Skip to main content

Comparing Measured and Computed Nonlinear Frequency Responses to Calibrate Nonlinear System Models

  • Conference paper
  • First Online:
Topics in Nonlinear Dynamics, Volume 3

Abstract

Many systems of interest contain nonlinearities that are difficult to accurately model from first principles, so it would be preferable to characterize the system experimentally. For many nonlinear systems, it is now possible to measure frequency response curves with stepped sine testing and to compute frequency response curves with numerical continuation. Nonlinear frequency response curves are very sensitive to the system model and the nonlinearities and they provide a lot of insight into the response of the system to a variety of inputs. This paper explores the feasibility of a nonlinear model updating approach based on nonlinear frequency response and the experimental and analytical tools that are needed. For the experiment, a cantilever beam with an unknown nonlinearity is driven with a harmonic force at various frequencies. The steady-state response is measured and processed with the fast Fourier transform to obtain the frequency response curve. Some subtle yet important details regarding how this is implemented are discussed. An analytical model is also constructed and its frequency response computed using a recently developed technique. The measured and simulated frequencies are then compared and used to tune the analytical model.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 169.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Kim TC et al (2005) Super- and sub-harmonic response calculations for a torsional system with clearance nonlinearity using the harmonic balance method. J Sound Vib 281:965–993

    Article  Google Scholar 

  2. DeSmidt HA (2010) Analysis of a dual gearbox/shaft system with nonlinear dynamic mesh phase interactions. Presented at the 51st AIAA/ASME/ASCE/AHS/ASC structures, structural dynamics and materials conference, Orlando, 12–16 Apr 2010

    Google Scholar 

  3. Padmanabhan C, Singh R (1995) Analysis of periodically excited non-linear systems by a parametric continuation technique. J Sound Vib 184:35–58

    Article  Google Scholar 

  4. Ribeiro P (2004) Non-linear forced vibrations of thin/thick beams and plates by the finite element and shooting methods. Comput Struct 82:1413–1423

    Article  Google Scholar 

  5. Nayfeh AH, Mook DT (1979) Nonlinear oscillations. Wiley, New York

    MATH  Google Scholar 

  6. Gibert C (2003) Fitting measured frequency response using non-linear modes. Mech Syst Signal Process 17:211–218

    Article  Google Scholar 

  7. Sracic MW, Allen MA (2011) Numerical continuation of periodic orbits for harmonically forced nonlinear systems. Presented at the 29th international modal analysis conference (IMAC XXIX), Jacksonville

    Google Scholar 

  8. Peeters M et al (2009) Nonlinear normal modes, part II: Towards a practical computation using numerical continuation techniques. Mech Syst Signal Process 23:195–216

    Article  Google Scholar 

  9. Carrella A, Ewins DJ (2011) Identifying and quantifying structural nonlinearities in engineering applications from measured frequency response functions. Mech Syst Signal Process 25:1011–1027

    Article  Google Scholar 

  10. Imregun M et al (1995) Finite element model updating using frequency response function data: I. Theory and initial investigation. Mech Syst Signal Process 9:187–202

    Article  Google Scholar 

  11. Guckenheimer J, Holmes P (1983) Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, vol 42. Springer, New York

    MATH  Google Scholar 

  12. Kerschen G, Peeters M, Golinval JC, Vakakis AF (2009) Nonlinear normal modes, part I: A useful framework for the structural dynamicist. Mech Syst Signal Process 23:170–194

    Article  Google Scholar 

  13. Sracic MW (2011) A new experimental method for nonlinear system identification based on linear time periodic approximations. Ph.D. Engineering Mechanics, Department of Engineering Physics, University of Wisconsin-Madison, Madison

    Google Scholar 

  14. Sracic MW, Allen MS (2011) Identifying parameters of nonlinear structural dynamic systems using linear time-periodic approximations. Presented at the 29th international modal analysis conference (IMAC XXIX), Jacksonville, 2011

    Google Scholar 

  15. Sracic MW, Allen MS. Identifying parameters of multi-degree-of-freedom nonlinear structural dynamic systems using linear time periodic approximations. Mech Syst Signal Process, submitted July 2011.

    Google Scholar 

  16. Thouverez F (2003) Presentation of the ECL benchmark. Mech Syst Signal Process 17:195–202

    Article  Google Scholar 

  17. Kerschen G, Lenaerts V, Golinval JC (2003) Identification of a continuous structure with a geometrical non-linearity. Part I: Conditioned reverse path method. J Sound Vib 262:889–906

    Article  Google Scholar 

  18. Kerschen G et al (2006) Past, present and future of nonlinear system identification in structural dynamics. Mech Syst Signal Process 20:505–592

    Article  Google Scholar 

  19. Worden K, Tomlinson GR (2007) A review of nonlinear dynamics applications to structural health monitoring. Struct Control Health Monitor 15:540–567

    Article  Google Scholar 

  20. Mayes RL, Gomez AJ (2006) What’s Shakin’, Dude? Effective use of modal shakers. Exp Tech 30:51–61

    Article  Google Scholar 

  21. Ginsberg JH (2001) Mechanical and structural vibrations theory and applications, 1st edn. Wiley, New York

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael W. Sracic .

