Skip to main content
Log in

Transient dynamic analysis of cracked structures with multiple contact pairs using generalized HSNC

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

Abstract

The development of efficient computational methods for cracked structures is critical in the fields of civil, mechanical, and aerospace engineering since the influence of cracks on structural dynamics can play an important role in design, prognosis, and health monitoring. The nonlinearity caused by the intermittent contact on the crack surfaces typically excludes the use of fast linear methods, without which the computation of the dynamics of complex cracked structures becomes very challenging. In this paper, an efficient computational scheme for predicting both the transient and steady-state responses of cracked structures with multiple contact pairs is introduced. The new algorithm is referred to as the generalized hybrid symbolic–numeric computational (HSNC) method. The generalized HSNC method extends the original HSNC method, which was recently developed for bilinear systems, to general piecewise-linear nonlinear systems. This work also combines the HSNC with the \(\hbox {X}-\hbox {X}_{r}\) method, a reduced-order modeling technique for cracked structures, to efficiently predict the dynamics of complex structures with cracks. The generalized HSNC approach is based on the idea that the nonlinear response of a cracked structure with multiple contact pairs can be obtained by combining linear responses of the system in each of its linear states. These linear responses can be symbolically expressed as functions of the initial conditions at starting time points in each time range where the system behaves linearly. The transition time where the system switches from one linear state to another is found using a nonlinear optimization solver with the initial values provided by an incremental search process. The method is able to individually track status of each contact pair; therefore, it can be used to predict the dynamics of the system when the crack surfaces are not completely open or closed. Moreover, both the transient and steady-state responses of complex cracked structures under various forcing conditions can be captured by the new method. The generalized HSNC method provides a flexible computational framework that is several orders of magnitude faster than traditional numerical integration methods. The dynamics of a spring–mass system and cantilever beam models that contain one crack and multiple cracks are investigated using the proposed method.

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. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

References

  1. Allemang, R.: Investigation of Some Multiple Input/Output Frequency Response Experimental Modal Analysis Techniques. Ph.D .Thesis, University of Cincinnati, Mechanical Engineering Department (1980)

  2. Bennighof, J.K., Lehoucq, R.B.: An automated multilevel substructuring method for eigenspace computation in linear elastodynamics. SIAM J. Sci. Comput. 25(6), 2084–2106 (2004). https://doi.org/10.1137/S1064827502400650

    Article  MathSciNet  MATH  Google Scholar 

  3. Bovsunovsky, A., Surace, C.: Non-linearities in the vibrations of elastic structures with a closing crack: a state of the art review. Mech. Syst. Signal Process. 62–63, 129–148 (2015). https://doi.org/10.1016/j.ymssp.2015.01.021

    Article  Google Scholar 

  4. Brake, M.: A hybrid approach for the modal analysis of continuous systems with discrete piecewise-linear constraints. J. Sound Vib. 330(13), 3196–3221 (2011). https://doi.org/10.1016/j.jsv.2011.01.028

    Article  Google Scholar 

  5. Brownjohn, J.M.W., De Stefano, A., Xu, Y.L., Wenzel, H., Aktan, A.E.: Vibration-based monitoring of civil infrastructure: challenges and successes. J. Civ. Struct. Health Monit. 1(3), 79–95 (2011). https://doi.org/10.1007/s13349-011-0009-5

    Article  Google Scholar 

  6. Burlayenko, V.N., Sadowski, T.: Influence of skin/core debonding on free vibration behavior of foam and honeycomb cored sandwich plates. Int. J. Non-Linear Mech. 45, 959–968 (2009). https://doi.org/10.1016/j.ijnonlinmec.2009.07.002

    Article  Google Scholar 

  7. Castanier, M.P., Óttarsson, G., Pierre, C.: A reduced order modeling technique for mistuned bladed disks. J. Vib. Acoust. 119(3), 439–447 (1997). https://doi.org/10.1115/1.2889743

