Skip to main content
Log in

Harmonic balance-based nonsmooth modal analysis of unilaterally constrained discrete systems

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

Abstract

Nonsmooth modal analysis of a unilaterally constrained one-dimensional bar with constant cross-sectional area was recently proposed. The corresponding formulation took advantage of the d’Alembert solution available for such systems and does not require any space semi-discretization of the governing equations. However, it is unable to cope with non-constant cross-sectional area bars, for instance. The present work suggests a formulation relying on various space semi-discretization methodologies (such as finite elements, Rayleigh–Ritz techniques, component mode synthesis, modal superposition and other reduced-order models) where the complementarity Signorini condition, reflecting the unilateral contact constraint, is enforced in a weighted residual sense in time through the harmonic balance method. Importantly, an impact law, such as Newton’s impact law, for instance, classically required for uniqueness purposes in a space semi-discrete dynamical framework, is here explicitly ignored in the proposed formulation and is, instead, implicitly satisfied in a weighted residual sense. As required for the existence of periodic solutions, the predicted vibratory responses would then correspond to an energy-preserving impact law in case the latter had to be explicitly implemented. Periodic responses are investigated in the form of classical energy-frequency backbone curves along with the associated displacement fields. It is found that for the constant cross-section benchmark system, the results compare well with existing works and the proposed methodology stands as a viable option in the field of interest when semi-discretization in space is required.

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
Fig. 21

Similar content being viewed by others

Data availability

The Matlab scripts used to generate the results exposed in this contribution are available at the permalink: 10.5281/zenodo.7577381.

Notes

  1. The following operations are performed starting from Eq. 1: (1) a Fourier transform in time, (2) a semi-discretization in space and (3) a prescribed periodicity in time. It is straightforward to show that this is strictly equivalent to (1) a semi-discretization in space, (2) a prescribed periodicity in time and (3) a Fourier transform. These last two operations form what is commonly known as the harmonic balance method, applied here on a system of Ordinary Differential Equations. This second approach is used in [26, Eq. (3.24)]. The above three operations commute and there is no preferred order.

  2. Note that the participation \({\hat{u}}_{\!N}(\omega )\) of shape function \(\phi _{\!N}(x)\) should not be confused with the discretized displacement \({\hat{u}}_{\!N}(x,\omega )\).

  3. Note that the scalar \(G(\omega )\) should not be confused with the matrix \(\textbf{G}(\omega )\).

  4. The notation \(a_i\), \(i=1,\ldots ,N\) is used instead of \(u_i\) to avoid any possible confusion with the term \(u_{\!N}(x)\) in Eq. B.2. Also the terms \(a_i\) in this section should not be confused with coefficients \(a_k\) in Sect. 3.2.

References

  1. Acary, V., Brogliato, B.: Numerical Methods for Nonsmooth Dynamical Systems: Applications in Mechanics and Electronics (2008)

  2. Bayliss, A., Goldstein, C.I., Turkel, E.: On accuracy conditions for the numerical computation of waves. J. Comput. Phys. 59, 396–404 (1985)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  4. Charleux, D., Gibert, C., Thouverez, F., Dupeux, J.: Numerical and experimental study of friction damping blade attachments of rotating bladed disks. Int. J. Rotating Mach. 2006, 1–13 (2006)

    Article  Google Scholar 

  5. Clarke, F.H.: Generalized gradients and applications. Trans. Am. Math. Soc. 205, 247–262 (1975)

    Article  MathSciNet  Google Scholar 

  6. Craig, R.R., Jr., Bampton, M.C.: Coupling of substructures for dynamic analyses. AIAA J. 6(7), 1313–1319 (1968). ([OAI])

    Article  Google Scholar 

  7. Doyen, D., Ern, A., Piperno, S.: Time-integration schemes for the finite element dynamic Signorini problem. SIAM J. Sci. Comput. 33(1), 223–249 (2011). ([OAI])

    Article  MathSciNet  Google Scholar 

  8. Duchon, C.E.: Lanczos filtering in one and two dimensions. J. Appl. Meteorol. Climatol. 18(8), 1016–1022 (1979)

    Article  Google Scholar 

  9. Hüeber, S., Stadler, G., Wohlmuth, B.I.: A primal-dual active set algorithm for three-dimensional contact problems with coulomb friction. SIAM J. Sci. Comput. 30(2), 572–596 (2008). ([OAI])

    Article  MathSciNet  Google Scholar 

  10. Ihlenburg, F., Babuka, I.: Finite element solution of the Helmholtz equation with high wave number Part I: the h-version of the FEM. Comput. Math. Appl. 30(9), 9–37 (1995)

    Article  MathSciNet  Google Scholar 

  11. Jones, S., Legrand, M.: Forced vibrations of a turbine blade undergoing regularized unilateral contact conditions through the wavelet balance method. Int. J. Numer. Methods Eng. 101(5), 351–374 (2014). ([OAI])

    Article  MathSciNet  Google Scholar 

  12. Jones, S., Legrand, M.: On solving one-dimensional partial differential equations with spatially dependent variables using the wavelet-galerkin method. J. Appl. Mech. 80(6), 061012 (2013)

    Article  Google Scholar 

  13. Legrand, M., Junca, S., Heng, S.: Nonsmooth modal analysis of a N-degree-of-freedom system undergoing a purely elastic impact law. Commun. Nonlinear Sci. Numer. Simul. 45, 190–219 (2017). ([OAI])

    Article  MathSciNet  Google Scholar 

  14. Legrand, M., Pierre, C.: Compact weighted residual formulation for periodic solutions of systems undergoing unilateral contact and frictional occurrences. In: 10th European Nonlinear Oscillations Conference (ENOC 2022). Lyon, France, (2022)

  15. Lu, T., Legrand, M.: Nonsmooth modal analysis via the boundary element method for one-dimensional bar systems. Nonlinear Dyn. 107(1), 227–246 (2022). ([OAI])

    Article  Google Scholar 

  16. Lu, T., Legrand, M.: Nonsmooth modal analysis with boundary element method. In: XI International Conference on Structural Dynamics. Greece, pp. 205–212 (2020)

  17. MacNeal, R.H.: A hybrid method of component mode synthesis. Comput. Struct. 1(4), 581–601 (1971). (Special Issue on Structural Dynamics)

    Article  Google Scholar 

  18. Nacivet, S., Pierre, C., Thouverez, F., Jézéquel, L.: A dynamic Lagrangian frequency-time method for the vibration of dry-friction-damped systems. J. Sound Vib. 265(1), 201–219 (2003). ([OAI])

    Article  MathSciNet  Google Scholar 

  19. Nayfeh, A.H., Balachandran, B.: Applied Nonlinear Dynamics: Analytical, Computational, and Experimental Methods. Wiley, Hoboken (2008)

    Google Scholar 

  20. Nunes, A.W., da Silva, S., Ruiz, A.: Exact general solutions for the mode shapes of longitudinally vibrating nonuniform rods via Lie symmetries. J. Sound Vib. 538, 117216 (2022)

    Article  Google Scholar 

  21. Powell, M.J.: A Fortran subroutine for solving systems of nonlinear algebraic equations (1968)

  22. Rubin, S.: Improved component-mode representation for structural dynamic analysis. AIAA J. 13(8), 995–1006 (1975)

    Article  Google Scholar 

  23. Salles, L., Blanc, L., Thouverez, F., Gouskov, A.M.: Dynamic analysis of fretting-wear in friction contact interfaces. Int. J. Solids Struct. 48(10), 1513–1524 (2011). ([OAI])

    Article  Google Scholar 

  24. Salles, L., Blanc, L., Thouverez, F., Gouskov, A.M.: Dynamic analysis of fretting-wear in friction contact interfaces. J. Eng. Gas Turbines Power 132(1) (2010)

  25. Schreyer, F., Leine, R.I.: A mixed shooting-harmonic balance method for unilaterally constrained mechanical systems. Arch. Mech. Eng. 63(2), 297–314 (2016)

  26. Shi, Y.: Computation of Nonlinear Modes of Vibration of Systems Undergoing Unilateral Contact through the Semi-smooth Newton Approach. M.A. thesis, McGill University, Canada (2016)

  27. Stadler, G.: Semismooth Newton and augmented Lagrangian methods for a simplified friction problem. SIAM J. Optim. 15(1), 39–62 (2004)

    Article  MathSciNet  Google Scholar 

  28. Stewart, D.E.: Dynamics with Inequalities: Impacts and Hard Constraints, vol. 59. SIAM, Philadelphia (2011)

    Book  Google Scholar 

  29. Thorin, A., Delezoide, P., Legrand, M.: Non-smooth modal analysis of piecewise-linear impact oscillators. SIAM J. Appl. Dyn. Syst. 16(3), 1710–1747 (2017). ([OAI])

    Article  MathSciNet  Google Scholar 

  30. Urman, D., Legrand, M., Junca, S.: D’Alembert function for exact non-smooth modal analysis of the bar in unilateral contact. Nonlinear Anal. Hybrid Syst. 43, 101115 (2021)

    Article  MathSciNet  Google Scholar 

  31. Venkatesh, J., Thorin, A., Legrand, M.: Nonlinear modal analysis of a one-dimensional bar undergoing unilateral contact via the time-domain boundary element method. In: ASME 2017 International Design Engineering Technical Conferences. Cleveland, United States (2017)

  32. Yoong, C.: Nonsmooth modal analysis of a finite elastic bar subject to a unilateral contact constraint. Ph.D. thesis, McGill University, Canada (2019)

  33. Yoong, C., Legrand, M.: Nonsmooth modal analysis of a non-internally resonant finite bar subject to a unilateral contact constraint. In: 37th IMAC: A Conference and Exposition on Structural Dynamics, Vol. 1, pp. 1–10. Nonlinear Structures and Systems. USA (2019)

  34. Yoong, C., Thorin, A., Legrand, M.: Nonsmooth modal analysis of an elastic bar subject to a unilateral contact constraint. Nonlinear Dyn. 91, 1–24 (2018)

    Article  Google Scholar 

Download references

Funding

The authors gratefully acknowledge the financial support by the Natural Sciences and Engineering Research Council of Canada through the Discovery Grants Program (421542-2018).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tianzheng Lu.

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

A Periodic solutions of one-dof systems

This appendix briefly discusses the proposed weak enforcement of the Signorini boundary condition for two academic one-dof systems featuring exact solutions in the sense of distributions.

1.1 A.1 Forced one-dof system: the bouncing ball

A rigid mass m subject to gravity, illustrated in Fig. 22, is considered. The (partial) governing equations are:

$$\begin{aligned} m \ddot{u}(t)= & {} -mg +p(t),\quad u(t)\ge 0,\quad p(t)\ge 0,\nonumber \\{} & {} \qquad u(t)p(t)=0, \quad u(0) = u_0, \quad v(0) = 0\nonumber \\ \end{aligned}$$
(A.1)
Fig. 22
figure 22

System of interest: bouncing ball

where u(t) is the displacement of the ball; g stands for gravity and p(t) is the contact force. The unilateral contact condition is supplemented with an impact law reading: for all t such that \(u(t)=0\) with \({\dot{u}}(t-)<0\), then \({\dot{u}}(t+)=-e {\dot{u}}(t-)\) to guarantee solution uniqueness [7]. The restitution coefficient is set to \(e = 1\) to preserve energy and thus existence of periodic solution. Normalization \(\bar{u} = u/2u_0\), \(\bar{t} = t/\sqrt{2u_0/g}\) and \(\bar{p} = p/mg\) yields \(\ddot{u}(t) = -g+p(t)\), \(u(0) = 0.5\), \(v(0) = 0\) and \(g = 1\). The exact solution, shown in Fig. 23a, is periodic with period \(T = 2\) and the solution in the range of \(t\in [0,T]\) reads

$$\begin{aligned} u(t) = \left\{ \begin{aligned}&-0.5t^2+0.5{} & {} 0\le t\le 1\\&-0.5t^2+2t-1.5{} & {} 1< t\le 2 \end{aligned}\right. \end{aligned}$$
(A.2)

along with \(p(t) = \delta (t-1)\) where \(\delta (\cdot )\) is the Dirac delta distribution. The corresponding Fourier series are

$$\begin{aligned}{} & {} u(t) = \frac{1}{3}-\frac{2}{\pi ^2}\sum ^{\infty }_{m=1}\frac{(-1)^m}{m^2}\cos (m\pi t)\nonumber \\{} & {} \quad \text {and}\quad p(t) =1+2\sum ^{\infty }_{m = 1}(-1)^m\cos (m\pi t) \end{aligned}$$
(A.3)

where p(t) is also the time derivative of u(t) twice, termwise. Pointwise convergence is achieved for the displacement and distributional convergence for the contact force even away from the delta.

Fig. 23
figure 23

Time-domain system response: a displacement, b contact force. Exact solution (black lines) and approximation with \(M = 10\) (blue lines), 20 (red lines) and 40 (yellow lines). (Color figure online)

Fig. 24
figure 24

Frequency-domain system response: a displacement, b contact force. Exact solution (black lines) and approximation with \(M = 10\) (blue lines), 20 (red lines) and 40 (yellow lines). (Color figure online)

The methodology detailed in Sects. 3.2 and 3.3 is now used with

$$\begin{aligned}{} & {} u(t) \approx u_{\!M}(t)=\sum ^M_{m = 0}{\hat{u}}_m\cos (m\Omega t),\nonumber \\{} & {} \quad p(t)\approx p_{\!M}(t) =\sum ^M_{m = 0}{\hat{p}}_m\cos (m\Omega t). \end{aligned}$$
(A.4)

The unknowns of the problem are the \(2M+3\) quantities \(\{{\hat{u}}_m,{\hat{p}}_m,\Omega \}\), \(m=0,\ldots ,M\). A Fourier transform on Eq. A.1 leads to the \(M+1\) equations \({\hat{p}}_m = -(m\Omega )^2{\hat{u}}_m\), \(m=1,\ldots ,M\) along with \({\hat{p}}_0 = g \). The weak enforcement of the Signorini condition as developed in the paper leads to the \(M+1\) additional equations

$$\begin{aligned}{} & {} g_m(\hat{\textbf{u}},\hat{\textbf{p}},\Omega )\nonumber \\{} & {} \quad \equiv \frac{1}{T}\int _0^T \cos (m\Omega t)(-p_{\!M}(t)\nonumber \\{} & {} \qquad +\max [-\rho u_{\!M}(t)+p_{\!M}(t),0])\textrm{d} t=0,\nonumber \\{} & {} \quad \qquad m=0,1,\ldots , M, \end{aligned}$$
(A.5)

where \(\hat{\textbf{u}}=({\hat{u}}_0,{\hat{u}}_1,\ldots ,{\hat{u}}_{\!M})\) and \(\hat{\textbf{p}}=({\hat{p}}_0,{\hat{p}}_1,\ldots ,{\hat{p}}_{\!M})\). The last equation is \(\sum ^M_{m = 0}{\hat{u}}_m = u_0\) since the initial displacement of the bouncing ball is prescribed and the frequency of the response is unknown. The initial velocity \(v_0 = 0\) is implicitly enforced since Fourier series are limited to \(\cos \) terms only. The above equations are collectively solved using a nonlinear solver. In practice \(\{\Omega ,\hat{\textbf{u}}\}\) are the sole independent unknowns and Eq. A.5 only should now be solved. We set \(\rho =10\).

The resulting numerically approximated contact force and displacement are shown in Fig. 23. For a low \(M\approx 10\), the approximation exhibits a slightly longer period compared to the exact solution and the instantaneous impact is not accurately captured. For a higher \(M\approx 40\), the approximate contact force converges to Dirac delta function (in a distributional sense). Convergence on the displacement and period is also achieved. Frequency-domain convergence analysis is illustrated in Fig. 24. It is self-explanatory and it is concluded that the proposed solution strategy to enforce a Signorini boundary condition via HBM is able to reproduces, in a weak sense, an energy-preserving impact law.

