Skip to main content
Log in

Input/output reduced model of a damped nonlinear beam based on Volterra series and modal decomposition with convergence results

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

Abstract

This paper addresses the model reduction and the simulation of a damped Euler–Bernoulli–von Kármán pinned beam excited by a distributed force. This nonlinear problem is formulated as a PDE and reformulated as a well-posed state-space system. The model order reduction and simulation are derived by combining two approaches: a Volterra series expansion and truncation and a pseudo-modal truncation defined from the eigenbasis of the linearized problem. The interest of this approach lies in the large class of input waveshapes that can be considered and in the simplicity of the simulation structure. This structure only involves cascades of finite-dimensional decoupled linear systems and multilinear functions. Closed-form bounds depending on the model coefficients and the truncation orders are provided for the Volterra convergence domain and the approximation error. These theoretical results are generalized to a large class of nonlinear models, and refinement of bounds are also proposed for a large sub-class. Numerical experiments confirm that the beam model is well approximated by the very first Volterra terms inside the convergence domain.

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

Notes

  1. Every eigenvalue of A has strictly negative real part.

  2. The maximal real part of the eigenvalues of A is the smallest optimal bound \(\alpha \).

  3. few oscillatory signals, sweeps or sequences of impulses.

  4. see “Appendix A.3” for the detailed construction of the functional setting.

  5. characterized by the smallest absolute value of the real part of \(\lambda _n^\pm \).

  6. This case involves some so-called secular modes due to internal and undamped resonances.

  7. Thanks to filter commutations and using an oversampler (starting in the block-chain at input v or at the linear contribution \(y_1\)), this can be done by simply combining each W with the anti-aliasing filter in the grey blocks.

  8. For recall, \(\partial _t^2 w + 2\big [a I +b {\mathcal {B}}\big ](\partial _t w) + {\mathcal {B}}(w) = f\) on \([0,1] \times {\mathbb {T}}\).

  9. \(H^k(0,1)\) is the standard Sobolev space of functions in \(L^2(0,1)\) whose first k derivatives are in \(L^2(0,1)\).

  10. This justifies the notation introduced in (29) where \(\Vert x_1\Vert _{{\mathbb {H}}^\frac{1}{2}}=\Vert {\mathcal {K}}\;x_1\Vert _{\mathbb {H}}=\Vert x_1^{(2)}\Vert _{\mathbb {H}}\).

References

  1. Antoulas, A.C.: Approximation of large-scale dynamical systems. SIAM, Philadelphia (2005)

    Book  Google Scholar 

  2. Chatterjee, A., Vyas, N.S.: Convergence analysis of Volterra series response of nonlinear systems subjected to harmonic excitation. J. Sound Vib. 236(2), 339–358 (2000). https://doi.org/10.1006/jsvi.2000.2967

    Article  MathSciNet  MATH  Google Scholar 

  3. Chen, K.K., Tu, J.H., Rowley, C.W.: Variants of dynamic mode decomposition: boundary condition, Koopman, and Fourier analyses. J. Nonlinear Sci. 22(6), 887–915 (2012)

    Article  MathSciNet  Google Scholar 

  4. Cheng, C., Peng, Z., Zhang, W., Meng, G.: Volterra-series-based nonlinear system modeling and its engineering applications: a state-of-the-art review. Mech. Syst. Signal Process. 87, 340–364 (2017). https://doi.org/10.1016/j.ymssp.2016.10.029

    Article  Google Scholar 

  5. Cheng, C., Peng, Z., Zhang, W., Meng, G.: Volterra-series-based nonlinear system modeling and its engineering applications: a state-of-the-art review. Mech. Syst. Signal Process. 87, 340–364 (2017)

    Article  Google Scholar 

  6. Crouch, P.E., Collingwood, P.C.: The observation space and realizations of finite Volterra series. SIAM J. Control Optim. 25(2), 316–333 (1987)

    Article  MathSciNet  Google Scholar 

  7. Curtain, R.F., Zwart, H.J.: An Introduction to Infinite-Dimensional Linear Systems Theory. Springer, Berlin (1995)

    Book  Google Scholar 

  8. Fliess, M., Lamnabhi, M., Lamnabhi-Lagarrigue, F.: An algebraic approach to nonlinear functional expansions. IEEE Trans. Circuits Syst. 30(8), 554–570 (1983)

    Article  MathSciNet  Google Scholar 

  9. Géradin, M., Rixen, D.: Théorie des vibrations. Masson (1996)

  10. Gilbert, E.G.: Functional expansions for the response of nonlinear differential systems. IEEE Trans. Autom. Control 22, 909–921 (1977)

    Article  MathSciNet  Google Scholar 

  11. Graff, K.F.: Wave Motion in Elastic Solids. Dover, New York (1991)

    MATH  Google Scholar 

  12. Hassanabadi, M.E., Heidarpour, A., Azam, S.E., Arashpour, M.: Recursive principal component analysis for model order reduction with application in nonlinear bayesian filtering. Comput. Methods Appl. Mech. Eng. 371, 113334 (2020)