    Article  Google Scholar 

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

    Article  MATH  Google Scholar 

  9. Della, C.N., Shu, D.: Vibration of delaminated composite laminates: a review. Appl. Mech. Rev. 60(1), 1–20 (2007). https://doi.org/10.1115/1.2375141

    Article  Google Scholar 

  10. Dimarogonas, A.D.: Vibration of cracked structures: a state of the art review. Eng. Fract. Mech. 55(5), 831–857 (1996)

    Article  Google Scholar 

  11. Doebling, S.W., Farrar, C.R., Prime, M.B.: A summary review of vibration-based damage identification methods. Shock Vib. Dig. 30(2), 91–105 (1998)

    Article  Google Scholar 

  12. Doebling, S.W., Farrar, C.R., Prime, M.B., Shevitz, D.W.: Damage identification and health monitoring of structural and mechanical systems from changes in their vibration characteristics: a literature review. In: Los Alamos National Laboratory Report LA-13070-MS. Los Alamos, NM (1996)

  13. Dormand, J., Prince, P.: A family of embedded Runge–Kutta formulae. J. Comput. Appl. Math. 6(1), 19–26 (1980). https://doi.org/10.1016/0771-050X(80)90013-3

    Article  MathSciNet  MATH  Google Scholar 

  14. D’Souza, K., Epureanu, B.I.: Multiple augmentations of nonlinear systems and generalized minimum rank perturbations for damage detection. J. Sound Vib. 316(1–5), 101–121 (2008). https://doi.org/10.1016/j.jsv.2008.02.018

    Article  Google Scholar 

  15. D’Souza, K., Epureanu, B.I., Pascual, M.: Forecasting bifurcations from large perturbation recoveries in feedback ecosystems. PLOS ONE 10(9), 1–19 (2015). https://doi.org/10.1371/journal.pone.0137779

    Google Scholar 

  16. Ewins, D.J.: Modal Testing: Theory and Practice. Research Studies Press, Taunton (1984)

    Google Scholar 

  17. Friswell, M.I., Penny, J.E.T., Garvey, S.D.: Using linear model reduction to investigate the dynamics of structures with local non-linearities. Mech. Syst. Signal Process. 9(3), 317–328 (1995)

    Article  Google Scholar 

  18. Guyan, R.J.: Reduction of stiffness and mass matrices. AIAA J. 3(2), 380 (1965). https://doi.org/10.2514/3.2874

    Article  Google Scholar 

  19. Hein, H., Feklistova, L.: Computationally efficient delamination detection in composite beams using Haar wavelets. Mech. Syst. Signal Process. 25(6), 2257–2270 (2011). https://doi.org/10.1016/j.ymssp.2011.02.003

    Article  Google Scholar 

  20. Hestenes, M.R.: Multiplier and gradient methods. J. Optim. Theory Appl. 4, 303–320 (1969)

    Article  MathSciNet  MATH  Google Scholar 

  21. Irons, B.: Structural eigenvalue problems: elimination of unwanted variables. AIAA J. 3(5), 961–962 (1965). https://doi.org/10.2514/3.3027

    Article  Google Scholar 

  22. Jaumouill, V., Sinou, J.J., Petitjean, B.: An adaptive harmonic balance method for predicting the nonlinear dynamic responses of mechanical systems–application to bolted structures. J. Sound Vib. 329(19), 4048–4067 (2010). https://doi.org/10.1016/j.jsv.2010.04.008

    Article  Google Scholar 

  23. Jung, C., D’Souza, K., Epureanu, B.I.: Nonlinear amplitude approximation for bilinear systems. J. Sound Vib. 333(13), 2909–19 (2014)

    Article  Google Scholar 

  24. Kim, J.G., Lee, P.S.: An accurate error estimator for guyan reduction. Comput. Methods Appl. Mech. Eng. 278, 1–19 (2014). https://doi.org/10.1016/j.cma.2014.05.002

    Article  MathSciNet  MATH  Google Scholar 

  25. Kim, J.G., Park, Y.J., Lee, G.H., Kim, D.N.: A general model reduction with primal assembly in structural dynamics. Comput. Methods Appl. Mech. Eng. 324, 1–28 (2017). https://doi.org/10.1016/j.cma.2017.06.007

    Article  MathSciNet  Google Scholar 

  26. Kurstak, E., D’Souza, K.: Multistage blisk and large mistuning modeling using fourier constraint modes and PRIME. J. Eng. Gas Turbines Power Trans. ASME 140(7), 072505 (2018). https://doi.org/10.1115/1.4038613

    Article  Google Scholar 

  27. Ma, O., Wang, J.: Model order reduction for impact-contact dynamics simulations of flexible manipulators. Robotica 25(04), 397–407 (2007)

    Article  Google Scholar 

  28. Marinescu, O., Epureanu, B., Banu, M.: Reduced order models of mistuned cracked bladed disks. J. Vib. Acoust. 133(5), 051,014 (2011)

    Article  Google Scholar 

  29. MATLAB: version: R2017b. The MathWorks Inc., Natick, Massachusetts (2017)

  30. Newmark, N.M.: A method of computation for structural dynamics. J. Eng. Mech. 85(EM3), 67–94 (1959)

    Google Scholar 

  31. Poudou, O.: Modeling and analysis of the dynamics of dry-friction-damped structural systems. Ph.D. thesis, The University of Michigan (2007)

  32. Saito, A., Castanier, M.P., Pierre, C., Poudou, O.: Efficient nonlinear vibration analysis of the forced response of rotating cracked blades. J. Comput. Nonlinear Dyn. Trans. ASME 4(1), 011,005 (2009)

    Article  Google Scholar 

  33. Shampine, L.F., Reichelt, M.W.: The matlab ode suite. SIAM J. Sci. Comput. 18(1), 1–22 (1997). https://doi.org/10.1137/S1064827594276424

    Article  MathSciNet  MATH  Google Scholar 

  34. Shiiryayev, O.V., Slater, J.C.: Detection of fatigue cracks using random decrement signatures. Struct. Health Monit. 9(4), 347–360 (2010)

    Article  Google Scholar 

  35. Theodosiou, C., Natsiavas, S.: Dynamics of finite element structural models with multiple unilateral constraints. Int. J. Non-Linear Mech. 44(4), 371–382 (2009). https://doi.org/10.1016/j.ijnonlinmec.2009.01.006

    Article  MATH  Google Scholar 

  36. Tien, M.H., D’Souza, K.: A generalized bilinear amplitude and frequency approximation for piecewise-linear nonlinear systems with gaps or prestress. Nonlinear Dyn. 88(4), 2403–2416 (2017). https://doi.org/10.1007/s11071-017-3385-5

    Article  Google Scholar 

  37. Tien, M.H., D’Souza, K.: Analyzing bilinear systems using a new hybrid symbolic-numeric computational method. J. Vibr. Acoust. 141(3), 031008 (2019). https://doi.org/10.1115/1.4042520

    Article  Google Scholar 

  38. Tien, M.H., Hu, T., D’Souza, K.: Generalized bilinear amplitude approximation and X-Xr for modeling cyclically symmetric structures with cracks. Journal of Vibration and Acoustics 140(4), 041,012–041,012–10 (2018). https://doi.org/10.1115/1.4039296

  39. Tien, M.H., Hu, T., D’Souza, K.: Efficient analysis of cyclic symmetric structures with mistuning and cracks. In: AIAA Scitech 2019 Forum, pp. AIAA 2019–0489 (2019). https://doi.org/10.2514/6.2019-0489

  40. Zhou, T., Xu, J., Sun, Z.: Dynamic analysis and diagnosis of a cracked rotor. J. Vib. Acoust. Trans. ASME 123, 534–539 (2001)

    Article  Google Scholar 

  41. Zucca, S., Epureanu, B.I.: Reduced order models for nonlinear dynamic analysis of structures with intermittent contacts. J. Vib. Control 24(12), 2591–2604 (2018). https://doi.org/10.1177/1077546316689214

    Article  MathSciNet  Google Scholar 

  42. Zucca, S., Firrone, C.M., Gola, M.M.: Numerical assessment of friction damping at turbine blade root joints by simultaneous calculation of the static and dynamic contact loads. Nonlinear Dyn. 67(3), 1943–1955 (2012). https://doi.org/10.1007/s11071-011-0119-y

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kiran D’Souza.

Ethics declarations

Conflict of interest

The authors declared 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: Changing coordinates along crack surfaces

This appendix details how the local coordinates along the crack surfaces are defined and the transformation that is applied to the system matrices.

Fig. 19
figure 13

Normal vector of the pth contact pair with \(N_e^p=4\)

The normal vector of the pth contact pair schematically shown in Fig. 13 is defined as the mean norm of the vectors normal to the adjacent element surfaces [32]:

$$\begin{aligned} \mathbf{n}^{p} = \frac{\sum _{m=1}^{N_{e}^{p}}\mathbf{c}_{m}^{p}}{||\sum _{m=1}^{N_{e}^{p}}{} \mathbf{c}_{m}^{p}||}, \text {where }{} \mathbf{c}_{m}^{p} = \frac{\mathbf{a}_{1}^{m}\times \mathbf{a}_{2}^{m}}{||\mathbf{a}_{1}^{m}\times \mathbf{a}_{2}^{m}||}. \end{aligned}$$
(19)

Note that \(\mathbf{a}_{1}^{m}\) and \(\mathbf{a}_{2}^{m}\) are the vectors from the contacting node to its adjacent nodes on the mth element, and \(N_{e}^{p}\) is the number of elements on the crack surfaces that share the contacting node. For example, \(N_{e}^{p} = 4\) represents the case where the contacting node forms a vertex of four elements as shown in Fig. 13. The contact status of the pth contact pair is tracked along the \(\mathbf{n}^{p}\) direction for \(p=1 \cdots N_{p}\), where \(N_{p}\) is the number of contact pairs in the structure.

Next, a coordinate transformation is applied such that the coordinates of the contacting nodes are aligned along the normal direction. The new local coordinates are defined using three mutually perpendicular unit vector \(\mathbf{n}^{p}\), \(\mathbf{t}_{1}^{p}\), and \(\mathbf{t}_{2}^{p}\), where (\(\mathbf{t}_{1}^{p}\), \(\mathbf{t}_{2}^{p}\)) is an arbitrary set of perpendicular unit vectors tangent to the contacting surface and perpendicular to \(\mathbf{n}^{p}\). The coordinate transformation can be expressed as

$$\begin{aligned} \mathbf{x}_{c1}^{p}= & {} \mathbf{P}_{c}^{p}\hat{\mathbf{x}}_{c1}^{p},\nonumber \\ \mathbf{x}_{c2}^{p}= & {} \mathbf{P}_{c}^{p}\hat{\mathbf{x}}_{c2}^{p}, \end{aligned}$$
(20)

where \(\mathbf{P}_{c}^{p} = \begin{bmatrix}{} \mathbf{n}^{p}&\mathbf{t}_{1}^{p}&\mathbf{t}_{2}^{p}\end{bmatrix}\). The transformation matrix can be assembled into the global form

$$\begin{aligned} \mathbf{x}= \begin{bmatrix} \mathbf{x}_{c1} \\ \mathbf{x}_{c2} \\ \mathbf{x}_{l} \end{bmatrix} = \begin{bmatrix} \mathbf{P}_{c}&\mathbf{0}&\mathbf{0}\\ \mathbf{0}&\mathbf{P}_{c}&\mathbf{0}\\ \mathbf{0}&\mathbf{0}&\mathbf{I} \end{bmatrix} \begin{bmatrix} \hat{\mathbf{x}}_{c1} \\ \hat{\mathbf{x}}_{c2} \\ \mathbf{x}_{l} \end{bmatrix} =\mathbf{P}{\hat{\mathbf{x}}}, \end{aligned}$$
(21)

where \(\mathbf{P}_{c}\) is a block diagonal matrix with submatrices \(\mathbf{P}_{c}^{p}\) on the diagonal. Applying this transformation to Eq. (4) yields

$$\begin{aligned}&\hat{\mathbf{M}}{\ddot{\hat{\mathbf{x}}}}(t)+\hat{\mathbf{C}}{\dot{\hat{\mathbf{x}}}}(t)+\hat{\mathbf{K}}{\hat{\mathbf{x}}}(t) = \hat{\mathbf{A}}_{0}^{*} \nonumber \\&\quad + \sum _{k=1}^{N_k}\big [{\hat{\mathbf{A}}}_{k}\cos (\alpha _{k}{t} + \tau _{k}) + {\hat{\mathbf{B}}}_{k}\sin (\alpha _{k}{t} + \tau _{k})\big ],\nonumber \\ \end{aligned}$$
(22)

where

$$\begin{aligned} \hat{\mathbf{M}}= & {} \mathbf{P}^{T}{} \mathbf{M}{} \mathbf{P} = \begin{bmatrix} \hat{\mathbf{M}}_{c1,c1}&\hat{\mathbf{M}}_{c1,c2}&\hat{\mathbf{M}}_{c1,l}\\ \hat{\mathbf{M}}_{c1,c2}^{T}&\hat{\mathbf{M}}_{c2,c2}&\hat{\mathbf{M}}_{c2,l}\\ \hat{\mathbf{M}}_{c1,l}^{T}&\hat{\mathbf{M}}_{c2,l}^{T}&\mathbf{M}_{l,l} \end{bmatrix},\nonumber \\ \hat{\mathbf{C}}= & {} \mathbf{P}^{T}{} \mathbf{C}{} \mathbf{P} = \begin{bmatrix} \hat{\mathbf{C}}_{c1,c1}&\hat{\mathbf{C}}_{c1,c2}&\hat{\mathbf{C}}_{c1,l}\\ \hat{\mathbf{C}}_{c1,c2}^{T}&\hat{\mathbf{C}}_{c2,c2}&\hat{\mathbf{C}}_{c2,l}\\ \hat{\mathbf{C}}_{c1,l}^{T}&\hat{\mathbf{C}}_{c2,l}^{T}&\mathbf{C}_{l,l} \end{bmatrix},\nonumber \\ \hat{\mathbf{K}}= & {} \mathbf{P}^{T}{} \mathbf{K}{} \mathbf{P} = \begin{bmatrix} \hat{\mathbf{K}}_{c1,c1}&\hat{\mathbf{K}}_{c1,c2}&\hat{\mathbf{K}}_{c1,l}\\ \hat{\mathbf{K}}_{c1,c2}^{T}&\hat{\mathbf{K}}_{c2,c2}&\hat{\mathbf{K}}_{c2,l}\\ \hat{\mathbf{K}}_{c1,l}^{T}&\hat{\mathbf{K}}_{c2,l}^{T}&\mathbf{K}_{l,l} \end{bmatrix},\nonumber \\ \hat{\mathbf{A}}_{0}^{*}= & {} \mathbf{P}^{T}{} \mathbf{A}_{0}^{*}, \nonumber \\ \hat{\mathbf{A}}_{k}= & {} \mathbf{P}^{T}{} \mathbf{A}_{k}, \nonumber \\ \hat{\mathbf{B}}_{k}= & {} \mathbf{P}^{T}{} \mathbf{B}_{k}. \end{aligned}$$
(23)