Fig. 25
figure 25

1-dof spring-mass system with unilateral contact

1.2 A.2 Autonomous one-dof system

Another simple spring-mass example, illustrated in Fig. 25, closer in spirit to the one considered in the present paper is briefly explored. A mass m is attached to a spring k and subject to a unilateral condition with initial gap \(g_0\). Normalized quantities are: time \(\bar{t} = \sqrt{k/m}~t\), displacement \(\bar{u} = u/g_0\) and impact force \(\bar{p} = p/(g_0k)\). The corresponding normalized governing equation reads \(u({t}) + \ddot{u}({t}) = {p}({t})\) where the upper bar notation is dropped. Unilateral contact is governed by the complementarity Signorini condition along with the Newton impact law with \(e = 1\). The main difference with the previous system is the absence of external forcing. Accordingly, the exact solution feature a continuum of (nonlinear) natural frequency \(\Omega \in ]1,2[\) and reads

$$\begin{aligned} u(t) = \frac{1}{\cos (T/2)}\left\{ \begin{aligned}&\cos (t){} & {} 0\le n\le T/2\\&\cos (t-T){} & {} T/2< n\le T \end{aligned}\right. \nonumber \\ \end{aligned}$$
(A.6)

with \(T=2\pi /\Omega \) and \(p(t) = 2\tan (T/2)\delta (T/2)\). The exact solution for \(T = 4.5\) is shown in Fig. 26. The corresponding Fourier series read

$$\begin{aligned} u(t)= & {} \frac{2\tan (T/2)}{T}\Bigl (1 +\sum ^{\infty }_{m = 1}\frac{2(-1)^m}{1-m^2\Omega ^2}\cos (m\Omega t)\Bigr )\quad \text {and}\nonumber \\ p(t)= & {} \frac{2\tan (T/2)}{T}\Bigl (1 +\sum ^{\infty }_{m = 1}2(-1)^m\cos (m\Omega t)\Bigr ). \end{aligned}$$
(A.7)
Fig. 26
figure 26

Time-domain system response: a displacement, b contact force. Exact solution (black lines) and approximation with \(M = 10\) (blue lines), 20 (red lines), and 40 (yellow lines). (Color figure online)

Fig. 27
figure 27

Frequency-domain system response: a displacement, b contact force. Exact solution (black lines) and approximation with \(M = 10\) (blue lines), 20 (red lines), and 40 (yellow lines). (Color figure online)

Fig. 28
figure 28

Convergence analysis: a displacement, b strain. Exact solution (black lines) and LM solution with \(N = 5\) (blue lines) and \(N = 20\) (red lines). (Color figure online)

The governing equations are again solved using the proposed solution strategy with \(T = 4.5\). A total of \(2M+2\) unknowns \(\{\hat{\textbf{u}},\hat{\textbf{p}}\}\) are to be found by solving the \(M+1\) equations \(((1-(m\Omega )^2){\hat{u}}_m = {\hat{p}}_m\), \(m=0,1,2,\ldots ,M\) and the \(M+1\) additional Signorini conditions in the form (A.5) where \(\Omega \), and thus T, is specified.

The displacement and contact force are shown in Fig. 26 for various M (assuming a discretization of the type (A.4)). Pointwise convergence is achieved for the displacement and distributional convergence for the contact force with a rate slightly slower than O(1/M) (not shown). Frequency-domain convergence analysis is illustrated in Fig. 27. Again, the proposed scheme finds approximate periodic solutions satisfying the impact law \(e = 1\) in a weak sense, in agreement with results exposed in [26].

B Linear mode expansion convergence analysis

In this paper, the LM method was developed by considering a family of clamped-free (homogeneous Dirichlet-homogeneous Neumann) modeshapes. It is here shown how the method can generate an approximate solution to a bar problem with homogeneous Dirichlet-non-homogeneous Neumann boundary condition. To this end, we reduce the Helmholtz equation to a time-independent configuration. It the becomes \(u_{,xx}(x) = 0\), \(x\in ]0,1[\), with \(u(0)=0\) and \(u_{x}(1)=1\) (non-homogeneous Neumann BC). The exact solution is \(u(x) = x\), \(x\in [0,1]\). The clamp-free modes of the bar are listed in Eq. 14. Using them to expand the approximation leads to the weak form

$$\begin{aligned} \int _0^1\bigl (\varvec{\phi }_{x}(x)^{\top } \varvec{\phi }_{x}(x)\bigr ) \textrm{d} x ~\textbf{u}=\varvec{\phi }(1) \end{aligned}$$
(B.1)

withFootnote 4\(\textbf{u}\equiv \left[ a_1, \ldots , a_{\!N}\right] ^{\top }\) storing the contributions of the shape functions (ie, the linear modes) collected in \(\varvec{\phi }(x) \equiv \left[ \phi _1(x), \ldots , \phi _{\!N}(x)\right] ^{\top }\). The system of linear equations simplifies to \(\omega _i^2 a_i = (-1)^{i-1}\sqrt{2}\), \(i=1,\ldots ,N\) so that \(a_i = (-1)^{i-1}\sqrt{2}/\omega _i^2\) and the approximated displacement solution of the system reads

$$\begin{aligned} u_{\!N}(x) = \sum _{i = 1}^{N}(-1)^{i-1}\frac{2}{\omega _i^2}\sin (\omega _i x),\quad \omega _i = \frac{2i-1}{2}\pi ,\nonumber \\ \end{aligned}$$
(B.2)

which converges pointwise to the exact solution \(u(x) = x\) as \(N\rightarrow \infty \).

The approximate transfer function defined in the paper reduces here to \(G_{\!N}=u_{\!N}(1)\) since it is assumed that \(p=1\).Accordingly, the convergence of \(u_{\!N}(1)\) as \(N\rightarrow \infty \) is clarified. Equation B.2 yields

$$\begin{aligned} \lim _{N\rightarrow \infty }u_{\!N}(1) =\frac{8}{\pi ^2}\sum _{i=1}^\infty \frac{1}{(2i-1)^2}=1=u(1). \end{aligned}$$
(B.3)

The error caused by truncation at order N is

$$\begin{aligned} E_{\!N} =|u(1)-u_{\!N}(1)|= \frac{8}{\pi ^2}\sum _{i=N+1}^\infty \frac{1}{(2i-1)^2} \end{aligned}$$
(B.4)

with the following bounds:

$$\begin{aligned} \int _{N+1}^\infty \frac{1}{(2x-1)^2}\mathop {}\!\textrm{d}x\le & {} \sum _{i=N+1}^\infty \frac{1}{(2i-1)^2}\nonumber \\\le & {} \int _{N}^\infty \frac{1}{(2x-1)^2}\mathop {}\!\textrm{d}x. \end{aligned}$$
(B.5)

Further evaluation of the upper and lower bounds leads to

$$\begin{aligned} \frac{8}{\pi ^2}\frac{1}{4N+2}<E_{\!N}<\frac{8}{\pi ^2}\frac{1}{4N-2} \end{aligned}$$
(B.6)

indicating that the sequence \(\{u_{\!N}(1)\}\) for N sufficiently large has the rate of convergence O(1/N) and so has \(G_{\!N}\), which matches the derivations reported in Sect. 4.1.

Although \(u_{\!N}(x)\) converges pointwise to the exact solution, see Fig. 28a, the approximated strain

$$\begin{aligned} u_{\!N,x}(x) = 2\sum _{i = 1}^{N}\frac{\cos (\omega _i x)}{\omega _i},\quad \omega _i = \frac{2i-1}{2}\pi \end{aligned}$$
(B.7)

only converges in \(L^2\) sense to \(u_x\). Obviously, the approximated strain vanishes at \(x=1\), that is \(u_{\!N,x}(1)=0\), which is independent of N, and does not converge to the exact strain \(u_{x}(1)=1\). A Gibbs phenomenon is thus generated (here in space) around \(x = 1\). The conclusion is the following: in the approximate solution, the LM method generates a Gibbs phenomenon in space which adds to the Gibbs phenomenon in time induced by HBM. The associated expansion family is clearly not optimal for the problem at hand but convergence is still achieved in a weak sense.

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

Lu, T., Legrand, M. Harmonic balance-based nonsmooth modal analysis of unilaterally constrained discrete systems. Nonlinear Dyn 112, 1619–1640 (2024). https://doi.org/10.1007/s11071-023-09014-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-023-09014-4

Keywords

Navigation