    Article  MathSciNet  Google Scholar 

  13. Hélie, T.: Modélisation physique d’instruments de musique et de la voix : systèmes dynamiques, problèmes directs et inverses. Tech. rep., Université Pierre et Marie Curie-Paris 6 (2013). Habilitation à Diriger des Recherches

  14. Hélie, T., Laroche, B.: Computation of convergence bounds for Volterra series of linear analytic single input systems. IEEE Transactions on Automatic Control 56-9, 2062–2072 (2011). http://articles.ircam.fr/textes/Helie10b/

  15. Hélie, T., Laroche, B.: Computable convergence bounds of series expansions for infinite dimensional linear-analytic systems and application. Automatica 50–9, 2334–2340 (2014)

    Article  MathSciNet  Google Scholar 

  16. Jacob, B., Trunk, C., Winklmeier, M.: Analyticity and Riesz basis property of semigroups associated to damped vibrations. J. Evol. Eq. 8(2), 263–281 (2008)

    Article  MathSciNet  Google Scholar 

  17. Kappagantu, R., Feeny, B.: Part 1: Dynamical characterization of a frictionally excited beam. Nonlinear Dyn. 22(4), 317–333 (2000)

    Article  Google Scholar 

  18. Kappagantu, R., Feeny, B.: Part 2: Proper orthogonal modal modeling of a frictionally excited beam. Nonlinear Dyn. 23(1), 1–11 (2000)

    Article  Google Scholar 

  19. Li, L., Billings, S.: Analysis of nonlinear oscillators using Volterra series in the frequency domain Part I : convergence limits. Research report, acse research report no.988. automatic control and systems engineering, University of Sheffield (2009)

  20. Li, L., Billings, S.: Analysis of nonlinear oscillators using Volterra series in the frequency domain. J. Sound Vib. 330(2), 337–355 (2011). https://doi.org/10.1016/j.jsv.2010.08.016

    Article  Google Scholar 

  21. Liang, Y., Lee, H., Lim, S., Lin, W., Lee, K., Wu, C.: Proper orthogonal decomposition and its applications-part i: Theory. J. Sound Vib. 252(3), 527–544 (2002)

    Article  Google Scholar 

  22. Nayfeh, A., Mook, D.: Nonlinear Oscillations. Wiley Classics Library. Wiley (2008). https://books.google.fr/books?id=sj3ebg7jRaoC

  23. Pazy, A.: Semigroups of Linear Operators and Applications to Partial Differential Equations, Applied mathematical sciences, vol. 44. Springer, Berlin (1983)

    MATH  Google Scholar 

  24. Peitz, S., Klus, S.: Koopman operator-based model reduction for switched-system control of pdes. Automatica 106, 184–191 (2019)

  25. Peng, Z., Lang, Z.: On the convergence of the Volterra-series representation of the duffing’s oscillators subjected to harmonic excitations. J. Sound Vib. 305(1), 322–332 (2007). https://doi.org/10.1016/j.jsv.2007.03.062

    Article  MathSciNet  MATH  Google Scholar 

  26. Roze, D., Hélie, T.: Introducing a Green-Volterra series formalism to solve weakly nonlinear boundary problems: application to the Kirchhoff’s string. J. Sound Vib. 333–7, 2073–2086 (2014)

    Article  Google Scholar 

  27. Rugh, W.J.: Nonlinear System Theory. The Johns Hopkins University Press, The Volterra/Wiener approach (1981)

    MATH  Google Scholar 

  28. Schetzen, M.: The Volterra and Wiener theories of nonlinear systems. Wiley-Interscience, New york (1989)

    MATH  Google Scholar 

  29. Volterra, V.: Theory of Functionnals and of Integral and Integro-Differential Equations. Dover Publications, New York (1959)

    Google Scholar 

  30. Zhu, Y.P., Lang, Z.Q.: A new convergence analysis for the Volterra series representation of nonlinear systems. Automatica 111, 108599 (2020)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Hélie.

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

Well-posedness and definitions

1.1 General setting

The class of systems (12) is considered for the general following setting:

  • \({\mathbb {T}}\) denotes the time set \({\mathbb {R}}_+\),

  • \({\mathbb {U}}\) and \({\mathbb {X}}\) are Banach spaces on the field \({\mathbb {R}}\),

  • \({\mathcal {L}}({\mathbb {U}},{\mathbb {X}})\) and \({\mathcal {L}}({\mathbb {X}})\) are the sets of bounded linear operators from \({\mathbb {U}}\) to \({\mathbb {X}}\), and from \({\mathbb {X}}\) to \({\mathbb {X}}\), respectively,

  • \(\mathcal {ML}_k({\mathbb {X}})\) (\(k\ge 2\)) is the set of bounded multilinear operators from \(\underbrace{{\mathbb {X}}\times \dots \times {\mathbb {X}}}_{k}\) to \({\mathbb {X}}\), equipped with norm

    $$\begin{aligned} \displaystyle \Vert A_k\Vert = \sup _{ \underset{\Vert x_1\Vert =\dots =\Vert x_k\Vert =1}{(x_1,\dots ,x_k)\in {\mathbb {X}}^k} } \Vert A_k(x_1,\dots ,x_k)\Vert _{\mathbb {X}}, \end{aligned}$$
  • \({\mathcal {U}}=L^\infty ({\mathbb {T}},{\mathbb {U}})\) and \({\mathcal {X}}=L^\infty ({\mathbb {T}},{\mathbb {X}})\) are standard Lebesgue spaces (in which trajectories u and x live).

1.2 Mild solutions

For the linearized version of systems (12), mild solutions are defined in the following way [7, 23].

Definition 1

(Mild solution of a linear system) Let \(u\in {\mathcal {L}}^\infty _{loc}({\mathbb {T}},{\mathbb {U}})\). The mild solution of the linearized system is the function \(x\in {\mathcal {C}}_0({\mathbb {T}},{\mathbb {X}})\) defined for all \(t\in {\mathbb {T}}\) as

$$\begin{aligned} x(t) = S(t){x_{\mathrm {ini}}}+ \int _0^t S(t-\tau ) B u(\tau ) \, \mathrm {d}\tau . \end{aligned}$$

A similar definition is given for nonlinear systems.

Definition 2

(Mild solution of a nonlinear system) Let \(u\in {\mathcal {L}}^\infty _{loc}({\mathbb {T}},{\mathbb {U}})\). Then, x is said to be a mild solution of (12) iff \(x\in {\mathcal {C}}_0({\mathbb {T}},{\mathbb {X}})\) and satisfies, for all \(t\in {\mathbb {T}},\)

$$\begin{aligned} x(t)&= S(t){x_{\mathrm {ini}}}+\int _0^t S(t-\tau ) \times \nonumber \\&\left( B\,u(\tau ) + \sum _{k=2}^K A_k \big ( x(\tau ),\dots ,x(\tau )\big ) \right) \,\mathrm {d}\tau \end{aligned}$$
(89)

As a standard result from [23], for a fixed \(u\in {\mathcal {C}}_0({\mathbb {T}},{\mathbb {U}})\), there exist \(t_{\max }\in (0,T]\) and a unique function \(x\in {\mathcal {C}}_0([0,t_{\max }),{\mathbb {X}})\) such that x is a mild solution in the sense of (89). For the class of systems (12), it can be easily shown that the local existence and uniqueness of mild solutions still holds when the input u is taken in \({\mathcal {L}}^\infty _{loc}({\mathbb {T}},{\mathbb {U}})\), as stated in Definition 2.

1.3 Functional setting for the linearized beam problem

In the linear problemFootnote 8 (24), the bi-Laplacian \({\mathcal {B}}\) is defined as the unbounded operator on \({\mathbb {H}}=L^2(0,1)\) with domainFootnote 9

$$\begin{aligned} D({\mathcal {B}})&= \big \{ w \in H^4(0,1) ~\text{ s.t. }~ w(0)=w(1)=0,\;\\&\quad w^{\prime \prime }(0)=w^{\prime \prime }(1)=0 \big \}, \end{aligned}$$

such that \({\mathcal {B}}(w)=w^{(4)}\) for all \(w\in D({\mathcal {B}})\).

This operator is closed, densely defined, self-adjoint and positive on \({\mathbb {H}}\). Hence, we can introduce its uniquely defined positive square root \({\mathcal {K}}\), with domain

$$\begin{aligned} D({\mathcal {K}})=\big \{ w \in H^2(0,1) \text{ s.t. } w(0)=w(1)=0\big \}, \end{aligned}$$

such that \({\mathcal {K}}(w)=-\varDelta w\). The domain \(D({\mathcal {K}})\) equipped with the \({\mathcal {K}}\)-norm defines a Hilbert spaceFootnote 10:

$$\begin{aligned} {\mathbb {H}}^{\frac{1}{2}}:= D({\mathcal {K}}) ~~\text {with}~ \Vert .\Vert _{{\mathbb {H}}^\frac{1}{2}}=\Vert {\mathcal {K}}\;.\Vert _{\mathbb {H}}. \end{aligned}$$

Then, according to [16], the linearized problem (24) admits the state-space representation (2527), which is well-posed for the functional setting \({\mathbb {U}}\) and \({\mathbb {X}}\) introduced in (2829) and defining operator A on domain D(A) as

$$\begin{aligned} A :\;&D(A) \longrightarrow {\mathbb {X}}~\text {with}~ D(A) = \{(u,v)\in {\mathbb {H}}^\frac{1}{2}\times {\mathbb {H}}^\frac{1}{2}~\text{ s.t. }~ \nonumber \\&\quad 2(a+b\,{\mathcal {B}})v+{\mathcal {B}}u \in {\mathbb {H}}\}. \end{aligned}$$