Editor information

Editors and Affiliations

Appendix A: Ritz-Galerkin Discrete Model

Appendix A: Ritz-Galerkin Discrete Model

A Galerkin approach was used to create a finite-order model of the experimental structure. Assuming that the beam behaves linear-elastically, mode shapes corresponding to transverse bending motion were used as shape functions to construct the Ritz-Galerkin representation [21]. The displacement of the beam at a position x was approximated as

$$ y\left( {x,t} \right) = \sum\limits_{{r = 1}}^{{{N_m}}} \it Psi {}_r(x){q_r}(t) $$
(A.1)

where ψ r (x) is the rth Euler-Bernoulli beam mode shape for a cantilever, q r (t) is the rth generalized coordinate, and N m is the number of modes used. The system’s undamped equations of motion are provided in the following equation, where the coordinates are the amplitudes of the basis functions.

$$ \rho {A_b}L\left[ M \right]\ddot{q} + \frac{{{E_b}I}}{{{L^3}}}\left[ K \right]q = Q = \sum {{f_{{ext}}}{\it Psi_r}({x_f})} $$
(A.2)

For the generalized coordinates, a time derivative is denoted with an over-dot (e.g. the generalized acceleration vector is \( \ddot{q} \)). Modal damping was added to the system by performing an eigenvector analysis on the linear system and then using,

$$ \left[ C \right] = {(\rho {A_b}L)^2}\left[ M \right]\,\left[ {{\phi_b}} \right]\,\left[ {diag\left( {2{\zeta_r}{\omega_r}} \right)} \right]\,{\left[ {{\phi_b}} \right]^{\text{T}}}\,\left[ M \right] $$
(A.3)

where [φ b ] is a matrix containing the eigenvectors in the columns, ω r is the rth circular natural frequency, and ζ r is the rth desired damping ratio. The generalized force vector Q is a sum of the product between all external forces and the value of the shape functions at the point where the force is applied, x f . Therefore, Q includes the applied or external forces, F ext in Fig. 21.5, as well as the nonlinear restoring force due to the spring [21]. The beam provides linear stiffness at the tip due to its flexural rigidity, so the discrete spring’s stiffness was chosen to be purely nonlinear

$$ {k_{{nl}}} = {k_3}y(L)^2 $$
(A.4)

where k 3 is a stiffness constant associated with the nonlinear spring. The physical restoring force due to the spring is then equal to \( {f_{{sp}}} = {k_3}y{(L)^3} \). The generalized force vector then has components corresponding to the nonlinear spring located at x = L and the externally applied force located at x = x f .

$$ \left\{ Q \right\} = {k_3}y(L)^3\left\{ {\begin{array}{*{20}{c}} {{\psi_1}(L)} \\ \vdots \\ {{\psi_N}(L)} \\ \end{array} } \right\} + {A_{{ext}}}\sin \left( {{\omega_T}t} \right)\left\{ {\begin{array}{*{20}{c}} {{\psi_1}({x_f})} \\ \vdots \\ {{\psi_N}({x_f})} \\ \end{array} } \right\} $$
(A.5)

A ext is the amplitude and \( {\omega_T} \) the frequency of the external forcing term that produces the limit cycle.

After using the Ritz-Galerkin method to form the discrete beam model and to account for the nonlinear applied force of the spring, the equations of motion were transformed back into physical coordinates using the relationship in (A.1). The differential equations of motion can then be arranged in state space format.

$$ \begin{gathered} \left\{ {\begin{array}{*{20}{c}} {\dot{y}} \\ {\ddot{y}} \\ \end{array} } \right\} = \left\{ {\begin{array}{*{20}{c}} {\dot{y}} \\ { - {{\left[ {{M_p}} \right]}^{{ - 1}}}\left( {\left[ {{C_p}} \right]\dot{y} + \left[ {{K_p}} \right]y + \left\{ F \right\}} \right)} \\ \end{array} } \right\} \hfill \\ \left[ {{M_p}} \right] = \rho {A_b}L{\left[ \Psi \right]^{{ - {\text{T}}}}}\left[ M \right]{\left[ \Psi \right]^{{ - 1}}},\quad \left[ {{C_p}} \right] = {\left[ \Psi \right]^{{ - {\text{T}}}}}\left[ C \right]{\left[ \Psi \right]^{{ - 1}}}, \hfill \\ \left[ {{K_p}} \right] = \frac{{{E_b}I}}{{{L^3}}}{\left[ \Psi \right]^{{ - {\text{T}}}}}\left[ K \right]{\left[ \Psi \right]^{{ - 1}}},\quad \left\{ F \right\} = {\left[ \Psi \right]^{{ - {\text{T}}}}}\left\{ Q \right\} \hfill \\ \end{gathered} $$
(A.6)

The matrix \( \left[ \Psi \right] \)has the numerical values of the mode vectors for specific position coordinates on the beam. Then, \( \left[ \Psi \right] \)can contain shape functions evaluated at the nodal degrees of freedom on the beam. In this study the number of mode shapes used in the expansion and the number of degrees of measurement points (shown in Fig. 21.5) was N = N m  = 3. The nodes were located at the center and tip of the beam as shown in Fig. 21.5.

Rights and permissions

Reprints and permissions

Copyright information

© 2012 The Society for Experimental Mechanics, Inc. 2012

About this paper

Cite this paper

Sracic, M.W., Yang, S., Allen, M.S. (2012). Comparing Measured and Computed Nonlinear Frequency Responses to Calibrate Nonlinear System Models. In: Adams, D., Kerschen, G., Carrella, A. (eds) Topics in Nonlinear Dynamics, Volume 3. Conference Proceedings of the Society for Experimental Mechanics Series. Springer, New York, NY. https://doi.org/10.1007/978-1-4614-2416-1_21

Download citation

  • DOI: https://doi.org/10.1007/978-1-4614-2416-1_21

  • Published:

  • Publisher Name: Springer, New York, NY

  • Print ISBN: 978-1-4614-2415-4

  • Online ISBN: 978-1-4614-2416-1

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics