_{1} } {{\hat{\text{d}}}t^{2} }}} \right. \kern-0pt} {{\hat{\text{d}}}t^{2} }} = 2\dot{p}_{m}^{2} - 2\omega_{m}^{2} p_{m}^{2} \\ {{{\hat{\text{d}}}e_{2} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}e_{2} } {{\hat{\text{d}}}t}}} \right. \kern-0pt} {{\hat{\text{d}}}t}} & = {{{\text{d}}\dot{p}_{m}^{2} } \mathord{\left/ {\vphantom {{{\text{d}}\dot{p}_{m}^{2} } {{\text{d}}t}}} \right. \kern-0pt} {{\text{d}}t}} = - 2\omega_{m}^{2} p_{m} \dot{p}_{m}^{{}} ,\quad {{{\hat{\text{d}}}^{2} e_{2} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}^{2} e_{2} } {{\hat{\text{d}}}t^{2} }}} \right. \kern-0pt} {{\hat{\text{d}}}t^{2} }} = 2\omega_{m}^{4} p_{m}^{2} - 2\omega_{m}^{2} \dot{p}_{m}^{2} \\ \end{aligned} $$
(56)

where \({{{\text{d}}^{2} p_{m} } \mathord{\left/ {\vphantom {{{\text{d}}^{2} p_{m} } {{\text{d}}t^{2} }}} \right. \kern-0pt} {{\text{d}}t^{2} }} \approx - \omega_{m}^{2} p_{m}\) and \({{{\hat{\text{d}}}} \mathord{\left/ {\vphantom {{{\hat{\text{d}}}} {{\hat{\text{d}}}t{ = }{\partial \mathord{\left/ {\vphantom {\partial {\partial p_{m} \cdot \dot{p}_{m} }}} \right. \kern-0pt} {\partial p_{m} \cdot \dot{p}_{m} }}}}} \right. \kern-0pt} {{\hat{\text{d}}}t{ = }{\partial \mathord{\left/ {\vphantom {\partial {\partial p_{m} \cdot \dot{p}_{m} }}} \right. \kern-0pt} {\partial p_{m} \cdot \dot{p}_{m} }}}}{ + }{\partial \mathord{\left/ {\vphantom {\partial {\partial \dot{p}_{m} }}} \right. \kern-0pt} {\partial \dot{p}_{m} }} \cdot \left( { - \omega_{m}^{2} p_{m} } \right)\) have been used.

For the second-order differentiation operator acting on \(\hat{H}_{2}^{3}\), i.e., \({{{\hat{\text{d}}}^{2} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}^{2} } {{\hat{\text{d}}}t^{2} }}} \right. \kern-0pt} {{\hat{\text{d}}}t^{2} }}:\;\;\hat{H}_{2}^{3} \to \hat{H}_{2}^{3}\), which is spanned by two monomial bases \(\{ e_{1} = p_{m}^{3} ,\;\;e_{2} = \dot{p}_{m}^{2} p_{m} \}\), its matrix representation \(D^{(3)}\) is

$$ \frac{{{\hat{\text{d}}}^{2} }}{{{\hat{\text{d}}}t^{2} }}\left[ {e_{1} ,\;e_{2} } \right] = \left[ {e_{1} ,\;e_{2} } \right]\;\underbrace {{\left[ {\begin{array}{*{20}c} { - 3\omega_{m}^{2} } & {2\omega_{m}^{4} } \\ 6 & { - 7\omega_{m}^{2} } \\ \end{array} } \right]}}_{{D^{(3)} }} $$
(57)

if noting

$$ \begin{aligned} {{{\hat{\text{d}}}e_{1} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}e_{1} } {{\hat{\text{d}}}t}}} \right. \kern-0pt} {{\hat{\text{d}}}t}} & = {{{\text{d}}p_{m}^{3} } \mathord{\left/ {\vphantom {{{\text{d}}p_{m}^{3} } {{\text{d}}t}}} \right. \kern-0pt} {{\text{d}}t}} = 3p_{m}^{2} \dot{p}_{m} ,\quad {{{\hat{\text{d}}}^{2} e_{1} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}^{2} e_{1} } {{\hat{\text{d}}}t^{2} }}} \right. \kern-0pt} {{\hat{\text{d}}}t^{2} }} = 6p_{m} \dot{p}_{m}^{2} - 3\omega_{m}^{2} p_{m}^{3} \\ {{{\hat{\text{d}}}e_{2} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}e_{2} } {{\hat{\text{d}}}t}}} \right. \kern-0pt} {{\hat{\text{d}}}t}} & = {{{\text{d}}\dot{p}_{m}^{2} p_{m} } \mathord{\left/ {\vphantom {{{\text{d}}\dot{p}_{m}^{2} p_{m} } {{\text{d}}t}}} \right. \kern-0pt} {{\text{d}}t}} = - 2\omega_{m}^{2} p_{m}^{2} \dot{p}_{m} + \dot{p}_{m}^{3} ,\\&{{{\hat{\text{d}}}^{2} e_{2} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}^{2} e_{2} } {{\hat{\text{d}}}t^{2} }}} \right. \kern-0pt} {{\hat{\text{d}}}t^{2} }} = 2\omega_{m}^{4} p_{m}^{3} - 7\omega_{m}^{2} p_{m} \dot{p}_{m}^{2} \\ \end{aligned} $$
(58)

where \({{{\text{d}}^{2} p_{m} } \mathord{\left/ {\vphantom {{{\text{d}}^{2} p_{m} } {{\text{d}}t^{2} }}} \right. \kern-0pt} {{\text{d}}t^{2} }} \approx - \omega_{m}^{2} p_{m}\) and \({{{\hat{\text{d}}}} \mathord{\left/ {\vphantom {{{\hat{\text{d}}}} {{\hat{\text{d}}}t{ = }{\partial \mathord{\left/ {\vphantom {\partial {\partial p_{m} \cdot \dot{p}_{m} }}} \right. \kern-0pt} {\partial p_{m} \cdot \dot{p}_{m} }}}}} \right. \kern-0pt} {{\hat{\text{d}}}t{ = }{\partial \mathord{\left/ {\vphantom {\partial {\partial p_{m} \cdot \dot{p}_{m} }}} \right. \kern-0pt} {\partial p_{m} \cdot \dot{p}_{m} }}}}{ + }{\partial \mathord{\left/ {\vphantom {\partial {\partial \dot{p}_{m} }}} \right. \kern-0pt} {\partial \dot{p}_{m} }} \cdot \left( { - \omega_{m}^{2} p_{m} } \right)\) have been used.

One more remark is regarding the two subspaces \(\hat{H}_{2}^{2} \subset H_{2}^{2} ,\;\;\hat{H}_{2}^{3} \subset H_{2}^{3}\). Following dynamical system literature [54], a first-order state space formulation consisting of displacement and velocity fields can be used, e.g., \(X = \left[ {w,\;\dot{w}} \right]^{T}\), and one has

$$ \begin{aligned} H_{2}^{2} & = {\text{span}}\left\{ {p_{m}^{2} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right], \;\dot{p}_{m}^{2} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right],\; p_{m} \dot{p}_{m} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right],\; p_{m}^{2} \left[ {\begin{array}{*{20}c} 0 \\ 1 \\ \end{array} } \right],\; \dot{p}_{m}^{2} \left[ {\begin{array}{*{20}c} 0 \\ 1 \\ \end{array} } \right],p_{m} \dot{p}_{m} \left[ {\begin{array}{*{20}c} 0 \\ 1 \\ \end{array} } \right]} \right\}; \\ H_{2}^{3} & = {\text{span}}\left\{ {p_{m}^{3} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right],\; p_{m} \dot{p}_{m}^{2} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right],\;\dot{p}_{m}^{3} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right],\; p_{m}^{2} \dot{p}_{m} \left[ {\begin{array}{*{20}c} 1 \\ 0 \\ \end{array} } \right],\; p_{m}^{3} \left[ {\begin{array}{*{20}c} 0 \\ 1 \\ \end{array} } \right],\;p_{m} \dot{p}_{m}^{2} \left[ {\begin{array}{*{20}c} 0 \\ 1 \\ \end{array} } \right],\; \dot{p}_{m}^{3} \left[ {\begin{array}{*{20}c} 0 \\ 1 \\ \end{array} } \right],\;p_{m}^{2} \dot{p}_{m} \left[ {\begin{array}{*{20}c} 0 \\ 1 \\ \end{array} } \right]} \right\} \\ \end{aligned} $$
(59)

By contrast, for the current second-order oscillator formulation, only displacement field occurs in the homological equations, say, Eqs. (18) or (24), meaning that only the first three (four) monomial bases in Eq. (59) are required for describing nonlinear displacement in quadratic (cubic) problems, respectively,

$$ \begin{aligned} \overline{H}_{2}^{2} = {\text{span}}\{ e_{1} = p_{m}^{2} ,\;\;e_{2} = \dot{p}_{m}^{2} \;\} \oplus {\text{span}}\{ e_{3} = p_{m} \dot{p}_{m} \} ; \hfill \\ \overline{H}_{2}^{3} = {\text{span}}\{ e_{1} = p_{m}^{3} ,\;\;e_{2} = p_{m} \dot{p}_{m}^{2} \} \oplus {\text{span}}\{ e_{3} = \dot{p}_{m}^{3} ,\;\;\;e_{4} = p_{m}^{2} \dot{p}_{m} \} \hfill \\ \end{aligned} $$
(60)

Both are denoted as direct sum of two invariant subspaces, with respect to the operator \({{{\hat{\text{d}}}^{2} } \mathord{\left/ {\vphantom {{{\hat{\text{d}}}^{2} } {{\hat{\text{d}}}t^{2} }}} \right. \kern-0pt} {{\hat{\text{d}}}t^{2} }}\). This can be directly verified by similar procedures used in Eqs. (56) or (58).

More explicitly, corresponding to quadratic nonlinearity \(p_{m}^{2}\), only the invariant subspace \(\hat{H}_{2}^{2} = {\text{span}}\{ e_{1} = p_{m}^{2} ,\;\;e_{2} = \dot{p}_{m}^{2} \}\) is demanded for constructing passive patterns. Similarly, corresponding to cubic nonlinearity \(p_{m}^{3}\) and/or \(\dot{p}_{m}^{2} p_{m}\), only the invariant subspace \(\hat{H}_{2}^{3} = {\text{span}}\{ e_{1} = p_{m}^{3} ,\;\;e_{2} = \dot{p}_{m}^{2} p_{m} \}\) is required for passive patterns. This explains why subspaces \(\hat{H}_{2}^{2} ,\;\;\hat{H}_{2}^{3}\) are employed in Sect. 2.3. However, for cubic nonlinearity due to damping, for example, \(\dot{p}_{m}^{3}\) or \(p_{m}^{2} \dot{p}_{m}\) [55], another invariant subspace \({\text{span}}\{ e_{3} = \dot{p}_{m}^{3} ,\;\;\;e_{4} = p_{m}^{2} \dot{p}_{m} \}\) should be employed to construct passive patterns. The current framework using invariant subspaces of real-valued polynomials is different from normal form analysis using complex-valued notations, built around either linear eigen-frequency [26] or nonlinear response frequency [56, 57].

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

Guo, T., Rega, G. Reduced-order modeling of geometrically nonlinear structures. Part I: A low-order elimination technique. Nonlinear Dyn 111, 19629–19654 (2023). https://doi.org/10.1007/s11071-023-08822-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-023-08822-y

Keywords

Navigation