In this setting, A generates a \(C_0\) contraction semigroup on \({\mathbb {X}}\) ([16, (A1–A2), p. 6]), so that there exists a negative growth bound \(\alpha <0\). More precisely (see [16, corollary 5.2]), it is a Riesz spectral operator on \({\mathbb {X}}\) which generates an analytic semigroup S, provided that \(-\frac{1}{b}\) is not in the point spectrum of A. In addition, operator B belongs \({\mathcal {L}}({\mathbb {U}},{\mathbb {X}})\) and \(\Vert B\Vert =1\). Therefore, this linear problem is well-posed and is in the class of systems defined in Sect. 2.1.

Estimates for the nonlinear beam model

1.1 Proof of Eq. (32)

For all \(X_1\in {\mathbb {H}}^{\frac{1}{2}}\) and \(z\in [0,1]\), \(X_1(z)=X_1^\prime (0) \, z + \int _0^z (z-\zeta ) X_1^{\prime \prime }(\zeta )\,\mathrm {d}\zeta \) which is zero at \(z=1\). Hence,

$$\begin{aligned} X_1(z)=\int _0^1 K(z,\zeta ) X_1^{\prime \prime }(\zeta )\,\mathrm {d}\zeta , \end{aligned}$$

where \(K(z,\zeta )=\zeta (z-1)\) if \(0\le \zeta \le z\) and \(K(z,\zeta )=z(\zeta -1)\) if \(z\le \zeta \le 1\). Using the Cauchy-Schwartz inequality, it comes

\(\big | X_1(z)\big | \le \mu (z) \Big (\int _0^1 X_1^{\prime \prime }(\zeta )^2\,\mathrm {d}\zeta \Big )^{\frac{1}{2}}\), where \(\mu (z) = \Big (\int _0^1 K(z,\zeta )^2 \,\mathrm {d}\zeta \Big )^{\frac{1}{2}} = \frac{z(1-z)}{\sqrt{3}}\).

Then, \(\Big ( \int _0^1 \big | X_1(z)\big |^2 \, \mathrm {d}z \Big )^{\frac{1}{2}} \le \nu \big \Vert X_1 \Vert _{{\mathbb {H}}^\frac{1}{2}}^2,\) where \(\nu = \Big (\int _0^1 \mu (z)^2 \,\mathrm {d}z\Big )^{\frac{1}{2}} =\frac{1}{3\sqrt{10}}\). Finally, for all \((X,Y,Z)\in {\mathbb {X}}^3\),

$$\begin{aligned}&\big \Vert A_3(X,Y,Z)\Vert _{\mathbb {X}}\\&\quad \le \eta \Big | \int _0^1 X_1^\prime (z)\,Y_1^\prime (z)\, \mathrm {d}z \Big | \; \int _0^1 \Big | Z_1^{\prime \prime }(z) \Big |\, \mathrm {d}z\\&\quad \le \eta \Big |\int _0^1 X_1(z)\,Y_1^{\prime \prime }(z)\, \mathrm {d}z \Big | \; \Big (\int _0^1 \Big |Z_1^{\prime \prime }(z)\Big |^2\, \mathrm {d}z\Big )^{\frac{1}{2}}\\&\quad \le \eta \; \Big ( \int _0^1 \big | X_1(z)\big |^2 \, \mathrm {d}z \Big )^{\frac{1}{2}} \, \big \Vert Y_1 \Vert _{{\mathbb {H}}^\frac{1}{2}} \, \big \Vert Z_1 \Vert _{{\mathbb {H}}^\frac{1}{2}}\\&\quad \le \nu \; \eta \; \big \Vert X_1 \Vert _{{\mathbb {H}}^\frac{1}{2}} \, \big \Vert Y_1 \Vert _{{\mathbb {H}}^\frac{1}{2}} \, \big \Vert Z_1 \Vert _{{\mathbb {H}}^\frac{1}{2}}\\&\quad \le \nu \; \eta \; \big \Vert X \Vert _{{\mathbb {X}}} \, \big \Vert Y \Vert _{{\mathbb {X}}} \, \big \Vert Z \Vert _{{\mathbb {X}}}, \end{aligned}$$

proves the result.

