Skip to main content
Log in

On the frequency response computation of geometrically nonlinear flat structures using reduced-order finite element models

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

Abstract

This paper presents a general methodology to compute nonlinear frequency responses of flat structures subjected to large amplitude transverse vibrations, within a finite element context. A reduced-order model (ROM) is obtained by an expansion onto the eigenmode basis of the associated linearized problem, including transverse and in-plane modes. The coefficients of the nonlinear terms of the ROM are computed thanks to a non-intrusive method, using any existing nonlinear finite element code. The direct comparison to analytical models of beams and plates proves that a lot of coefficients can be neglected and that the in-plane motion can be condensed to the transverse motion, thus giving generic rules to simplify the ROM. Then, a continuation technique, based on an asymptotic numerical method and the harmonic balance method, is used to compute the frequency response in free (nonlinear mode computation) or harmonically forced vibrations. The whole procedure is tested on a straight beam, a clamped circular plate and a free perforated plate for which some nonlinear modes are computed, including internal resonances. The convergence with harmonic numbers and oscillators is investigated. It shows that keeping a few of them is sufficient in a range of displacements corresponding to the order of the structure’s thickness, with a complexity of the simulated nonlinear phenomena that increase very fast with the number of harmonics and oscillators.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20

Similar content being viewed by others

Notes

  1. In fact, the in-plane frequencies normalized values depend on parameter \(\mu \) defined by Eqs. (2), (53) but have a priori a negligible influence on the transverse motion.

  2. We consider here only the transverse displacement degrees of freedom and not the eventual rotation ones, encountered in Timoshenko (shear deformation) 1D finite elements.

  3. See footnote 2.

References

  1. Chen, C., Zanette, D.H., Czaplewsk, D.A., Shaw, S., López, D.: Direct observation of coherent energy transfer in nonlinear micromechanical oscillators. Nat. Commun. 8, 15523 (2017)

    Article  Google Scholar 

  2. Thomas, O., Mathieu, F., Mansfield, W., Huang, C., Trolier-McKinstry, S., Nicu, L.: Efficient parametric amplification in MEMS with integrated piezoelectric actuation and sensing capabilities. Appl. Phys. Lett. 102(16), 163504 (2013). https://doi.org/10.1063/1.4802786

    Article  Google Scholar 

  3. Dezest, D., Thomas, O., Mathieu, F., Mazenq, L., Soyer, C., Costecalde, J., Remiens, D., Deü, J.-F., Nicu, L.: Wafer-scale fabrication of self-actuated piezoelectric nanoelectromechanical resonators based on lead zirconate titanate (PZT). J. Micromech. Microengineering 25(3), 035002 (2015)

    Article  Google Scholar 

  4. Quinn, D.D., Triplett, A.L., Vakakis, A.F., Bergman, L.A.: Energy harvesting from impulsive loads using intentional essential nonlinearities. J. Vibr. Acoust. 133(1), 011004 (2011)

    Article  Google Scholar 

  5. Ducceschi, M., Touzé, C.: Modal approach for nonlinear vibrations of damped impacted plates: application to sound synthesis of gongs and cymbals. J. Sound Vibr. 344, 313–331 (2015)

    Article  Google Scholar 

  6. Monteil, M., Thomas, O., Touzé, C.: Identification of mode couplings in nonlinear vibrations of the steelpan. Appl. Acoust. 89, 1–15 (2015). https://doi.org/10.1016/j.apacoust.2014.08.008

    Article  Google Scholar 

  7. Grolet, A., Thouverez, F.: Free and forced vibration analysis of a nonlinear system with cyclic symmetry: application to a simplified model. J. Sound Vibr. 331(12), 2911–2928 (2012)

    Article  Google Scholar 

  8. Renson, L., Noël, J.P., Kerschen, G.: Complex dynamics of a nonlinear aerospace structure: numerical continuation and normal modes. Nonlinear Dyn. 79(2), 1293–1309 (2015)

    Article  Google Scholar 

  9. Noor, A.K.: Recent advances in reduction methods for nonlinear problems. Comput. Struct. 13(1–3), 31–44 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  10. Slaats, P.M.A., Jongh, J.D., Sauren, A.A.H.J.: Model reduction tools for nonlinear structural dynamics. Comput. Struct. 54(6), 1155–1171 (1995)

    Article  MATH  Google Scholar 

  11. von Kármán, T.: Festigkeitsprobleme im Maschinenbau. Encykl. Math. Wiss. 4(4), 311–385 (1910)

    MATH  Google Scholar 

  12. Nayfeh, A.H., Mook, D.T.: Nonlinear Oscillations. Wiley, London (2008)

    MATH  Google Scholar 

  13. Chu, H.-N., Herrmann, G.: Influence of large amplitudes on free flexural vibrations of rectangular elastic plates. J. Appl. Mech 23, 532–540 (1956)

    MathSciNet  MATH  Google Scholar 

  14. Thomas, O., Bilbao, S.: Geometrically non-linear flexural vibrations of plates: in-plane boundary conditions and some symmetry properties. J. Sound Vibr. 315(3), 569–590 (2008)

    Article  Google Scholar 

  15. Woinowski-Krieger, S.: The effect of axial force on the vibration of hinged bars. J. Appl. Mech. 17(2), 35–36 (1950)

    MathSciNet  MATH  Google Scholar 

  16. Eisley, J.G.: Nonlinear vibration of beams and rectangular plates. Z. Angew. Math. Phys. (ZAMP) 15(2), 167–175 (1964)

    Article  MathSciNet  MATH  Google Scholar 

  17. Ho, C.H., Scott, R.A., Eisley, J.G.: Non-planar, non-linear oscillations of a beam. part II: Free motions. J. Sound Vibr. 47(3), 333–339 (1976)

    Article  MATH  Google Scholar 

  18. Sridhar, S., Mook, D.T., Nayfeh, A.H.: Non-linear resonances in the forced responses of plates, part II: asymmetric responses of circular plates. J. Sound Vibr. 59(2), 159–170 (1975)

