Skip to main content
Log in

High-order direct parametrisation of invariant manifolds for model order reduction of finite element structures: application to generic forcing terms and parametrically excited systems

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

Abstract

The direct parametrisation method for invariant manifolds is used for model order reduction of forced-damped mechanical structures subjected to geometric nonlinearities. Nonlinear mappings are introduced, allowing one to pass from the degrees of freedom of the finite-element model to the normal coordinates. Arbitrary orders of expansions are considered for the unknown mappings and the reduced dynamics, which are then solved sequentially through the homological equations for both autonomous and time-dependent terms. It is emphasised that the two problems share a similar structure, which can be used for an efficient implementation of the non-autonomous added terms. Special emphasis is also put on the new resonance conditions arising due to the presence of the external forcing frequencies, which allow predicting phenomena such as parametric excitation and isolas formation. The method is then applied to structures of academic and industrial interest. First, the large amplitude vibrations of a forced-damped cantilever beam are studied. This example highlights that high-order non-autonomous terms are compulsory to correctly estimate the maximum vibration amplitude experienced by the structure. The birth of isolated solutions is also illustrated on this example. The cantilever is then used to show how quadratic coupling creates conditions for the excitation of the parametric instability, and that this feature is correctly embedded in the reduction process. A shallow arch excited with multi-modal forcing is then studied to detail different forcing effects. Finally, the approach is validated on a structure of industrial relevance, i.e. a comb-driven micro-electro-mechanical resonator. The accuracy and computational performance reported suggest that the proposed methodology can accurately predict the nonlinear dynamic response of a large class of nonlinear vibratory systems.

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

Similar content being viewed by others

Data availability

The codes developed for the presented analyses are available upon request to the authors.

References

  1. Zega, V., Gattere, G., Koppaka, S., Alter, A., Vukasin, G., Frangi, A., Kenny, T.W.: Numerical modelling of non-linearities in MEMS resonators. J. Microelectromech. Syst. 29(6), 1443–1454 (2020)

    Article  Google Scholar 

  2. Hajjaj, A.Z., Hafiz, M.A., Younis, M.I.: Mode coupling and nonlinear resonances of MEMS arch resonators for bandpass filters. Sci. Rep. 7(1), 1–7 (2017)

    Article  Google Scholar 

  3. Rhoads, J.F., Shaw, S.W., Turner, K.L., Baskaran, R.: Tunable microelectromechanical filters that exploit parametric resonance. J. Vib. Acoust. 127(5), 423–430 (2005)

    Article  Google Scholar 

  4. Mignolet, M.P., Przekop, A., Rizzi, S.A., Spottswood, S.M.: A review of indirect/non-intrusive reduced order modeling of nonlinear geometric structures. J. Sound Vib. 332, 2437–2460 (2013)

    Article  Google Scholar 

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

  6. Amabili, M., Sarkar, A., Païdoussis, M.P.: Reduced-order models for nonlinear vibrations of cylindrical shells via the proper orthogonal decomposition method. J. Fluids Struct. 18(2), 227–250 (2003)

    Article  Google Scholar 

  7. Amabili, M., Touzé, C.: Reduced-order models for non-linear vibrations of fluid-filled circular cylindrical shells: comparison of POD and asymptotic non-linear normal modes methods. J. Fluids Struct. 23(6), 885–903 (2007)

    Article  Google Scholar 

  8. Vizzaccaro, A., Givois, A., Longobardi, P., Shen, Y., Deü, J.-F., Salles, L., Touzé, C., Thomas, O.: Non-intrusive reduced order modelling for the dynamics of geometrically nonlinear flat structures using three-dimensional finite elements. Comput. Mech. 66, 1293–1319 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  9. Kerschen, G., Golinval, J.-C., Vakakis, A.F., Bergmann, L.A.: The method of proper orthogonal decomposition for dynamical characterization and order reduction of mechanical systems: an overview. Nonlinear Dyn. 41(1), 147–169 (2005)

    Article  MathSciNet  Google Scholar 

  10. Shaw, S.W., Pierre, C.: Non-linear normal modes and invariant manifolds. J. Sound Vib. 150(1), 170–173 (1991)

    Article  Google Scholar 

  11. Shaw, S.W., Pierre, C.: Normal modes for non-linear vibratory systems. J. Sound Vib. 164(1), 85–124 (1993)

    Article  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  13. 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(1–2), 77–101 (2004)

    Article  Google Scholar 

  14. Shaw, S.W., Pierre, C.: Normal modes of vibration for non-linear continuous systems. J. Sound Vib. 169(3), 85–124 (1994)

    Article  Google Scholar 

  15. Lyapunov, A.M.: Problème général de la stabilité du mouvement. Ann. de la Fac. des Sci. de Toulouse, Sér. 2(9), 203–474 (1907)

    Google Scholar 

  16. Kelley, A.F.: Analytic two-dimensional subcenter manifolds for systems with an integral. Pac. J. Math. 29, 335–350 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  17. 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)

    Article  Google Scholar 

  18. Haller, G., Ponsioen, S.: Nonlinear normal modes and spectral submanifolds: existence, uniqueness and use in model reduction. Nonlinear Dyn. 86(3), 1493–1534 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  19. Haller, G., Ponsioen, S.: Exact model reduction by a slow-fast decomposition of nonlinear mechanical systems. Nonlinear Dyn. 90(1), 617–647 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  20. Pesheck, E., Pierre, C., Shaw, S.: A new Galerkin-based approach for accurate non-linear normal modes through invariant manifolds. J. Sound Vib. 249(5), 971–993 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  21. Jiang, D.: Nonlinear modal analysis based on invariant manifolds. Application to rotating blade systems. PhD thesis, University of Michigan, (2004)

  22. Touzé, C., Amabili, M.: Non-linear normal modes for damped geometrically non-linear systems: application to reduced-order modeling of harmonically forced structures. J. Sound Vib. 298(4–5), 958–981 (2006)

    Article  Google Scholar 

  23. Pesheck, E., Boivin, N., Pierre, C., Shaw, S.: Nonlinear modal analysis of structural systems using multi-mode invariant manifolds. Nonlinear Dyn. 25, 183–205 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  24. Mazzilli, C.E.N., Soares, M.E.S., Baracho Neto, O.G.P.: Non-linear normal modes of a simply supported beam: continuous system and finite-element models. Comput. Struct. 82(31), 2683–2691 (2004)

    Article  Google Scholar 

  25. Cabré, X., Fontich, E., de la Llave, R.: The parameterization method for invariant manifolds. i. manifolds associated to non-resonant subspaces. Indiana Univ. Math. J. 52(2), 283–328 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  26. Cabré, X., Fontich, E., de la Llave, R.: The parameterization method for invariant manifolds. ii. regularity with respect to parameters. Indiana Univ. Math. J. 52(2), 329–360 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  27. Cabré, X., Fontich, E., de la Llave, R.: The parameterization method for invariant manifolds. iii. Overview and applications. J. Differ. Equs. 218(2), 444–515 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  28. Haro, A., Canadell, M., Figueras, J.-L., Luque, A., Mondelo, J.-M.: The parameterization method for invariant manifolds. From rigorous results to effective computations. Springer, Switzerland (2016)

    Book  MATH  Google Scholar 

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

    Article  Google Scholar 

  30. 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(2213), 20180083 (2018)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  32. 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)

    Article  MathSciNet  MATH  Google Scholar 

  33. Opreni, A., Vizzaccaro, A., Touzé, C., Frangi, A.: Model order reduction based on direct normal form: application to large finite element MEMS structures featuring internal resonance. Nonlinear Dyn. 105, 1237–1272 (2021)

    Article  Google Scholar 

  34. 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)

    Article  Google Scholar 

  35. 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. 110, 525–571 (2022)

    Article  Google Scholar 

  36. Li, M., Jain, S., Haller, G.: Nonlinear analysis of forced mechanical systems with internal resonance using spectral submanifolds–part i: periodic response and forced response curve. arXiv preprint arXiv:2106.05162, (2021)

  37. Li, M., Haller, G.: Nonlinear analysis of forced mechanical systems with internal resonance using spectral submanifolds–part ii: bifurcation and quasi-periodic response. arXiv preprint arXiv:2108.08152, (2021)

  38. Jiang, D., Pierre, C., Shaw, S.: Nonlinear normal modes for vibratory systems under harmonic excitation. J. Sound Vib. 288(4–5), 791–812 (2005)

    Article  Google Scholar 

  39. Haro, A., de la Llave, R.: A parameterization method for the computation of invariant tori and their whiskers in quasi-periodic maps: rigorous results. J. Differ. Equs. 228(2), 530–579 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  40. Haro, A., de la Llave, R.: A parameterization method for the computation of invariant tori and their whiskers in quasi-periodic maps: numerical algorithms. Discr. Contin. Dyn. Syst.-B 6(6), 1261 (2006)

    MathSciNet  MATH  Google Scholar 

  41. Haro, A., de La Llave, R.: A parameterization method for the computation of invariant tori and their whiskers in quasi-periodic maps: explorations and mechanisms for the breakdown of hyperbolicity. SIAM J. Appl. Dyn. Syst. 6(1), 142 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  42. Ponsioen, S., Pedergnana, T., Haller, G.: Analytic prediction of isolated forced response curves from spectral submanifolds. Nonlinear Dyn. 98(4), 2755–2773 (2019)

    Article  MATH  Google Scholar 

  43. Ponsioen, S., Jain, S., Haller, G.: Model reduction to spectral submanifolds and forced-response calculation in high-dimensional mechanical systems. J. Sound Vib. 488, 115640 (2020)

    Article  Google Scholar 

  44. Holzapfel, G.: Nonlinear solid mechanics. Jhon Wiley & Sons, Chichester (2000)

  45. Golla, D.F., Hughes, P.C.: Dynamics of viscoelastic structures: a time-domain, finite element formulation. J. Appl. Mech. 52(4), 897–906 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  46. Pritz, T.: Frequency power law of material damping. Appl. Acoust. 65(11), 1027–1036 (2004)

    Article  Google Scholar 

  47. Deü, J.-F., Matignon, D.: Simulation of fractionally damped mechanical systems by means of a Newmark-diffusive scheme. Comput. Math. Appl. 59(5), 1745–1753 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  48. Parret-Fréaud, A., Cotté, B., Chaigne, A.: Time-domain damping models in structural acoustics using digital filtering. Mech. Syst. Signal Process. 68–69, 587–607 (2016)

    Article  Google Scholar 

  49. Sumali, H.: Squeeze-film damping in the free molecular regime: model validation and measurement on a MEMS. J. Micromech. Microeng. 17(11), 2231 (2007)

    Article  Google Scholar 

  50. Méndez, C., Paquay, S., Klapka, I., Raskin, J.-P.: Effect of geometrical nonlinearity on MEMS thermoelastic damping. Nonlinear Anal. Real World Appl. 10(3), 1579–1588 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  51. Frangi, A., Fedeli, P., Laghi, G., Langfelder, G., Gattere, G.: Near vacuum gas damping in MEMS: numerical modeling and experimental validation. J. Microelectromech. Syst. 25(5), 890–899 (2016)

    Article  Google Scholar 

  52. Lazarus, A., Thomas, O., Deü, J.-F.: Finite element reduced order models for nonlinear vibrations of piezoelectric layered beams with applications to NEMS. Finite Elem. Anal. Des. 49, 35–51 (2012)

    Article  MathSciNet  Google Scholar 

  53. Touzé, C., Vidrascu, M., Chapelle, D.: Direct finite element computation of non-linear modal coupling coefficients for reduced-order shell models. Comput. Mech. 54(2), 567–580 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  54. Touzé, C.: Normal form theory and nonlinear normal modes: theoretical settings and applications. In Kerschen, G. editor, Modal Analysis of nonlinear Mechanical Systems, pp. 75–160, New York, NY, (2014). Springer Series CISM courses and lectures, vol. 555

  55. Doedel, E.J., Paffenroth, R., Champneys, A.R., Fairgrieve, T.F., Kuznetsov, Y.A., Oldeman, B.E., Sandstede, B., Wang, X.: Auto 2000: continuation and bifurcation software for ordinary differential equations. Technical report, Concordia University, (2002)

  56. Thomsen, J.J.: Vibrations and stability. Advanced theory, analysis and tools. Springer, Berlin, Heidelberg, (2003). 2nd edition

  57. Poincaré, H.: Les méthodes nouvelles de la mécanique céleste. Gauthiers-Villars, Paris (1892)

    MATH  Google Scholar 

  58. Guckenheimer, J., Holmes, P.: Nonlinear oscillations, dynamical systems and bifurcations of vector fields. Springer-Verlag, New-York (1983)

    Book  MATH  Google Scholar 

  59. Jézéquel, L., Lamarque, C.H.: Analysis of non-linear dynamical systems by the normal form theory. J. Sound Vib. 149(3), 429–459 (1991)

    Article  Google Scholar 

  60. Neild, S.A., Champneys, A.R., Wagg, D.J., Hill, T.L., Cammarano, A.: The use of normal forms for analysing nonlinear mechanical vibrations. Proc. R. Soc. A. 373, 20140404 (2015)

    MathSciNet  MATH  Google Scholar 

  61. Wagg, D.J.: Normal form transformations for structural dynamics: an introduction for linear and nonlinear systems. J. Struct. Dyn. 1, 138–216 (2022) https://popups.uliege.be/2684-6500/index.php?id=84

  62. Opreni, A., Boni, N., Carminati, R., Frangi, A.: Analysis of the nonlinear response of piezo-micromirrors with the harmonic balance method. Actuators 10(2), 21 (2021)

    Article  Google Scholar 

  63. Dhooge, A., Govaerts, W., Kuznetsov, Y.A.: Matcont: a matlab package for numerical bifurcation analysis of odes. ACM Trans. Math. Softw. (TOMS) 29(2), 141–164 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  64. Cochelin, B., Vergez, C.: A high order purely frequency-based harmonic balance formulation for continuation of periodic solutions. J. Sound Vib. 324(1), 243–262 (2009)

    Article  Google Scholar 

  65. Veltz, R.: BifurcationKit.jl (2020).https://hal.archives-ouvertes.fr/hal-02902346

  66. Faraday, M.: On a peculiar class of acoustical figures; and on certain forms assumed by groups of particles upon vibrating elastic surfaces. In Abstracts of the Papers Printed in the Philosophical Transactions of the Royal Society of London, number 3, pp. 49–51. The Royal Society London (1837)

  67. Mathieu, É.: Mémoire sur le mouvement vibratoire d’une membrane de forme elliptique. J. Math. 2(13), 137–203 (1968)

    MATH  Google Scholar 

  68. Nayfeh, A.H., Mook, D.T.: Nonlinear oscillations. John Wiley & sons, New-York (1979)

    MATH  Google Scholar 

  69. Zavodney, L.D., Nayfeh, A.H.: The non-linear response of a slender beam carrying a lumped mass to a principal parametric excitation: theory and experiment. Int. J. Non-Linear Mech. 24(2), 105–125 (1989)

    Article  MATH  Google Scholar 

  70. Lacarbonara, W., Yabuno, H., Hayashi, K.: Non-linear cancellation of the parametric resonance in elastic beams: theory and experiment. Int. J. Solids Struct. 44(7–8), 2209–2224 (2007)

    Article  MATH  Google Scholar 

  71. Givois, A., Grolet, A., Thomas, O., Deü, J.-F.: On the frequency response computation of geometrically nonlinear flat structures using reduced-order finite element models. Nonlinear Dyn. 97(2), 1747–1781 (2019)

    Article  MATH  Google Scholar 

  72. Comi, C., Corigliano, A., Zega, V., Zerbini, S.: Non linear response and optimization of a new z-axis resonant micro-accelerometer. Mechatronics 40, 235–243 (2016)

    Article  Google Scholar 

  73. Zega, V., Opreni, A., Mussi, G., Kwon, H.-K., Vukasin, G., Gattere, G., Langfelder, G., Frangi, A., Kenny, T. W.: Thermal stability of DETF MEMS resonators: numerical modelling and experimental validation. In 2020 IEEE 33rd international conference on micro electro mechanical systems (MEMS), pp. 1207–1210. IEEE, (2020)

  74. Dou, S., Strachan, B.S., Shaw, S.W., Jensen, J.S.: Structural optimization for nonlinear dynamic response. Philos. Trans. R. Soc. A: Math., Phys. Eng. Sci. 373(2051), 20140408 (2015)

  75. Dou, S., Jensen, J.S.: Optimization of nonlinear structural resonance using the incremental harmonic balance method. J. Sound Vib. 334, 239–254 (2015)

    Article  Google Scholar 

  76. Czaplewski, D.A., Chen, C., López, D., Shoshani, O., Eriksson, A.M., Strachan, S., Shaw, S.W.: Bifurcation generated mechanical frequency comb. Phys. Rev. Lett. 121(24), 244302 (2018)

    Article  Google Scholar 

  77. Chen, C., Zanette, D.H., Czaplewski, D.A., Shaw, S.W., López, D.: Direct observation of coherent energy transfer in nonlinear micromechanical oscillators. Nat. Commun. 8(1), 1–7 (2017)

    Google Scholar 

Download references

Funding

The work received no additional funding.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andrea Opreni.

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

Weak formulation of the invariance equations

As remarked in Sect. 3, an alternative representation of the nonlinear change of coordinates is obtained by defining the nonlinear mappings between the normal coordinates and the continuous fields (here displacements and velocities), rather than operating from the nodal values from the FE discretisation. This alternative formulation will thus provide a more general setting, where one could for example use another numerical method for the space discretisation of the problem.

For the purpose of deriving this generalisation, let us first restart from the finite element approximation, which expressed the involved fields as a linear combination of a finite set of basis functions:

$$\begin{aligned} \varvec{u}_h = \sum _{i=1}^{\textrm{N}} \varvec{l}_{i}\textrm{U}_i, \qquad \varvec{w}_h = \sum _{i=1}^{\textrm{N}} \varvec{l}_{i}\textrm{W}_i, \end{aligned}$$
(128)

with \(\varvec{l}_{i}\) nodal shape functions. For the present treatment, the same shape functions are used to interpolate both displacement and test fields. This choice is adopted only to simplify the presentation. We can then introduce a coordinate change between normal coordinates and continuous fields of the type:

$$\begin{aligned} \varvec{u}_h = \varvec{\varPsi }_{h}(\textbf{z}), \end{aligned}$$
(129)

with:

$$\begin{aligned} \varvec{\varPsi }_{h}(\textbf{z}) = \sum _{p=0}^{o} [\varvec{\varPsi }_{h,\mathcal {I}}(\textbf{z})]_{p}. \end{aligned}$$
(130)

We can therefore expand each tensor order as:

$$\begin{aligned}{}[\varvec{\varPsi }_{h}(\textbf{z})]_{p} = \sum _{\mathcal {I}} \varvec{\varPsi }^{(p)}_{h,\mathcal {I}} \pi _{\mathcal {I}}^{(p)}, \end{aligned}$$
(131)

within this setting, each mapping field \(\varvec{\varPsi }^{(p)}_{h,\mathcal {I}}\) is a continuous field since it is defined pointwise over the entire domain \(\textrm{B}\). We can then express each mapping field using a finite element approximation identical to that used for the displacement field itself:

$$\begin{aligned} \varvec{\varPsi }^{(p)}_{h,\mathcal {I}} = \sum _{i=1}^{\textrm{N}} \Psi ^{(p)}_{h,\mathcal {I},i} \varvec{l}_i, \end{aligned}$$
(132)

where \(\Psi ^{(p)}_{h,\mathcal {I},i}\) corresponds to the nodal values of the mappings. We remark that these nodal values are mathematically equivalent to the entries of the mapping vectors used on the discretised problem presented in the work. This representation is exploited for the computation of the left-hand sides of the homological equations, without the need to compute the complete nonlinearity tensors, which is faster and more memory efficient. To better appreciate it, let us report the weak form associated with the first-order autonomous problem:

$$\begin{aligned}&\int _{\textrm{B}} \rho \, \dot{\varvec{v}}_{h} \cdot \varvec{w}_{h} \, \textrm{dB} + \int _{\textrm{B}} \varvec{\varepsilon }(\varvec{u}_{h}):\mathcal {A}:\varvec{\varepsilon }(\varvec{w}_{h}) \,\textrm{dB} \nonumber \\&+ \int _{\textrm{B}} \frac{1}{2} \varvec{\gamma }(\varvec{u}_{h},\varvec{u}_{h}):\mathcal {A}:\varvec{\varepsilon }(\varvec{w}_{h}) + \varvec{\varepsilon }(\varvec{u}_{h}):\mathcal {A}:\varvec{\gamma }(\varvec{u}_{h},\varvec{w}_{h}) \,\textrm{dB} \nonumber \\&{+} \int _{\textrm{B}} \frac{1}{2} \varvec{\gamma }(\varvec{u}_{h},\varvec{u}_{h}):\mathcal {A}:\varvec{\gamma }(\varvec{u}_{h},\varvec{w}_{h}) \,\textrm{dB} = 0, \quad \forall \, \varvec{w}_{h} {\in } \mathcal {C}_{h}(\textbf{0}), \end{aligned}$$
(133a)
$$\begin{aligned}&\int _{\textrm{B}} \rho \, \dot{\varvec{u}}_h \cdot \varvec{w}_{h} \,\textrm{dB} = \int _{\textrm{B}} \rho \, \varvec{v}_h \cdot \varvec{w}_{h} \textrm{dB} \quad \forall \, \varvec{w}_{h} \in \mathcal {C}_{h}(\textbf{0}), \end{aligned}$$
(133b)

where the finite-dimensional velocity field \(\dot{\varvec{u}}_h=\varvec{v}_h\) has been introduced. Time derivatives are taken exactly as for the discrete problem:

$$\begin{aligned} \dot{\varvec{u}}_h = \nabla _{\textbf{z}}\varvec{\varPsi }_{h}(\textbf{z})\varvec{f}(\textbf{z}), \quad \dot{\varvec{v}}_h = \nabla _{\textbf{z}}\varvec{\varUpsilon }_{h}(\textbf{z})\varvec{f}(\textbf{z}), \end{aligned}$$
(134)

And upon substitution of mapping fields and their time derivatives into Eq. (133), we retrieve again the invariance equations, which are now formulated at the level of the weak form of the partial differential equation:

$$\begin{aligned}&\int _{\textrm{B}} \rho \, \nabla _{\textbf{z}}\varvec{\varUpsilon }_{h}(\textbf{z})\varvec{f}(\textbf{z}) \cdot \varvec{w}_{h} \, \textrm{dB} + \int _{\textrm{B}} \varvec{\varepsilon }(\varvec{\varPsi }_{h}(\textbf{z})):\mathcal {A}:\varvec{\varepsilon }(\varvec{w}_{h}) \,\textrm{dB} \nonumber \\&+~\int _{\textrm{B}} \!\frac{1}{2} \varvec{\gamma }(\varvec{\varPsi }_{h}(\textbf{z}),\varvec{\varPsi }_{h}(\textbf{z}))\!:\!\mathcal {A}\!:\!\varvec{\varepsilon }(\varvec{w}_{h}) \!\! +\!\! \varvec{\varepsilon }(\varvec{\varPsi }_{h}(\textbf{z}))\!:\!\mathcal {A}\!:\!\varvec{\gamma }(\varvec{\varPsi }_{h}(\textbf{z}),\varvec{w}_{h}) \textrm{dB} \nonumber \\&+ \int _{\textrm{B}} \frac{1}{2} \varvec{\gamma }(\varvec{\varPsi }_{h}(\textbf{z}),\varvec{\varPsi }_{h}(\textbf{z})):\mathcal {A}:\varvec{\gamma }(\varvec{\varPsi }_{h}(\textbf{z}),\varvec{w}_{h}) \,\textrm{dB} = 0, \quad \forall \, \varvec{w}_{h} \in \mathcal {C}_{h}(\textbf{0}), \end{aligned}$$
(135a)
$$\begin{aligned}&\int _{\textrm{B}} \rho \, \nabla _{\textbf{z}}\varvec{\varPsi }_{h}(\textbf{z})\varvec{f}(\textbf{z}) \cdot \varvec{w}_{h} \,\textrm{dB} = \int _{\textrm{B}} \rho \, \varvec{\varUpsilon }_{h}(\textbf{z}) \cdot \varvec{w}_{h} \textrm{dB}, \quad \forall \, \varvec{w}_{h} \in \mathcal {C}_{h}(\textbf{0}), \end{aligned}$$
(135b)

These two invariance equations are completely equivalent to Eq. (29), if the finite element approximation is identical and if the same integration rules are used. The interesting feature of Eq. (135) is that it represents, for each monomial of the asymptotic expansion, a linear elliptic partial differential equation, already discretised in a finite-dimensional space. As a result, all matrices and left-hand sides can be computed from the integration of Eq. (135) without the need to export the full nonlinearity tensor of the system. Also, Eq. (135) can be used in a more general setting and one could enlarge the scope and select another numerical method than the finite element procedure to solve out the elliptic PDEs. Furthermore, no constraint on the finite element scheme is required, hence any Galerkin method can be applied.

Multivariate tensor notation

Most quantities adopted in the present work are coefficients of multivariate polynomials. In this Section we recall the tensorial representation introduced in [35] to derive such quantities.

A generic vector of multivariate polynomials \(\textbf{a}(\textbf{z})\) can be expressed as a summation of terms of increasing order from zero, up to the maximum order of the expansion o:

$$\begin{aligned} \textbf{a}(\textbf{z}) = \sum _{p=0}^{o} [\textbf{a}(\textbf{z})]_{p}. \end{aligned}$$
(136)

for each expansion order, polynomial terms are explicitly defined as:

$$\begin{aligned}{}[\textbf{a}(\textbf{z})]_{p} = \sum _{i_1=1}^{n} \sum _{i_2=1}^{n} \ldots \sum _{i_p=1}^{n} \textbf{a}^{(p)}_{\{i_1 i_2\ldots i_p\}} {z}_{i_1} {z}_{i_2} \ldots {z}_{i_p},\nonumber \\ \end{aligned}$$
(137)

with n the number of variables of the polynomial expansion. Alternatively, the following expansion can be used:

$$\begin{aligned}{}[\textbf{a}(\textbf{z})]_{p} = \sum _{\mathcal {I}} \textbf{a}^{(p)}_{\mathcal {I}} \pi _{\mathcal {I}}^{(p)}, \end{aligned}$$
(138)

where indexes and variables products are collected with the following notation:

$$\begin{aligned} \mathcal {I}= \{ i_1 i_2 \ldots i_{p} \}, \qquad \pi _{\mathcal {I}}^{(p)} = {z}_{i_1} {z}_{i_2} \ldots {z}_{i_p}. \end{aligned}$$
(139)

For a given polynomial with order p, there exist a finite number of indexes combinations \(\mathcal {I}\) corresponding to the expansion monomials. Such combinations are collected in sets \({\mathcal {H}}^{(p)}\).

As a natural consequence of the above notation, vectors belonging to an asymptotic expansion are in general expressed as:

$$\begin{aligned} \textbf{a}^{(p)}_{\mathcal {I}} = \textbf{a}^{(p)}_{\{i_1 i_2\ldots i_p\}}, \end{aligned}$$
(140)

or alternatively, instead of writing vectors using the compact bold-typed notation, it is possible to express vector components using the following notation:

$$\begin{aligned} \textrm{a}_{s\mathcal {I}}^{(p)} = \textrm{a}_{s\{i_1 i_2\ldots i_p\}}^{(p)}, \end{aligned}$$
(141)

where the first subscript s stands for the s-th vector entry. If multiple vectors are present for a given monomial and order, as for instance when multiple forcing eigenvalues are present, a further index may be needed. Expansion are in general expressed as:

$$\begin{aligned} \textbf{a}^{(p)}_{j,\mathcal {I}} = \textbf{a}^{(p)}_{j,\{i_1 i_2\ldots i_p\}}, \end{aligned}$$
(142a)
$$\begin{aligned} \textrm{a}_{j,s\mathcal {I}}^{(p)} = \textrm{a}_{j,s\{i_1 i_2\ldots i_p\}}^{(p)}, \end{aligned}$$
(142b)

where index j refers to the j-th forcing eigenvalue.

Nonlinearity tensors computation

An important remark on dynamical systems stemming from partial differential equations is that their discretised algebraic formulation is equivalent to their finite-dimensional weak formulation up to numerical integration errors. This has already been underlined in Appendix A, where a finite-dimensional weak form for the invariance equation is derived. The natural result is that nonlinearities can be computed through integration over the domain. Let us consider the expressions reported in Eqs. (38d)–(38e). The most efficient technique to compute such term is simply to derive them through numerical integration of their associated weak formulation. Regarding the quadratic operator, its assemblage is performed as:

$$\begin{aligned}&\textrm{G}_{\mathcal {I}}^{(p)} =\sum _{k=1}^{p-1} \,\varvec{G}\left( \varvec{\varPsi }^{(k)}_{\{i_1\ldots i_{k}\}}, \varvec{\varPsi }^{(p-k)}_{\{i_{k+1}\ldots i_{p}\}} \right) \nonumber \\&\,=\sum _{k=1}^{p-1} \frac{1}{2} \int _{\textrm{B}} \,\varvec{\gamma }\left( \varvec{\varPsi }^{(k)}_{h,\{i_1\ldots i_{k}\}},\varvec{\varPsi }^{(p-k)}_{h,\{i_{k+1}\ldots i_{p}\}}\right) :\mathcal {A}:\varvec{\varepsilon }(\varvec{w}_{h}) \nonumber \\&\, + \varvec{\varepsilon }\left( \varvec{\varPsi }^{(k)}_{h,\{i_1\ldots i_{k}\}}\right) :\mathcal {A}:\varvec{\gamma }\left( \varvec{\varPsi }^{(p-k)}_{h,\{i_{k+1}\ldots i_{p}\}},\varvec{w}_{h}\right) \nonumber \\&\, + \varvec{\varepsilon }\left( \varvec{\varPsi }^{(p-k)}_{h,\{i_{k+1}\ldots i_{p}\}}\right) :\mathcal {A}:\varvec{\gamma }\left( \varvec{\varPsi }^{(k)}_{h,\{i_1\ldots i_{k}\}},\varvec{w}_{h}\right) \,\textrm{dB}, \end{aligned}$$
(143)

where the operator has been recasted in symmetric form. Similar treatment is performed for the \(\varvec{H}(\textbf{U},\textbf{U},\textbf{U})\) operator:

$$\begin{aligned} \textbf{H}_{\mathcal {I}}^{(p)} = \sum _{k=1}^{p-2}\sum _{l=1}^{p-k-1}&\,\varvec{H}\left( \varvec{\varPsi }^{(k)}_{\{i_1\ldots i_{k}\}},\varvec{\varPsi }^{(l)}_{\{i_{k+1}\ldots i_{k+l}\}},\right. \nonumber \\&\quad \left. \varvec{\varPsi }^{(p-k-l)}_{\{i_{k+l+1}\ldots i_{p}\}}\right) = \nonumber \\ \sum _{k=1}^{p-2}\sum _{l=1}^{p-k-1} \frac{1}{6} \int _{\textrm{B}}&\, \varvec{\gamma }\left( \varvec{\varPsi }^{(k)}_{h,\{i_1\ldots i_{k}\}},\varvec{\varPsi }^{(l)}_{h,\{i_{k+1}\ldots i_{k+l}\}} \right) :\nonumber \\&\quad \mathcal {A}:\varvec{\gamma }\left( \varvec{\varPsi }^{(p-k-l)}_{h,\{i_{k+l+1}\ldots i_{p}\}},\varvec{w}_{h} \right) \nonumber \\&\, + \varvec{\gamma }\left( \varvec{\varPsi }^{(p-k-l)}_{h,\{i_{k+l+1}\ldots i_{p}\}}, \varvec{\varPsi }^{(k)}_{h,\{i_1\ldots i_{k}\}} \right) :\nonumber \\&\quad \mathcal {A}:\varvec{\gamma }\left( \varvec{\varPsi }^{(l)}_{h,\{i_{k+1}\ldots i_{k+l}\}},\varvec{w}_{h} \right) \nonumber \\&\, + \varvec{\gamma }\left( \varvec{\varPsi }^{(l)}_{h,\{i_{k+1}\ldots i_{k+l}\}},\varvec{\varPsi }^{(p-k-l)}_{h,\{i_{k+l+1}\ldots i_{p}\}} \right) :\nonumber \\&\quad \mathcal {A}:\varvec{\gamma }\left( \varvec{\varPsi }^{(k)}_{h,\{i_1\ldots i_{k}\}},\varvec{w}_{h} \right) \,\textrm{dB}, \end{aligned}$$
(144)

Non-autonomous gradients

In Eq. (32), terms involving a gradient with respect to the normal coordinate require a dedicated treatment. In particular, two new terms depending on such gradient are involved. The first involves the gradient of the non-autonomous mappings:

(145a)
(145b)

where the shortcut notation has been introduced to refer to a term with order strictly larger than one and smaller than p. Expanding the product in the right-hand side, one can rewrite:

(146a)
(146b)

The second term at the right-hand side of Eq. (146) can be explicitly rewritten as:

(147a)
(147b)

where all the terms have been grouped and expanded onto the monomials \(\pi _{\mathcal {I}}^{(p)}\), thus making appear new coefficients \(\bar{\varvec{\mu }}_{\mathcal {I}}^{(p)}\) and \(\bar{\varvec{\nu }}_{\mathcal {I}}^{(p)}\), the full expressions of which can be written as:

$$\begin{aligned} \bar{\varvec{\mu }}_{\mathcal {I}}^{(p)} =&\, \sum _{s=1}^{2\textrm{n}} \sum _{k=2}^{p} \sum _{l=0}^{p-k} \hat{\varvec{\Psi }}^{(p-k+1)}_{\{i_1 \ldots i_{l}\,s\,i_{l+k+1} \ldots i_{p}\}} f_{s\{i_{l+1}\ldots i_{l+k}\}}^{(k)}, \end{aligned}$$
(148a)
$$\begin{aligned} \bar{\varvec{\nu }}_{\mathcal {I}}^{(p)} =&\, \sum _{s=1}^{2\textrm{n}} \sum _{k=2}^{p} \sum _{l=0}^{p-k} \hat{\varvec{\Upsilon }}^{(p-k+1)}_{\{i_1 \ldots i_{l}\,s\,i_{l+k+1} \ldots i_{p}\}} f_{s\{i_{l+1}\ldots i_{l+k}\}}^{(k)}. \end{aligned}$$
(148b)

These new terms have exactly the same structure as those appearing in Eqs. (38b)–(38c) computed for the autonomous part, where one just need to replace the terms of the nonlinear mapping, which has important consequence at the computational level and ease the writing.

The second term involving a gradient and needing a dedicated development involves the gradient of the known autonomous mappings with respect to the normal coordinates, now multiplied with the reduced dynamics of the unknown time-dependent part. It can be expressed as:

(149a)
(149b)

which in turn can be rewritten as:

(150a)
(150b)

The second term on the right-hand side can be explicitly defined as:

(151a)
(151b)

which provides the following compact expression for Eq. (150):

$$\begin{aligned} \left[ \nabla _{\textbf{z}}\varvec{\varPsi }(\textbf{z})\hat{\varvec{f}}(\textbf{z}) \right] =&\, \sum _{\mathcal {I}}\left( \sum _{s=1}^{2\textrm{n}} (\pmb {\phi }_s \hat{f}_{s\mathcal {I}}^{(p)}) + \bar{\bar{\varvec{\mu }}}_{\mathcal {I}}^{(p)} \right) \pi _{\mathcal {I}}^{(p)}, \end{aligned}$$
(152a)
$$\begin{aligned} \left[ \nabla _{\textbf{z}}\varvec{\varUpsilon }(\textbf{z})\hat{\varvec{f}}(\textbf{z}) \right] =&\, \sum _{\mathcal {I}}\left( \sum _{s=1}^{2\textrm{n}} (\pmb {\phi }_s \lambda _s \hat{f}_{s\mathcal {I}}^{(p)}) + \bar{\bar{\varvec{\nu }}}_{\mathcal {I}}^{(p)} \right) \pi _{\mathcal {I}}^{(p)}, \end{aligned}$$
(152b)

where the new terms provided by the above developments are explicitly computed as:

(153a)
(153b)

Again, these coefficients can be directly compared to those reported in Eqs. (38b)–(38c) and in Eqs. (148), since sharing exactly the same structure, where now the non-autonomous part (tilde terms) is on the reduced dynamics coefficients. At the computational level, this symmetric structure can be used to compute easily these new terms.

Explicit expressions for the RHS of low-order \({\varepsilon }^{1}\)-homological equations and reduced dynamics coefficients

In the present section, we report explicit expressions of the right-hand sides \(\hat{\textbf{E}}_{\mathcal {I}}^{(p)}\) for low-order developments. This serves the purpose of highlighting how different terms in the reduced dynamics are obtained and how they are affected by the forcing and the geometrical parameters of the structure.

For the zero-order homological equations, \(\hat{\textbf{E}}^{(0)}\) is simply given by the applied forcing:

$$\begin{aligned} \hat{\textbf{E}}^{(0)} =&\, \hat{\textbf{F}}, \end{aligned}$$
(154a)
$$\begin{aligned} \hat{\varvec{\mu }}^{(0)} =&\, \textbf{0}. \end{aligned}$$
(154b)

Furthermore, since \(\hat{\varvec{\mu }}^{(0)}=\textbf{0}\), the resulting mapping is co-linear with the applied forcing \(\hat{\textbf{F}}\) and reduced dynamics coefficients \(\hat{\textbf{F}}^{0}\) are proportional to the projection of the forcing vector onto the master modal subspace, as detailed in Section 4. This conclusion is obtained by considering that the reduced dynamics coefficient can be computed from the relation:

$$\begin{aligned} \hat{f}^{(p)}_{j\,\mathcal {I}}= & {} \textbf{X}_{j}^{\textrm{T}} \left[ \begin{array}{c} \hat{\textbf{E}}^{(p)}_{\mathcal {I}} \\ -\textbf{M}\hat{\varvec{\mu }}^{(p)}_{\mathcal {I}} \end{array}\right] = \frac{\pmb {\phi }_j^{\textrm{T}}}{\lambda _{j} - \bar{\lambda }_{j}} \hat{\textbf{E}}^{(p)}_{\mathcal {I}} \nonumber \\{} & {} + \frac{\bar{\lambda }_{j} \pmb {\phi }_j^{\textrm{T}}}{\lambda _{j} - \bar{\lambda }_{j}} \textbf{M}\hat{\varvec{\mu }}^{(p)}_{\mathcal {I}}, \end{aligned}$$
(155)

which for the zero-order development yields:

$$\begin{aligned} \hat{f}^{(0)}_{j} = \frac{\pmb {\phi }_j^{\textrm{T}}}{\lambda _{j} - \bar{\lambda }_{j}} \hat{\textbf{F}}. \end{aligned}$$
(156)

The first-order non-autonomous asymptotic development yields:

$$\begin{aligned} \hat{\textbf{E}}_{\mathcal {I}}^{(1)} = \hat{\textbf{E}}_{\{i_1\}}^{(1)} =&\, -\textbf{M}\sum _{s=1}^{2\textrm{n}}\left[ {\varvec{\Upsilon }}^{(2)}_{\{s \,i_1\}} \hat{f}^{(0)}_{s} + {\varvec{\Upsilon }}^{(2)}_{\{i_1 s\}} \hat{f}^{(0)}_{s} \right] \nonumber \\&\quad - 2\varvec{G}(\varvec{\Psi }^{(1)}_{\{i_1\}}, \hat{\varvec{\Psi }}^{(0)}), \end{aligned}$$
(157a)
$$\begin{aligned} \hat{\varvec{\mu }}^{(1)}_{\{i_1\}} =&\, \sum _{s=1}^{2\textrm{n}} \left[ {\varvec{\Psi }}^{(2)}_{\{s\,i_1\}} \hat{f}^{(0)}_{s} + {\varvec{\Psi }}^{(2)}_{\{i_1 s\}} \hat{f}^{(0)}_{s} \right] , \end{aligned}$$
(157b)

thus providing an explicit expression for the reduced dynamics coefficients:

$$\begin{aligned}{} & {} \hat{f}^{(1)}_{j\{i_1\}} = \frac{\pmb {\phi }_j^{\textrm{T}}}{\lambda _{j} - \bar{\lambda }_{j}} \left[ -\textbf{M}\sum _{s=1}^{2\textrm{n}}\left[ {\varvec{\Upsilon }}^{(2)}_{\{s\, i_1\}} \hat{f}^{(0)}_{s} +{\varvec{\Upsilon }}^{(2)}_{\{i_1 s\}} \hat{f}^{(0)}_{s} \right] \right. \nonumber \\{} & {} \left. - 2\varvec{G}(\varvec{\Psi }^{(1)}_{\{i_1\}},\hat{\varvec{\Psi }}^{(0)}) \right] + \frac{\bar{\lambda }_{j} \pmb {\phi }_j^{\textrm{T}}}{\lambda _{j} - \bar{\lambda }_{j}} \left[ \sum _{s=1}^{2\textrm{n}} \textbf{M}\left[ {\varvec{\Psi }}^{(2)}_{\{s\,i_1\}} \hat{f}^{(0)}_{s}\right. \right. \nonumber \\{} & {} \left. \left. + \varvec{\Psi }^{(2)}_{\{i_1 s\}} \hat{f}^{(0)}_{s} \right] \right] , \end{aligned}$$
(158)

which highlights how the reduced dynamics coefficients at first order are heavily affected by the \({{\varvec{G}}}(\varvec{\Psi }^{(1)}_{\{i_1\}},\hat{\varvec{\Psi }}^{(0)})\) vector. Finally, the second-order expansions yields:

$$\begin{aligned} \hat{\textbf{E}}_{\mathcal {I}}^{(2)} = \hat{\textbf{E}}_{\{i_1i_2\}}^{(2)} =&\, -\textbf{M}\sum _{s=1}^{2\textrm{n}}\left[ {\varvec{\Upsilon }}^{(3)}_{\{s\, i_1i_2\}} \hat{f}^{(0)}_{s} +{\varvec{\Upsilon }}^{(3)}_{\{i_1 s \,i_2\}} \hat{f}^{(0)}_{s}\right. \nonumber \\&\quad \left. + {\varvec{\Upsilon }}^{(3)}_{\{i_1 i_2 s\}} \hat{f}^{(0)}_{s} \right] \nonumber \\&\, - \textbf{M}\sum _{s=1}^{2\textrm{n}}\left[ {\varvec{\Upsilon }}^{(2)}_{\{i_1 s\}} \hat{f}^{(1)}_{s\, \{i_2\}} + {\varvec{\Upsilon }}^{(2)}_{\{s\, i_2\}} \hat{f}^{(1)}_{s \{i_1\}} \right] \nonumber \\&\quad - \textbf{M}\sum _{s=1}^{2\textrm{n}} \left[ \hat{\varvec{\Upsilon }}^{(1)}_{\{s\}} f^{(2)}_{s\{ i_1 i_2\}} \right] \nonumber \\&\, + 2\varvec{ G}(\varvec{\Psi }^{(2)}_{\{i_1 i_2\}}, \hat{\varvec{\Psi }}^{(0)} ) + 2\varvec{ G}(\varvec{\Psi }^{(1)}_{\{i_1\}}, \hat{\varvec{\Psi }}^{(1)}_{\{i_2\}} ) \nonumber \\&\quad + 3\varvec{ H}(\varvec{\Psi }^{(1)}_{\{i_1\}}, \varvec{\Psi }^{(1)}_{\{i_2\}}, \hat{\varvec{\Psi }}^{(0)}) \; , \end{aligned}$$
(159a)
$$\begin{aligned} \hat{\varvec{\mu }}_{\{i_1i_2\}}^{(2)} =&\, - \sum _{s=1}^{2\textrm{n}}\left[ \varvec{\Psi }^{(3)}_{\{s\,i_1i_2\}} \hat{f}^{(0)}_{s} \right. \nonumber \\&\quad \left. + \varvec{\Psi }^{(3)}_{\{i_1 s\,i_2\}} \hat{f}^{(0)}_{s} +\varvec{\Psi }^{(3)}_{\{i_1 i_2 s\}} \hat{f}^{(0)}_{s} \right] \nonumber \\&\, - \sum _{s=1}^{2\textrm{n}}\left[ \varvec{\Psi }^{(2)}_{\{i_1 s\}} \hat{f}^{(1)}_{s\{i_2\}} + \varvec{\Psi }^{(2)}_{\{s \,i_2\}} \hat{f}^{(1)}_{s \{i_1\}} \right] \nonumber \\&\quad - \sum _{s=1}^{2\textrm{n}} \left[ \hat{\varvec{\Psi }}^{(1)}_{\{s\}} f^{(2)}_{s\, \{i_1 i_2\}} \right] \; , \end{aligned}$$
(159b)

and the second-order reduced dynamics coefficients:

$$\begin{aligned} \hat{f}^{(2)}_{j\,\{i_1 i_2\}} =&\,\frac{\pmb {\phi }_j^{\textrm{T}}}{\lambda _{j} - \bar{\lambda }_{j}} \,\biggl [ \biggl . -\textbf{M}\sum _{s=1}^{2\textrm{n}}\left[ \varvec{\Upsilon }^{(3)}_{\{s\,i_1i_2\}} \hat{f}^{(0)}_{s} \right. \nonumber \\&\left. + \varvec{\Upsilon }^{(3)}_{\{i_1 s\,i_2\}} \hat{f}^{(0)}_{s} + \varvec{\Upsilon }^{(3)}_{\{i_1 i_2 s\}} \hat{f}^{(0)}_{s} \right] \nonumber \\&\, - \textbf{M}\sum _{s=1}^{2\textrm{n}}\left[ \varvec{\Upsilon }^{(2)}_{\{i_1 s\}} \hat{f}^{(1)}_{s\{i_2\}}\right. \nonumber \\&\left. + \varvec{\Upsilon }^{(2)}_{\{s\,i_2\}} \hat{f}^{(1)}_{s\{i_1\}} \right] - \textbf{M}\sum _{s=1}^{2\textrm{n}} \left[ \hat{\varvec{\Upsilon }}^{(1)}_{\{s\}} f^{(2)}_{s\{i_1 i_2\}} \right] \nonumber \\&\, + 2\varvec{ G}(\varvec{\Psi }^{(2)}_{\{i_1 i_2\}}, \hat{\varvec{\Psi }}^{(0)} ) + 2\varvec{ G}(\varvec{\Psi }^{(1)}_{\{i_1\}}, \hat{\varvec{\Psi }}^{(1)}_{\{i_2\}} ) \nonumber \\&\,+ 3\varvec{ H}(\varvec{\Psi }^{(1)}_{\{i_1\}}, \varvec{\Psi }^{(1)}_{\{i_2\}}, \hat{\varvec{\Psi }}^{(0)}) \biggl .\biggl ] \nonumber \\&+ \frac{\bar{\lambda }_{j} \pmb {\phi }_j^{\textrm{T}}}{\lambda _{j} - \bar{\lambda }_{j}} \, \biggl [ \biggl . - \textbf{M}\sum _{s=1}^{2\textrm{n}}\left[ \varvec{\Psi }^{(3)}_{\{s\, i_1i_2\}} \hat{f}^{(0)}_{s}\right. \nonumber \\&\left. + \varvec{\Psi }^{(3)}_{\{i_1 s\, i_2\}} \hat{f}^{(0)}_{s} + \varvec{\Psi }^{(3)}_{\{i_1 i_2 s\}} \hat{f}^{(0)}_{s} \right] \nonumber \\&\, - \sum _{s=1}^{2\textrm{n}} \textbf{M}\left[ \varvec{\Psi }^{(2)}_{\{i_1 s\}} \hat{f}^{(1)}_{s\, i_2} + \varvec{\Psi }^{(2)}_{\{s\,i_2\}} \hat{f}^{(1)}_{s\{i_1\}} \right] \nonumber \\&- \sum _{s=1}^{2\textrm{n}} \textbf{M}\left[ \hat{\varvec{\Psi }}^{(1)}_{\{s\}} f^{(2)}_{s\{i_1 i_2\}} \right] \biggl .\biggl ]. \end{aligned}$$
(160)

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

Opreni, A., Vizzaccaro, A., Touzé, C. et al. High-order direct parametrisation of invariant manifolds for model order reduction of finite element structures: application to generic forcing terms and parametrically excited systems. Nonlinear Dyn 111, 5401–5447 (2023). https://doi.org/10.1007/s11071-022-07978-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-022-07978-3

Keywords

Navigation