1.2 Proof of Eqs. (33,43)

Functions \(e^\pm _n\) form an orthogonal basis of \({\mathbb {X}}\), from which we define an orthonormal basis

$$\begin{aligned} E_n^+(z) =\frac{1}{k_n^2} \begin{bmatrix} 1 \\ 0 \end{bmatrix}s_n(z) ~~\text{ and }~~ E_n^-(z) = \begin{bmatrix} 0 \\ 1 \end{bmatrix}s_n(z) \end{aligned}$$

On each modal subspace, the linearized system behaves like second order system. Indeed, for all \(x\in {\mathbb {H}}\),

$$\begin{aligned} S(t)Bx=\sum _{n=1}^\infty \kappa _n^-S(t)E_n^-, \end{aligned}$$
(90)

where \(\kappa _n^-=<Bx,E_n^->_{\mathbb {X}}=<x,s_n>_{\mathbb {H}}\), and,

$$\begin{aligned} E_n^+(z)= & {} \frac{(e_n^+-e_n^-)\lambda _n^+\lambda _n^-}{k_n^2(\lambda _n^- - \lambda _n^+)}= \frac{1}{k_n^2}, \begin{bmatrix} 1 \\ 0 \end{bmatrix}s_n(z),\nonumber \\ E_n^-(z)= & {} \frac{(\lambda _n^+e_n^+-\lambda _n^-e_n^-)}{\lambda _n^+-\lambda _n^-}= \begin{bmatrix} 0 \\ 1 \end{bmatrix}s_n(z) \end{aligned}$$
(91)

From (91), we obtain

$$\begin{aligned}&S(t)Bx=\sum _{n=1}^\infty \kappa _n^-\left( \frac{\lambda _n^+e^{\lambda _n^+ t}}{\lambda _n^+ - \lambda _n^-}e_n^+ -\frac{\lambda _n^-e^{\lambda _n^-t}}{\lambda _n^+ - \lambda _n^-}e_n^-\right) =\nonumber \\&\quad \sum _{n=1}^\infty \kappa _n^-\left( k_n^2\frac{e^{\lambda _n^+t}-e^{\lambda _n^- t}}{\lambda _n^+ - \lambda _n^-}E_n^+ + \frac{\lambda _n^+e^{\lambda _n^+ t}-\lambda _n^-e^{\lambda _n^-t}}{\lambda _n^+ - \lambda _n^-}E_n^-\right) . \end{aligned}$$
(92)

Setting \(h_n(t)=\frac{e^{\lambda _n^+ t}-e^{\lambda _n^- t}}{\lambda _n^+ - \lambda _n^-}\) and using the standard expressions

$$\begin{aligned} h_n(t)= \left\{ \begin{array}{ll} e^{-k_n^2\xi _n t}\frac{\sin (k_n^2\sqrt{1-\xi _n^2}t)}{k_n^2\sqrt{1-\xi _n^2}} &{}\text { if }\xi _n<1\\ te^{-k_n^2 t} &{}\text { if }\xi _n=1\\ e^{-k_n^2\xi _n t}\frac{\sinh (k_n^2\sqrt{\xi _n^2-1}t)}{k_n^2\sqrt{\xi _n^2-1}} &{}\text { if }\xi _n>1 \end{array}\right. \end{aligned}$$

we obtain that

$$\begin{aligned} \Vert S(t)Bx\Vert _{\mathbb {X}}&=\sqrt{\sum _{n=1}^\infty (\kappa _n^-)^2 (k_n^4h_n(t)^2+h_n'(t)^2)}\\&\le \sqrt{\sup _n|k_n^4h_n(t)^2+h_n'(t)^2|}\,\Vert x\Vert _{\mathbb {H}}, \end{aligned}$$

so that we can set

$$\begin{aligned} \gamma =\int _{\mathbb {T}}\sqrt{\sup _n|k_n^4h_n(t)^2+h_n'(t)^2|}\mathrm {d}t. \end{aligned}$$
(93)

Now we notice that for all (xyz) in \({\mathbb {X}}\), \(A_3(x,y,z)=Bg(x,y,z)\) where g is a third order multilinear operator from \({\mathbb {X}}^3\) to \({\mathbb {H}}\), such that \(\Vert g\Vert _{\mathcal {ML}({\mathbb {X}},{\mathbb {H}})}=a_3\). We therefore obtain that for all (abc) in \({\mathcal {X}}\)

$$\begin{aligned}&\Vert \int _0^t S(t-\tau )Bg(a,b,c)(\tau ) \mathrm {d}\tau \Vert _{{\mathbb {X}}} \nonumber \\&\quad \le \int _0^t \Vert S(t)Bx\Vert _{{\mathcal {L}}({\mathbb {H}},{\mathbb {X}})}a_3\mathrm {d}\tau \Vert a\Vert _{\mathcal {X}}\Vert b\Vert _{\mathcal {X}}\Vert c\Vert _{\mathcal {X}}\nonumber \\&\quad \le \gamma a_3 \Vert a\Vert _{\mathcal {X}}\Vert b\Vert _{\mathcal {X}}\Vert c\Vert _{\mathcal {X}}. \end{aligned}$$
(94)

This proves that \(\zeta _3\le \gamma a_3\).

In the same way, for operator C, we obtain from (92) that

$$\begin{aligned} CS(t)Bx= \sum _{n=1}^\infty \kappa _n^- \frac{e^{\lambda _n^+ t}-e^{\lambda _n^- t}}{\lambda _n^+ - \lambda _n^-}s_n, \end{aligned}$$

and

$$\begin{aligned} \Vert CS(t)Bx\Vert _{{\mathcal {L}}({\mathbb {H}},{\mathbb {H}}^{\frac{1}{2}})}&=\sqrt{\sum _{n=1}^\infty (\kappa _n^-)^2 k_n^4h_n(t)^2}\\&\le \sup _n\left( k_n^2|h_n(t)|\right) \Vert x\Vert _{\mathbb {H}}. \end{aligned}$$

It follows that \(\Vert CS(t)Bx\Vert _{{\mathcal {L}}({\mathbb {H}},{\mathbb {H}}^{\frac{1}{2}})}\le \sup _n k_n^2|h_n(t)|\), so that we can set

$$\begin{aligned}&\gamma ^\dagger =\int _{\mathbb {T}}\sup _n\left( k_n^2|h_n(t)|\right) \mathrm {d}t,\\&r=\frac{\int _{\mathbb {T}}\sqrt{\sup _n|k_n^2h_n(t)^2+h_n'(t)^2|}\mathrm {d}t}{\int _{\mathbb {T}}\sup _n\left( k_n^2|h_n(t)|\right) \mathrm {d}t}. \end{aligned}$$

Proof of Proposition 1 (truncation error estimate)

We assume (A1–A3) in Sect. 5.3 and set \(\varPhi (z)=\sum _{m=1}^\infty \phi _m z^m\) where the sequence \((\phi _m)_{m\in {\mathbb {N}}^*}\) is defined by (see step 6 in page 4)

$$\begin{aligned} \phi _1\!=\!1 ~~\text{ and, } \text{ for } \text{ all }~ m\ge 2,\quad \phi _m\!=\!\gamma a_3\sum _{p\in {\mathbb {M}}_m^3}\phi _{p_1}\phi _{p_2}\phi _{p_3}, \end{aligned}$$

with \({\mathbb {M}}_m^{3}= \Big \{ {p}\in ({\mathbb {N}}^*)^{3} \,\, \big | \,\,{p}_1+{p}_2+{p}_{3} = m \Big \}\). Then, the trajectories \({\widetilde{x}}=\sum _{m=1}^\infty {\widetilde{x}}_m\) and \(x=\sum _{m=1}^\infty x_m\) are normally convergent and such that, for all \(m\ge 1\),

$$\begin{aligned} \Vert {\widetilde{x}}_m\Vert _{\mathcal {X}}&\le \phi _m \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m,\\ \Vert x_m\Vert _{\mathcal {X}}&\le \phi _m \Vert x_1\Vert _{\mathcal {X}}^m\\&\le \phi _m \left( (1+\epsilon )\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}\right) ^m \le \beta _m \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m, \end{aligned}$$

where the sequence \((\beta _m)_{m\in {\mathbb {N}}^*}\) defined by \(\beta _m=(1+\epsilon )^m\phi _m\) is such that

$$\begin{aligned} \beta _1\!=\!1\!+\!\epsilon ~~\text{ and } \text{ for } \text{ all }~ m\!\ge \! 2,\quad \beta _m\!=\!\gamma a_3\sum _{p\in {\mathbb {M}}_m^3}\beta _{p_1}\beta _{p_2}\beta _{p_3}. \end{aligned}$$

Now, for all \(m\ge 1\), denote the error terms \(e_m := x_m-{\widetilde{x}}_m\). The first terms are such that \(\Vert e_1\Vert _{\mathcal {X}}=\epsilon \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}\) (from (A2)), \(\Vert e_2\Vert _{\mathcal {X}}=0\). Moreover, for \(m\ge 3\), it follows from (9) that

$$\begin{aligned} e_m(t)&\!=\! \int _0^t S(t\!-\!\tau )\sum _{p\in {\mathbb {M}}_m^3} \left( A_3\big (x_{p_1}(\tau ),x_{p_2}(\tau ),x_{p_3}(\tau )\big )\right. \nonumber \\&\quad \left. -{\widehat{A_3}}\big ({\widetilde{x}}_{p_1}(\tau ),{\widetilde{x}}_{p_2}(\tau ),{\widetilde{x}}_{p_3}(\tau )\big )\right) \mathrm {d}\tau \nonumber \\&= \int _0^t S(t-\tau )\sum _{p\in {\mathbb {M}}_m^3}F_p(\tau ) \mathrm {d}\tau , \end{aligned}$$
(95)

where we set \(F_p(\tau )=A_3\big (x_{p_1}(\tau ),x_{p_2}(\tau ),x_{p_3}(\tau )\big ) -A_3\big ({\widetilde{x}}_{p_1}(\tau ),{\widetilde{x}}_{p_2}(\tau ),{\widetilde{x}}_{p_3}(\tau )\big )\). It should be noted that (95) crucially depends on Eq. (53), from which in \({\widehat{{\mathbb {X}}}}\), \({\widehat{A_3}}\) is identical to \(A_3\).

Now, replacing \(x_{p_i}\) by \({\widetilde{x}}_{p_i}+e_{p_i}\) and exploiting the multi-linearity of \(A_3\) yield the following expansion (omitting variable \(\tau \) for sake of legibility)

$$\begin{aligned} F_p&= A_3\big ({\widetilde{x}}_{p_1},{\widetilde{x}}_{p_2},e_{p_3}\big ) +A_3\big ({\widetilde{x}}_{p_1},e_{p_2},{\widetilde{x}}_{p_3}\big )\\&\quad +A_3\big (e_{p_1},{\widetilde{x}}_{p_2},{\widetilde{x}}_{p_3}\big ) \\&\quad +A_3\big ({\widetilde{x}}_{p_1},e_{p_2},e_{p_3}\big ) +A_3\big (e_{p_1},{\widetilde{x}}_{p_2},e_{p_3}\big )\\&\quad +A_3\big (e_{p_1},e_{p_2},{\widetilde{x}}_{p_3}\big ) \\&\quad +A_3\big (e_{p_1},e_{p_2},e_{p_3}\big ). \end{aligned}$$

Then, introducing \(\psi _m=\beta _m-\phi _m\) for all \(m\ge 1\), we prove by induction that (claim \(C_m\)) \(\Vert e_m\Vert _{\mathcal {X}}\le \psi _m\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m\) :

  • (\(m=1\)): the claim (\(C_1\)) is true for \(m=1\) by construction;

  • (\(m\ge 2\)):assume that \(C_p\) holds for all \(p\le m-1\), then using expressions above, it follows that

    $$\begin{aligned}&\Vert e_m\Vert _{\mathcal {X}}\\&\quad \le \gamma a_3 \sum _{p\in {\mathbb {M}}_m^3}\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^{(p_1+p_2+p_3)} \Big ( \phi _{p_1}\phi _{p_2}\psi _{p_3}\\&\qquad +\phi _{p_1}\psi _{p_2}\phi _{p_3}+\psi _{p_1}\phi _{p_2}\phi _{p_3} +\phi _{p_1}\psi _{p_2}\psi _{p_3}\\&\qquad +\psi _{p_1}\phi _{p_2}\psi _{p_3} +\psi _{p_1}\psi _{p_2}\phi _{p_3} +\psi _{p_1}\psi _{p_2}\psi _{p_3} \Big )\quad \\&\quad \le \gamma a_3 \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m \sum _{p\in {\mathbb {M}}_m^3} \Big ( \beta _{p_1}\beta _{p_2}\beta _{p_3}-\phi _{p_1}\phi _{p_2}\phi _{p_3} \big )\\&\quad \le (\beta _m-\phi _m) \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m = \psi _m \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m, \end{aligned}$$

so that \(C_m\) is satisfied. An immediate consequence is that

$$\begin{aligned} \Vert e\Vert _{\mathcal {X}}&\le \sum _{m=1}^\infty \psi _m\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m =\sum _{m=1}^\infty \beta _m\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m-\sum _{m=1}^\infty \phi _m\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m\\&\le \varPhi \left( (1+\epsilon )\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}\right) -\varPhi (\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}), \end{aligned}$$

which concludes the proof.

Proof of generalized result

We assume (A1–A3) in Sect. 5.3 and consider a beam model with a third order nonlinearity \(A_3\) for which \({\widehat{{\mathbb {X}}}}\) is not invariant. We assume that a bound \(a_3\) of \(\Vert A_3\Vert _{{\mathbb {X}}} \) is available, and define \(\varPhi (z)=\sum _{m=1}^\infty \phi _m z^m\) as in “Appendix C”.

We denote \({\widehat{A}}_3=\varPi _{\mathbb {X}}A_3\). Since \(\varPi _{\mathbb {X}}\) is an orthogonal projection, \(\Vert {\widehat{A}}_3\Vert _{{\widehat{{\mathbb {X}}}}}\le \Vert A_3\Vert _{{\mathbb {X}}} \le a_3\), and therefore, as in “Appendix C”, the trajectories \({\widetilde{x}}=\sum _{m=1}^\infty {\widetilde{x}}_m\) and \(x=\sum _{m=1}^\infty x_m\) are convergent and satisfy, for all \(m\ge 1\),

$$\begin{aligned} \Vert {\widetilde{x}}_m\Vert _{\mathcal {X}}&\le \phi _m \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m,\\ \Vert x_m\Vert _{\mathcal {X}}&\le \phi _m \Vert x_1\Vert _{\mathcal {X}}^m\le \phi _m \left( (1+\epsilon )\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}\right) ^m \le \beta _m \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m, \end{aligned}$$

where the sequence \((\beta _m)_{m\in {\mathbb {N}}^*}\) was defined in C. Now, for all \(m\ge 1\), denote the error terms \(e_m := x_m-{\widetilde{x}}_m\). The first terms are such that \(\Vert e_1\Vert _{\mathcal {X}}=\epsilon \Vert {\widehat{x}}_1\Vert _{\mathcal {X}}\) (from (A2)), \(\Vert e_2\Vert _{\mathcal {X}}=0\). Moreover, for \(m\ge 3\), it follows from (9) that

$$\begin{aligned} e_m(t)&= \int _0^t S(t-\tau )\sum _{p\in {\mathbb {M}}_m^3} \left( A_3\big (x_{p_1}(\tau ),x_{p_2}(\tau ),x_{p_3}(\tau )\big )\right. \nonumber \\&\quad \left. -{\widehat{A_3}}\big ({\widetilde{x}}_{p_1}(\tau ),{\widetilde{x}}_{p_2}(\tau ),{\widetilde{x}}_{p_3}(\tau )\big )\right) \mathrm {d}\tau \nonumber \\ \nonumber \\&= \int _0^t S(t-\tau )\sum _{p\in {\mathbb {M}}_m^3}F_p(\tau ) \mathrm {d}\tau \nonumber \\&\quad + \int _0^t S(t-\tau )\sum _{p\in {\mathbb {M}}_m^3}G_p(\tau )\mathrm {d}\tau , \end{aligned}$$
(96)

where we set \(F_p(\tau )=A_3\big (x_{p_1}(\tau ),x_{p_2}(\tau ),x_{p_3}(\tau )\big ) -A_3\big ({\widetilde{x}}_{p_1}(\tau ),{\widetilde{x}}_{p_2}(\tau ),{\widetilde{x}}_{p_3}(\tau )\big )\) and

$$\begin{aligned} G_p(\tau )=(A_3-{\widehat{A_3}})\big ({\widetilde{x}}_{p_1}(\tau ),{\widetilde{x}}_{p_2}(\tau ),{\widetilde{x}}_{p_3}(\tau )\big ) \end{aligned}$$

Setting

$$\begin{aligned} \Vert A_3-{\widehat{A}}_3\Vert _{{\widehat{{\mathbb {X}}}}}= & {} \sup _{\begin{array}{c} (a,b,c)\in ({\widehat{{\mathbb {X}}}})^3\\ \Vert a\Vert _{{\widehat{{\mathbb {X}}}}},\Vert b\Vert _{{\widehat{{\mathbb {X}}}}},\Vert c\Vert _{{\widehat{{\mathbb {X}}}}}=1 \end{array}}\!\!\!\!\!\!\Vert A_3(a,b,c)\\&-{\widehat{A}}_3(a,b,c)\Vert _{{\mathbb {X}}}, \end{aligned}$$

we have

$$\begin{aligned} \Vert G_p\Vert _{{\mathcal {X}}} \le \Vert A_3-{\widehat{A}}_3\Vert _{{\widehat{{\mathbb {X}}}}}\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^{m}\phi _{p_1}\phi _{p_2}\phi _{p_3}. \end{aligned}$$

Then, following the same steps as in “Appendix C”, we introduce \(\psi _m=\beta _m-\phi _m\) and obtain by induction that for all \(m\ge 1\), \(\Vert e_m\Vert _{\mathcal {X}}\le (\psi _m + \frac{\Vert A_3-{\widehat{A}}_3\Vert _{{\widehat{{\mathbb {X}}}}}}{a_3}\phi _m)\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m\). An immediate consequence is that

$$\begin{aligned} \Vert e\Vert _{\mathcal {X}}&\le \sum _{m=1}^\infty (\psi _m+ \frac{\Vert A_3-{\widehat{A}}_3\Vert _{{\widehat{{\mathbb {X}}}}}}{a_3}\phi _m)\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m \\&\le \sum _{m=1}^\infty \beta _m\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m\\&\quad -\left( 1-\frac{\Vert A_3-{\widehat{A}}_3\Vert _{{\widehat{{\mathbb {X}}}}}}{a_3}\right) \sum _{m=1}^\infty \phi _m\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}^m\\&\le \varPhi \left( (1+\epsilon )\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}\right) \\&\quad -\left( 1-\frac{\Vert A_3-{\widehat{A}}_3\Vert _{{\widehat{{\mathbb {X}}}}}}{a_3}\right) \varPhi (\Vert {\widehat{x}}_1\Vert _{\mathcal {X}}), \end{aligned}$$

which concludes the proof.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hélie, T., Laroche, B. Input/output reduced model of a damped nonlinear beam based on Volterra series and modal decomposition with convergence results. Nonlinear Dyn 105, 515–540 (2021). https://doi.org/10.1007/s11071-021-06529-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-021-06529-6

Keywords

Navigation