    Article  MATH  Google Scholar 

  19. Touzé, C., Thomas, O., Chaigne, A.: Asymmetric non-linear forced vibrations of free-edge circular plates, part I: theory. J. Sound Vibr. 258(4), 649–676 (2002). https://doi.org/10.1006/jsvi.2002.5143

    Article  Google Scholar 

  20. Ducceschi, M., Touzé, C., Bilbao, S., Webb, C.J.: Nonlinear dynamics of rectangular plates: investigation of modal interaction in free and forced vibrations. Acta Mech. 22(1), 213–232 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  21. Capiez-Lernout, E., Soize, C., Mignolet, M.P.: Computational stochastic statics of an uncertain curved structure with geometrical nonlinearity in three-dimensional elasticity. Computational Mechanics 49(1), 87–97 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  22. 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 

  23. Ribeiro, P., Petyt, M.: Non-linear vibration of beams with internal resonance by the hierarchical finite element method. J. Sound Vibr. 224(4), 591–624 (1999)

    Article  Google Scholar 

  24. Stoykov, S., Ribeiro, P.: Periodic geometrically nonlinear free vibrations of circular plates. J. Sound Vibr. 315, 536–555 (2008)

    Article  Google Scholar 

  25. McEwan, M.I., Wright, J.R., Cooper, J.E., Leung, A.Y.T.: A combined modal/finite element analysis technique for the dynamic response of a non-linear beam to harmonic excitation. J. Sound Vibr. 243(4), 601–624 (2001)

    Article  Google Scholar 

  26. Muravyov, A.A., Rizzi, S.A.: Determination of nonlinear stiffness with application to random vibration of geometrically nonlinear structures. Comput. Struct. 81(15), 1513–1523 (2003)

    Article  Google Scholar 

  27. Mignolet, M.P., Przekop, A., Rizzi, S., Spottswood, S.: A review of indirect/non-intrusive reduced-order modeling of nonlinear geometric structures. J. Sound Vibr. 332(10), 2437–2460 (2013)

    Article  Google Scholar 

  28. Perez, R., Wang, X.Q., Mignolet, M.P.: Nonintrusive structural dynamic reduced-order modeling for large deformations: enhancements for complex structures. J. Comput. Nonlinear Dyn. 9(3), 031008 (2014)

    Article  Google Scholar 

  29. Murthy, R., Wang, X.Q., Perez, R., Mignolet, M.P., Richter, L.A.: Uncertainty-based experimental validation of nonlinear reduced-order models. J. Sound Vibr. 331(5), 1097–1114 (2012)

    Article  Google Scholar 

  30. Claeys, M., Sinou, J.-J., Lambelin, J.-P., Alcoverro, B.: Multi-harmonic measurements and numerical simulations of nonlinear vibrations of a beam with non-ideal boundary conditions. Commun. Nonlinear Sci. Numer. Simul. 19, 4196–4212 (2014)

    Article  Google Scholar 

  31. O’Hara, P., Hollkamp, J.J.: Modeling vibratory damage with reduced-order models and the generalized finite element method. J. Sound Vibr. 333(24), 6637–6650 (2014)

    Article  Google Scholar 

  32. Ehrhardt, D.A., Allen, M.S., Beberniss, T.J., Neild, S.A.: Finite element model calibration of a nonlinear perforated plate. J. Sound Vibr. 392, 280–294 (2017)

    Article  Google Scholar 

  33. Kim, K., Radu, A.G., Wang, X.Q., Mignolet, M.P.: Nonlinear reduced-order modeling of isotropic and functionally graded plates. Int. J. Non-linear Mech. 49, 100–110 (2013)

    Article  Google Scholar 

  34. Lazarus, A., Thomas, O., Deü, J.-F.: Finite elements reduced order models for nonlinear vibrations of piezoelectric layered beams with applications to NEMS. Finite Elem. Anal. Des. 49(1), 35–51 (2012)

    Article  MathSciNet  Google Scholar 

  35. Hollkamp, J., Gordon, R.: Reduced-order models for nonlinear response prediction: implicit condensation and expansion. J. Sound Vibr. 318(4), 1139–1153 (2008)

    Article  Google Scholar 

  36. Kim, K., Khanna, V., Wang, X., Mignolet, M.: Nonlinear reduced order modeling of flat cantilevered structures. In: Proceedings of the 50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, p. 2492 (2009)

  37. Nash, M.: Nonlinear structural dynamics by finite element modal synthesis, Ph.D. thesis, Imperial College—University of London (1978)

  38. Rizzi, S.A., Przekop, A.: System identification-guided basis selection for reduced-order nonlinear response analysis. J. Sound Vibr. 315(3), 467–485 (2008)

    Article  Google Scholar 

  39. Przekop, A., Guo, X., Rizzi, S.A.: Alternative modal basis selection procedures for reduced-order nonlinear random response simulation. J. Sound Vibr. 331(17), 4005–4024 (2012)

    Article  Google Scholar 

  40. Kuether, R.J., Deaner, B., Hollkamp, J.J., Allen, M.S.: Evaluation of geometrically nonlinear reduced-order models with nonlinear normal modes. AIAA J. 53(11), 3273–3285 (2015)

    Article  Google Scholar 

  41. Idelsohn, S.R., Cardona, A.: A reduction method for nonlinear structural dynamic analysis. Comput. Methods Appl. Mech. Eng. 49(3), 253–279 (1985)

    Article  MATH  Google Scholar 

  42. Sombroek, C.S.M., Tiso, P., Renson, L., Kerschen, G.: Numerical computation of nonlinear normal modes in a modal derivative subspace. Comput. Struct. 195, 34–36 (2018)

    Article  Google Scholar 

  43. Jain, S., Tiso, P., Rutzmoser, J.B., Rixen, D.J.: A quadratic manifold for model order reduction of nonlinear structural dynamics. Comput. Struct. 188, 80–94 (2017)

    Article  Google Scholar 

  44. Rutzmoser, J.B., Rixen, D.J., Tiso, P., Jain, S.: Generalization of quadratic manifolds for reduced order modeling of nonlinear structural dynamics. Comput. Struct. 192, 196–209 (2017)

    Article  Google Scholar 

  45. Boumediene, F., Miloudi, A., Cadou, J., Duigou, L., Boutyour, E.: Nonlinear forced vibration of damped plates by an asymptotic numerical method. Comput. Struct. 87(23–24), 1508–1515 (2009)

    Article  MATH  Google Scholar 

  46. Boumediene, F., Duigou, L., Boutyour, E., Miloudi, A., Cadou, J.: Nonlinear forced vibration of damped plates coupling asymptotic numerical method and reduction models. Comput. Mech. 47(4), 359–377 (2011)

    Article  MATH  Google Scholar 

  47. 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(1), 170–194 (2009)

    Article  Google Scholar 

  48. Lamarque, C.-H., Touzé, C., Thomas, O.: An upper bound for validity limits of asymptotic analytical approaches based on normal form theory. Nonlinear Dyn. 70(3), 1931–1949 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  49. Renson, L., Kerschen, G., Cochelin, B.: Numerical computation of nonlinear normal modes in mechanical engineering. J. Sound Vibr. 364, 177–206 (2016)

    Article  Google Scholar 

  50. Peeters, M., Viguié, R., Sérandour, G., Kerschen, G., Golinval, J.-C.: Nonlinear normal modes, part II: toward a practical computation using numerical continuation techniques. Mech. Syst. Signal Process. 23(1), 195–216 (2009)

    Article  Google Scholar 

  51. Sombroek, C.S.M., Renson, L., Tiso, P., Kerschen, G.: Bridging the gap between nonlinear normal modes and modal derivatives. Nonlinear Dyn. 1, 349–361 (2016)

    Google Scholar 

  52. Kuether, R.J., Allen, M.S.: A numerical approach to directly compute nonlinear normal modes of geometrically nonlinear finite element models. Mech. Syst. Signal Process. 46(1), 1–15 (2014)

    Article  Google Scholar 

  53. Ciarlet, P.G.: A justification of the von-Kármán equations. Arch. Rat. Mech. Analysis 73, 349–389 (1980)

    Article  MATH  Google Scholar 

  54. Millet, O., Hamdouni, A., Cimetière, A.: A classification of thin plate models by asymptotic expansion of non-linear three-dimensional equilibrium equations. Int. J. Non-linear Mech. 36(1), 165–186 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  55. 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)

    Article  MATH  Google Scholar 

  56. Thomas, O., Sénéchal, A., Deü, J.-F.: Hardening/softening behaviour and reduced order modelling of nonlinear vibrations of rotating cantilever beams. Nonlinear Dyn. 86(2), 1293–1318 (2016)

    Article  Google Scholar 

  57. Bilbao, S., Thomas, O., Touzé, C., Ducceschi, M.: Conservative numerical methods for the full von Kármán plate equations 31(6), 1948–1970 (2015). https://doi.org/10.1002/num.21974

  58. Cottanceau, E., Thomas, O., Véron, P., Alochet, M., Deligny, R.: A finite element/quaternion/asymptotic numerical method for the 3D simulation of flexible cables. Finite Elem. Anal. Des. 139, 14–34 (2017)

    Article  MathSciNet  Google Scholar 

  59. Chang, Y., Wang, X., Capiez-Lernout, E., Mignolet, M., Soize, C.: Reduced order modelling for the nonlinear geometric response of some curved structures. In: International Forum on Aeroelasticity and Structural Dynamics, IFASD 2011, AAAF-AIAA, Paper–IFASD (2011)

  60. Grolet, A.: Dynamique non-linéaire des structures mécaniques: application aux systèmes à symétrie cyclique. Ecully, Ecole centrale de Lyon (2013). Ph.D. thesis

  61. Électricité de France: https://www.code-aster.org/forum2/viewtopic.php?id=21537

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

    Article  Google Scholar 

  63. Cochelin, B., Damil, N., Potier-Ferry, M.: Méthode asymptotique numérique. Hermes Lavoissier, Paris (2007)

    MATH  Google Scholar 

  64. Karkar, S., Cochelin, B., Vergez, C.: A high-order, purely frequency based harmonic balance formulation for continuation of periodic solutions: The case of non-polynomial nonlinearities. J. Sound Vibr. 332(4), 968–977 (2013)

    Article  Google Scholar 

  65. Munoz-Almaraz, F., Freire, E., Galán, J., Doedel, E., Vanderbauwhede, A.: Continuation of periodic orbits in conservative and Hamiltonian systems. Phys. D: Nonlinear Phenom. 181(1–2), 1–38 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  66. Manevitch, A., Manevitch, L.: Free oscillations in conservative and dissipative symmetric cubic two-degree-of-freedom systems with closed natural frequencies. Meccanica 38, 335–348 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  67. King, M., Vakakis, A.: Mode localization in a system of coupled flexible beams with geometric nonlinearities. Z. Angew. Math. Mech. 75(2), 127–139 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  68. Grolet, A., Thouverez, F.: Computing multiple periodic solutions of nonlinear vibration problems using the harmonic balance method and Groebner bases. Mech. Syst. Signal Process. 52–53, 529–547 (2015)

    Article  Google Scholar 

  69. Park, C.I.: Frequency equation for the in-plane vibration of a clamped circular plate. J. Sound Vibr. 313(1), 325–333 (2008)

    Article  Google Scholar 

Download references

Acknowledgements

The French Ministry of Research is warmly thanked for the financial support of this study, through the Ph.D. Grant of the first author.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arthur Givois.

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: The beam (u,w) model

The beam model of Eq. (1) is here discussed. Using Hamilton principle, one obtained the classical equations of motion for the axial and transverse motion of the beam, which are:

$$\begin{aligned} \rho S \ddot{w} + EIw_{,xxxx} - \left( N w_{,x}\right) _{,x}&= p(x,t) \end{aligned}$$
(44a)
$$\begin{aligned} \rho S \ddot{u} - N_{,x}&= n(x,t). \end{aligned}$$
(44b)

where p(xt) and n(xt) are distributed transverse and in-plane forces per unit length and where the axial force in the beam is:

$$\begin{aligned} N=ES\left( u_{,x} + \frac{1}{2} w_{,x}^2\right) . \end{aligned}$$
(45)

Equations (45) and (44a, b) constitute the classical von Kármán model describing the transverse motion of a beam with geometrical nonlinearities [12, 15, 16]. Then, eliminating N between Eqs. (45) and (44a, b) leads to the (uw) model of Eq. (1).

After the modal expansion of Sect. 2.1, the analytical expressions of the coefficients of nonlinear terms in Eq. (7a, b) are:

$$\begin{aligned} C_{pi}^k= & {} -\frac{\displaystyle {\int _0^1 (\varPsi _{p,x} \varPhi _{i,x})_{,x} \varPhi _{k} \,\text {d}x}}{\displaystyle {\int _0^1 \varPhi _k^2 \,\text {d}x}}, \nonumber \\ G_{ij}^p= & {} \frac{\displaystyle { \int _0^1 \varPsi _{p} \varPhi _{i,x} \varPhi _{j,xx} \,\text {d}x}}{\displaystyle { \int _0^1 \varPsi _p^2 \,\text {d}x}} \end{aligned}$$
(46a)
$$\begin{aligned} D_{ijl}^k= & {} - \frac{1}{2} \frac{\displaystyle { \int _0^1 (\varPhi _{i,x} \varPhi _{j,x} \varPhi _{l,x})_{,x} \varPhi _k\,\text {d}x}}{\displaystyle { \int _0^1 \varPhi _k^2 \,\text {d}x}}, \end{aligned}$$
(46b)
$$\begin{aligned} Q_k(t)= & {} \frac{\displaystyle { \int _0^1 \varPhi _k(x)p(x,t)\,\text {d}x}}{\displaystyle { \int _0^1 \varPhi _k^2 \text {d}x}},\nonumber \\ E_p(t)= & {} \frac{\displaystyle { \int _0^1 \varPsi _p(x)n(x,t)\,\text {d}x}}{\displaystyle { \int _0^1 \varPsi _p^2 \text {d}x}}. \end{aligned}$$
(46c)

The modes are here normalized as:

$$\begin{aligned} \int _0^1\varPhi _k^2(x)\,\text {d}x=\int _0^1\varPsi _p^2(x)\,\text {d}x=1. \end{aligned}$$
(47)
Table 9 Numerical values, computed with a trapezoidal rule, of some nonlinear coefficients of a the clamped–clamped beam model (Eqs. (46a, b)). Scaling of Eqs. (2) and mode normalization of Eqs. (47)

In the case of a clamped–clamped boundary conditions, the transverse mode shapes are solution of Eq. (6a) and can be written:

$$\begin{aligned} \varPhi _k(x)= & {} \kappa _k\left[ (\sin \beta _k-\sinh \beta _k)(\cos \beta _kx- \cosh \beta _kx)\right. \nonumber \\&\left. -(\cos \beta _k-\cosh \beta _k)(\sin \beta _kx-\sinh \beta _kx)\right] .\nonumber \\ \end{aligned}$$
(48)

with \(\cos \beta _k\cosh \beta _k-1=0,\quad \omega _k=\beta _k^2\) for all \(k>1\) and \(\kappa _k\) is numerically computed to verify Eq. (47). The axial mode shapes are solutions of Eq. (6b) and can be written:

$$\begin{aligned} \varPsi _p(x)=\sqrt{2}\sin \gamma _px,\quad \gamma _p=p\pi . \end{aligned}$$
(49)

They naturally verify Eq. (47). The values of some of these coefficients are given in Table 9.

Appendix B: On the plate \((\varvec{u},w)\)-formulation ROM

Some details about the analytical plate model of Sect. 2.2 are given here. The in-plane forces in the plate are represented by a two-dimensional tensor \(\varvec{N}\), which, with the von Kármán assumptions, can be written [14]:

$$\begin{aligned} \varvec{N}= & {} A\left[ (1-\nu )\varvec{\epsilon }+\nu {\text {tr}}\varvec{\epsilon }\varvec{1}\right] ,\nonumber \\ \varvec{\epsilon }= & {} \frac{1}{2}\left( \varvec{\nabla }\varvec{u}+\varvec{\nabla }^{{\text {T}}}\varvec{u} + \varvec{\nabla }w\otimes \varvec{\nabla }w\right) \end{aligned}$$
(50)

where \(\varvec{\epsilon }\) is the in-plane strain tensor, \(A=Eh/(1-\nu ^2)\) is the in-plane stiffness of the plate, \(\varvec{\nabla }\) denotes the vector/tensor gradients of scalar/vector fields, \({\text {tr}}\) is the trace of a tensor \(\otimes \) is the tensor product of two vectors and \(\varvec{1}\) is the identity tensor.

By eliminating \(\varvec{N}\) between Eqs. (50) and (11a, b), one obtains the following \((\varvec{u},w)\)-formulation:

$$\begin{aligned}&\rho h \ddot{w}+ D \Delta \Delta w {-} A \Big [ (1{-}\nu )\Big (\varvec{\nabla }\varvec{u}:\varvec{\nabla }\varvec{\nabla }w + \varvec{\Delta }\varvec{u}\cdot \varvec{\nabla }w\Big ) \nonumber \\&\quad +\nu \Big ({\mathrm{div}}\,\varvec{u} \,\Delta w + \varvec{\nabla }{\mathrm{div}}\,\varvec{u}\cdot \varvec{\nabla }w\Big ) \nonumber \\&\quad + \varvec{\nabla }w \cdot \varvec{\nabla }\varvec{\nabla }w \cdot \varvec{\nabla }w + \frac{1}{2} (\varvec{\nabla }w)^2 \,\Delta w\Big ] = 0. \end{aligned}$$
(51a)
$$\begin{aligned}&\rho h \ddot{\varvec{u}} - \frac{A}{2}\Big [ (1-\nu ) \varvec{\Delta }\varvec{u} + (1+\nu ) \varvec{\nabla }{\mathrm{div}}\,\varvec{u} \nonumber \\&\quad +(1-\nu ) \Delta w \varvec{\nabla }w + (1+\nu ) \varvec{\nabla }\varvec{\nabla }w \cdot \varvec{\nabla }w \Big ] = 0. \end{aligned}$$
(51b)

where  :  denotes the doubly contracted product of two tensors, \(\varvec{\Delta }\) the vector Laplacian of a vector field and \(\cdot \) the dot product.

With the following dimensionless variables:

$$\begin{aligned} {\bar{w}}= & {} \frac{w}{h} , \quad \bar{\varvec{r}} = \frac{\varvec{r}}{L} \quad {\bar{F}}=\frac{F}{Eh^3}, \nonumber \\ \varepsilon= & {} 12(1-\nu ^2),\quad {\bar{t}}=\frac{1}{L^2}\sqrt{\frac{D}{\rho h}}t,\quad {\bar{p}}= \frac{L^4}{Dh} p, \end{aligned}$$
(52)
$$\begin{aligned} {\tilde{\varepsilon }}= & {} \frac{Ah^2}{D}=\frac{\varepsilon }{1-\nu ^2}=12,\quad \mu = \frac{D}{AL^2},\nonumber \\ \bar{\varvec{u}}= & {} \frac{L}{h^2}\varvec{u},\quad \bar{\varvec{N}}=\frac{L^2}{Eh^3}\varvec{N}. \end{aligned}$$
(53)

where L is a characteristic dimension of the middle plane of the plate (its diameter for instance). The \((\varvec{u},w)\)-formulation is then rewritten:

$$\begin{aligned}&\ddot{w} + \Delta \Delta w - {\tilde{\varepsilon }}\Big [ (1-\nu )\Big (\varvec{\nabla }\varvec{u}:\varvec{\nabla }\varvec{\nabla }w + \varvec{\Delta }\varvec{u}\cdot \varvec{\nabla }w\Big ) \nonumber \\&\quad +\nu \Big ({\mathrm{div}}\,\varvec{u} \,\Delta w + \varvec{\nabla }{\mathrm{div}}\,\varvec{u}\cdot \varvec{\nabla }w\Big ) \nonumber \\&\quad + \varvec{\nabla }w \cdot \varvec{\nabla }\varvec{\nabla }w \cdot \varvec{\nabla }w + \frac{1}{2} (\varvec{\nabla }w)^2 \,\Delta w\Big ] = 0. \end{aligned}$$
(54a)
$$\begin{aligned}&\mu \ddot{\varvec{u}} - \Big [ (1-\nu ) \varvec{\Delta }\varvec{u} + (1+\nu ) \varvec{\nabla }{\mathrm{div}}\,\varvec{u} \nonumber \\&\quad +(1-\nu ) \Delta w \varvec{\nabla }w + (1+\nu ) \varvec{\nabla }\varvec{\nabla }w \cdot \varvec{\nabla }w \Big ] = 0. \end{aligned}$$
(54b)

The bending and in-plane displacements are expanded onto a linear basis of, respectively, \(N_w\) transverse modes \(\varPhi _k\) and \(N_u\) axial modes \(\varvec{\varPsi }_p\):

$$\begin{aligned}&w(\varvec{r},t) = \sum _{k=1}^{N_w} \varPhi _k(\varvec{r}) q_k(t) \quad \text {and} \nonumber \\&\quad \varvec{u}(\varvec{r},t) = \sum _{p=N_w+1}^{N_u+N_w} \varvec{\varPsi }_p(x) \eta _p(t). \end{aligned}$$
(55)

The modes are chosen to satisfy the following two eigenproblems associated with the linear parts of Eqs. (54a, b):

$$\begin{aligned}&\Delta \Delta \varPhi _{k} - \omega ^2_k \varPhi _k = 0 \end{aligned}$$
(56a)
$$\begin{aligned}&(1-\nu )\varvec{\Delta }\varvec{\varPsi }_{p} + (1+\nu )\varvec{\nabla }{\mathrm{div}}\,\varvec{\varPsi }_{p} - \gamma _p^2 \varPsi _p = 0 \end{aligned}$$
(56b)
Table 10 Numerical values of the cubic coefficients \(\varGamma _{ijl}^k\) of Eq. (9), for the clamped–clamped beam [(uw)-formulation, Eq. (10), scaling of Eq. (2) and mode normalization of Eq. (47)] and the clamped circular plate [(wF)-formulation of section C, Eq. (63), scaling of Eqs. (52) and mode normalization of Eqs. (65)]

where \(\omega _k\) and \(\gamma _p\) are the transverse and in-plane dimensionless eigenfrequencies (the frequency equation of the in-plane vibrations of a clamped circular plate has been derived in [69]). Then, by multiplying (54a, b), respectively, by \(\varPhi _k\) and \(\varvec{\varPsi }_p\), integrating on the mid-plane surface \({\mathcal {S}}\) and using the orthogonality properties, the ROM of Eqs. (7a, b) is obtained, with:

$$\begin{aligned} C_{pi}^k= & {} -\frac{\displaystyle {\int _{\mathcal {S}} \Big [(1-\nu )\Big (\varvec{\nabla }\varvec{\varPsi }_p:\varvec{\nabla }\varvec{\nabla }\varPhi _i + \varvec{\Delta }\varvec{\varPsi }_p\cdot \varvec{\nabla }\varPhi _i\Big )+\nu \Big ({\mathrm{div}}\,\varvec{\varPsi }_p \,\Delta \varPhi _i + \varvec{\nabla }{\mathrm{div}}\,\varvec{\varPsi }_p\cdot \varvec{\nabla }\varPhi _i\Big )\Big ]\varPhi _k\,\text {d}S}}{\displaystyle {\int _{\mathcal {S}} \varPhi _k^2 \,\text {d}S}}, \end{aligned}$$
(57a)
$$\begin{aligned} G_{ij}^p= & {} \frac{\displaystyle {\int _{\mathcal {S}} \Big [(1-\nu ) \Delta \varPhi _i \varvec{\nabla }\varPhi _j + (1+\nu ) \varvec{\nabla }\varvec{\nabla }\varPhi _i \cdot \varvec{\nabla }\varPhi _j \Big ]\cdot \varvec{\varPsi }_p \,\text {d}S}}{\displaystyle {\int _{\mathcal {S}} \varvec{\varPsi }_p^2 \,\text {d}S}},\nonumber \\ \end{aligned}$$
(57b)
$$\begin{aligned} D_{ijl}^k= & {} - \frac{\displaystyle {\int _{\mathcal {S}} \Big [\varvec{\nabla }\varPhi _i \cdot \varvec{\nabla }\varvec{\nabla }\varPhi _j \cdot \varvec{\nabla }\varPhi _l + \frac{1}{2} \varvec{\nabla }\varPhi _i\cdot \varvec{\nabla }\varPhi _j \,\Delta \varPhi _l\Big ]\varPhi _k \,\text {d}S}}{\displaystyle {\int _{\mathcal {S}} \varPhi _k^2 \,\text {d}S}}.\nonumber \\ \end{aligned}$$
(57c)

The above developments are not analytically investigated in this study: They have been reported here for a sake of completeness and to precisely demonstrate the results of Sect. 2.3.

Appendix C: On the plate (wF)-formulation ROM

In the literature dealing with analytical models of plates, the \((\varvec{u},w)\)-formulation of the previous section is never used since it would lead to tedious—maybe impossible—computations in an analytical context (see Eq. (51)) and also because there are three unknown displacement fields (w and the two components of \(\varvec{u}\)). In most of the cases, the in-plane inertia is neglected and Eq. (11b) reduces to \({\mathbf {div}}\,\varvec{N}=\varvec{0}\), which enables the introduction of an Airy stress function \(F(\varvec{r},t)\) defined by:

$$\begin{aligned} \Delta F\varvec{1} + \varvec{\nabla }\varvec{\nabla }F = \varvec{N}. \end{aligned}$$
(58)

Then, by adding a compatibility condition, one obtains the following (wF)-formulation (all details can be found in [14]):

$$\begin{aligned} \ddot{w} + \Delta \Delta w&= \varepsilon \phi (w,F) + p(\varvec{r},t) \end{aligned}$$
(59a)
$$\begin{aligned} \Delta \Delta F&= -\frac{1}{2} \phi (w,w) \end{aligned}$$
(59b)

where the overbars have been dropped and where \(\phi (\cdot ,\cdot )\) is a bilinear operator [14]. Similarly to the beam case, the transverse displacements and the force function are expanded on a linear basis of vibration modes \(\varPhi _k\) and \(\varUpsilon _j\):

$$\begin{aligned}&w(\varvec{r},t) = \sum _{k=1}^{N_w} \varPhi _k(\varvec{r}) q_k(t) \quad \text {and} \nonumber \\&\qquad F(\varvec{r},t) = \sum _{j=1}^{N_F} \varUpsilon _j(\varvec{r}) \eta _j(t) \end{aligned}$$
(60)

where the \(q_k\) and \(\eta _j\) are time-dependent modal coordinates, respectively, associated with bending and in-plane eigenmodes, which verify:

$$\begin{aligned} \Delta \Delta \varPhi _{k} - \omega ^2_k \varPhi _k&= 0, \end{aligned}$$
(61a)
$$\begin{aligned} \Delta \Delta \varUpsilon _{j} + \zeta _j^4 \varUpsilon _j&= 0. \end{aligned}$$
(61b)

After multiplying equation (59a) by a mode \(\varPhi _l\), equation (59b) by a mode \(\varUpsilon _m\) and integrating on the surface of the plate, the orthogonality property of the eigenmodes yields:

$$\begin{aligned}&\ddot{q}_k + \omega _k^2 q_k - \varepsilon \sum _{p=1}^{N_w} \sum _{j=1}^{N_F} E_{rj}^k q_r \eta _j = Q_k(t) \end{aligned}$$
(62a)
$$\begin{aligned}&\zeta _j^4 \eta _j = - \frac{1}{2}\sum _{p=1}^{N_w} \sum _{q=1}^{N_w} H_{pq}^j q_p q_q, \end{aligned}$$
(62b)

where nonlinear quadratic coefficients \((E_{rj}^k, H_{pq}^j)\) and modal load \(Q_k\) are defined in “Appendix C”. Finally, eliminating \(\eta _j(t)\) between Eqs (62a, b) leads to the same set of \(N_w\) coupled transverse oscillators with cubic nonlinearities than in the case of beams, Eq. (9), with the following values of the condensed cubic coefficients:

$$\begin{aligned} \varGamma _{pqr}^k = \sum _{j=1}^{N_F} \frac{E_{rj}^kH_{pq}^j}{2\zeta _j^4}. \end{aligned}$$
(63)

Again, the above-defined nonlinear coefficients depend on the mode normalization, defined by Eq. (65).

The analytical expressions of the coefficients of nonlinear terms in Eq. (63) are:

$$\begin{aligned} E_{rj}^k= & {} \frac{\displaystyle {\int _{\mathcal {S}} \varPhi _k L (\varPhi _r, \varUpsilon _j) \,\text {d}S}}{\displaystyle {\int _{\mathcal {S}} \varPhi _k^2 \,\text {d}S}}, \nonumber \\ H_{pq}^j= & {} \frac{\displaystyle {\int _{\mathcal {S}} \varUpsilon _j L(\varPhi _p, \varPhi _q) \,\text {d}S}}{\displaystyle {\int _{\mathcal {S}} \varUpsilon _j^2 \,\text {d}S}}, \nonumber \\ Q_k (t)= & {} \frac{\displaystyle {\int _{\mathcal {S}} \varPhi _{k} p(\varvec{r},t)\,\text {d}S}}{\displaystyle {\int _{\mathcal {S}} \varPhi _k^2 \,\text {d}S}}. \end{aligned}$$
(64)

The modes are here normalized as:

$$\begin{aligned} \int _{\mathcal {S}} \varPhi _k^2\,\text {d}S=\int _{\mathcal {S}} \varUpsilon _p^2\,\text {d}S=1. \end{aligned}$$
(65)

Appendix D: Numerical values of beam and plate cubic ROM coefficients

The numerical values of the \(\varGamma _{ijl}^k\) coefficients of Eq. (9), in the case of the clamped–clamped beam and the circular plate, are gathered in Table 10. They have been computed by the analytical models (Eq. (10) and (63)) and are identical to those computed by the STEP (Sect. 3.2).

Appendix E: Details of STEP

We develop here the details to compute only the nonzero coefficients of (23) with the method of [26]. We refer to the notations \(C_{pi}^k\), \(D_{ijl}^k\) and \(G_{ij}^p\) of Eq. (7). Four distinct steps are necessary, by separating bending modes (BM) and in-plane modes (MM). We also consider unitary modal masses (\(m_r=1\;\forall r\)).

  1. Step 1

    Imposing a displacement on a single BM, \(\varvec{x}_1=\lambda \varvec{\varPhi }_\alpha \), and expanding the result on either a BM \(\varvec{\varPhi }_k\) or a MM \(\varvec{\varPsi }_p\) leads to:

    $$\begin{aligned} \lambda ^3 D_{\alpha \alpha \alpha }^k&= \varvec{\varPhi }_k^{{\text {T}}} \varvec{f}_\text {nl}(\lambda \varvec{\varPhi }_\alpha ), \end{aligned}$$
    (66)
    $$\begin{aligned} \lambda ^2 G_{\alpha \alpha }^p&= \varvec{\varPsi }_p^{{\text {T}}} \varvec{f}_\text {nl}(\lambda \varvec{\varPhi }_\alpha ). \end{aligned}$$
    (67)

    Consequently, only \(N_w\) static computations are necessary to obtain coefficients \(D_{\alpha \alpha \alpha }^k\) and \(G_{\alpha \alpha }^p\).

  2. Step 2

    Imposing displacements on two BM, \(\varvec{x}_2=\lambda _\alpha \varvec{\varPhi }_\alpha \pm \lambda _\beta \varvec{\varPhi }_\beta \), with \(\beta >\alpha \), and expanding the result on either a BM \(\varvec{\varPhi }_k\) or a MM \(\varvec{\varPsi }_p\) leads to:

    $$\begin{aligned} \lambda _\alpha ^2\lambda _\beta D^k_{\alpha \alpha \beta } + \lambda _\alpha \lambda _\beta ^2 D^k_{\alpha \beta \beta } =&\varvec{\varPhi }_k^{{\text {T}}}\varvec{f}_\text {nl}(\lambda _\alpha \varvec{\varPhi }_\alpha + \lambda _\beta \varvec{\varPhi }_\beta ) \nonumber \\&- \lambda _\alpha ^3 D_{\alpha \alpha \alpha }^k - \lambda _\beta ^3 D_{\beta \beta \beta }^k, \end{aligned}$$
    (68)
    $$\begin{aligned} -\lambda _\alpha ^2\lambda _\beta D^k_{\alpha \alpha \beta } + \lambda _\alpha \lambda _\beta ^2 D^k_{\alpha \beta \beta } =&\varvec{\varPhi }_k^{{\text {T}}}\varvec{f}_\text {nl}(\lambda _\alpha \varvec{\varPhi }_\alpha - \lambda _\beta \varvec{\varPhi }_\beta )\nonumber \\&- \lambda _\alpha ^3 D_{\alpha \alpha \alpha }^k + \lambda _\beta ^3 D_{\beta \beta \beta }^k \end{aligned}$$
    (69)
    $$\begin{aligned} \lambda _\alpha \lambda _\beta G_{\alpha \beta }^p =&\varvec{\varPsi }_p^{{\text {T}}}\varvec{f}_\text {nl}(\lambda _\alpha \varvec{\varPhi }_\alpha + \lambda _\beta \varvec{\varPhi }_\beta ) \nonumber \\&- \lambda _\alpha ^2 G_{\alpha \alpha }^p + \lambda _\beta ^2 G_{\beta \beta }^p \end{aligned}$$
    (70)

    where the last part of the second members is known from the first step. At this step, since \(\beta >\alpha \), \(N_w(N_w-1)\) static computations are required.

  3. Step 3

    Imposing displacements on one BM and one MM, \(\varvec{x}_3=\lambda _\alpha \varvec{\varPhi }_\alpha +\lambda _\beta \varvec{\varPsi }_\beta \), and expanding the result on a BM \(\varvec{\varPhi }_k\) leads to:

    $$\begin{aligned} \lambda _\alpha \lambda _\beta C_{\alpha \beta }^k = \varvec{\varPhi }_k^{{\text {T}}}\varvec{f}_\text {nl}(\lambda _\alpha \varvec{\varPhi }_\alpha + \lambda _\beta \varvec{\varPhi }_\beta ) - \lambda _\alpha ^3 D_{\alpha \alpha \alpha }^k \end{aligned}$$
    (71)

    Since \(\alpha =1,\ldots N_w\) and \(\beta =1,\ldots N_u\), the computation of the coefficients \(C_{\alpha \beta }^k\) thus requires \(N_wN_u\) static computations.

  4. Step 4

    Imposing a displacement on a three BM, \(\varvec{x}_4=\lambda _\alpha \varvec{\varPhi }_\alpha + \lambda _\beta \varvec{\varPhi }_\beta + \lambda _\gamma \varvec{\varPhi }_\gamma \), and expanding the result on a BM \(\varvec{\varPhi }_k\) leads to:

    $$\begin{aligned} \lambda _\alpha \lambda _\beta \lambda _\gamma D_{\alpha \beta \gamma }^k&= \varvec{\varPhi }_k^{{\text {T}}}\varvec{f}_\text {nl}(\lambda _\alpha \varvec{\varPhi }_\alpha + \lambda _\beta \varvec{\varPhi }_\beta + \lambda _\gamma \varvec{\varPhi }_\gamma ) \nonumber \\&- \sum _{\begin{array}{c} i,j,l\in \{\alpha ,\beta ,\gamma \}\\ l\ge j\ge i\\ ijl\ne \alpha \beta \gamma \end{array}}\lambda _i\lambda _j\lambda _l D_{ijl}^k. \end{aligned}$$
    (72)

    This computation of the \(D_{\alpha \beta \gamma }^k\) coefficients thus requires \(N_w^3/6-N_w^2/2+N_w/3\) static computations.

Appendix F: Computation of the energy for the FEPs

The nonlinear part of the potential energy in Eq. (41) writes:

$$\begin{aligned} {\mathcal {V}}_{nl}&= \sum _{i=1}^{N _w}\varGamma _{iii}^i \frac{q_i^4}{4} + \sum _{i=1}^{N_w} \sum _{j>i}^{N_w} \bigg (\varGamma _{iij}^i \frac{q_i^3 q_j}{3}+\varGamma _{iij}^{j}q_{i}^2q_{j}^{2} \nonumber \\&\quad + \varGamma _{ijj}^j \frac{q_j^3 q_i}{3} \bigg ) \nonumber \\&\quad + \sum _{i=1}^{N_w} \sum _{j>i}^N \sum _{k>j}^{N_w} \bigg ( \frac{\varGamma _{ijk}^i}{2} q_i^2 q_j q_k + \frac{\varGamma _{ijk}^j}{2} q_i q_j^2 q_k \nonumber \\&\quad + \frac{\varGamma _{ijk}^k}{2} q_i q_j q_k^2 \bigg )\nonumber \\&\quad + \sum _{i=1}^{N_w} \sum _{j>i}^{N_w} \sum _{k>j}^{N_w} \sum _{l=1}^{N_w} \varGamma _{ijk}^l q_i q_j q_k q_l . \end{aligned}$$
(73)

To obtain this equation, the starting point is the general expression of a dynamical system with cubic nonlinearities:

$$\begin{aligned} \ddot{q}_k + \omega _k^2 q_k + \sum _{i=1}^{N_w} \sum _{j=i}^{N_w} \sum _{l=j}^{N_w} \varGamma _{ijl}^{k} q_i q_j q_l = Q_k, \end{aligned}$$
(74)

which can also be written \(\ddot{\varvec{q}} + \varvec{f}(\varvec{q}) = 0\), where \(\varvec{f}\) denotes the internal forces vector. It derives from a potential and can be integrated over time if and only if the following condition on the crossed derivatives is verified:

$$\begin{aligned} \frac{\partial f_i}{\partial q_j} = \frac{\partial f_j}{\partial q_i} \quad \forall i \ne j. \end{aligned}$$
(75)

According to the developments of [14], it is established for some classical boundary conditions of plates—including the clamped one—that the symmetry properties of the bilinear operator L lead to the following equality between the nonlinear coefficients of the (wF)-formulation ROM presented in “Appendix B”:

$$\begin{aligned} H_{pq}^j = E_{pj}^q, \qquad H_{pq}^j = H_{qp}^j. \end{aligned}$$
(76)

It results after some developments to some equalities between the non-upper triangular form coefficients \({\bar{\varGamma }}_{ijl}^k, {i,j,l,k} \in [1,N_w]\), which are written:

$$\begin{aligned} {\bar{\varGamma }}_{ijl}^k {=}{} \mathbf {\bar{\varGamma }}_{jil}^k = {\bar{\varGamma }}_{kjl}^i = {\bar{\varGamma }}_{klj}^i = {\bar{\varGamma }}_{lki}^j {=} {\bar{\varGamma }}_{kli}^j = {\bar{\varGamma }}_{jki}^l = {\bar{\varGamma }}_{kji}^l, \end{aligned}$$
(77)

thus, the upper triangular form yields

$$\begin{aligned}&\varGamma _{iij}^i = 3 \varGamma _{iii}^j, \quad \varGamma _{ijj}^i = \varGamma _{iij}^j,\quad \varGamma _{ijj}^j = 3\varGamma _{jjj}^i \quad \forall j>i \nonumber \\&\varGamma _{ijl}^j = 2 \varGamma _{jjl}^i = 2\varGamma _{ijj}^l, \quad \varGamma _{ijl}^i = 2 \varGamma _{iij}^l = 2\varGamma _{iil}^j, \nonumber \\&\varGamma _{ijl}^l = 2 \varGamma _{ill}^j =2 \varGamma _{jll}^i \quad \forall l>j>i\nonumber \\&\varGamma _{ijl}^k = \varGamma _{ijk}^l = \varGamma _{ilk}^j = \varGamma _{jlk}^i \quad \forall k>l>j>i, \end{aligned}$$
(78)

which allows to verify the condition of Eq. (75). The general expression of the energy in Eq. (73) is then obtained by multiplying the equation of the kth oscillator of Eq. (74) by the modal velocity \({\dot{q}}_k\) and integrating over time.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Givois, A., Grolet, A., Thomas, O. et al. On the frequency response computation of geometrically nonlinear flat structures using reduced-order finite element models. Nonlinear Dyn 97, 1747–1781 (2019). https://doi.org/10.1007/s11071-019-05021-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-019-05021-6

Keywords

Navigation