The submatrices of \(\hat{\mathbf{M}}\) in Eq. (23) can be expressed as

$$\begin{aligned} \hat{\mathbf{M}}_{c1,c1}= & {} \mathbf{P}_{c}^{T}{} \mathbf{M}_{c1,c1} \mathbf{P}_{c}, \nonumber \\ \hat{\mathbf{M}}_{c1,c2}= & {} \mathbf{P}_{c}^{T}{} \mathbf{M}_{c1,c2} \mathbf{P}_{c}, \nonumber \\ \hat{\mathbf{M}}_{c1,l}= & {} \mathbf{P}_{c}^{T}{} \mathbf{M}_{c1,l}, \nonumber \\ \hat{\mathbf{M}}_{c2,c2}= & {} \mathbf{P}_{c}^{T}{} \mathbf{M}_{c2,c2} \mathbf{P}_{c}, \nonumber \\ \hat{\mathbf{M}}_{c2,l}= & {} \mathbf{P}_{c}^{T}{} \mathbf{M}_{c2,l}. \end{aligned}$$
(24)

Note that \(\hat{\mathbf{C}}\) and \(\hat{\mathbf{K}}\) have the same expressions in their submatrices.

Appendix B: \(\hbox {X}-\hbox {X}_{r}\) reduction method

This appendix discusses the creation of a ROM of a cracked structure using the \(\hbox {X}-\hbox {X}_r\) method and CB-CMS [28]. First, the relative coordinate transformation is performed

$$\begin{aligned} \hat{\mathbf{x}}= & {} \begin{bmatrix} \hat{\mathbf{x}}_{c1} \\ \hat{\mathbf{x}}_{c2} \\ \mathbf{x}_{l} \end{bmatrix} ={{\varvec{\alpha }}} \begin{bmatrix} \hat{\mathbf{x}}_{r} \\ \hat{\mathbf{x}}_{c2} \\ \mathbf{x}_{l} \end{bmatrix} ={{\varvec{\alpha }}} \begin{bmatrix} \hat{\mathbf{x}}_{r} \\ \mathbf{x}_{h} \end{bmatrix} ={{\varvec{\alpha }}}{\overline{\mathbf{x}}},\nonumber \\&\text {where } {{\varvec{\alpha }}} = \begin{bmatrix} \mathbf{I}&\mathbf{I}&\mathbf{0}\\ \mathbf{0}&\mathbf{I}&\mathbf{0}\\ \mathbf{0}&\mathbf{0}&\mathbf{I} \end{bmatrix}. \end{aligned}$$
(25)

Note that \(\hat{\mathbf{x}}_{r} = \hat{\mathbf{x}}_{c1} - \hat{\mathbf{x}}_{c2}\) represents the relative displacement between the contact pairs and \(\mathbf{x}_{h}\) = \([\hat{\mathbf{x}}^{T}_{c2},\mathbf{x}^{T}_{l}]^{T}\) represents the coordinates of the healthy structure (i.e., the structure where the crack does not exist). A new set of equations of motion can be obtained by applying the transformation to Eq. (22)

$$\begin{aligned}&\overline{\mathbf{M}}{\ddot{\overline{\mathbf{x}}}}(t) + \overline{\mathbf{C}}{\dot{\overline{\mathbf{x}}}}(t)+\overline{\mathbf{K}}{\overline{\mathbf{x}}}(t) = \overline{\mathbf{A}}_{0}^{*}\nonumber \\&\quad + \sum _{k=1}^{N_k}\big [{\overline{\mathbf{A}}}_{k}\cos (\alpha _{k}{t} + \tau _{k}) + {\overline{\mathbf{B}}}_{k}\sin (\alpha _{k}{t} + \tau _{k})\big ], \end{aligned}$$
(26)

where

$$\begin{aligned} \overline{\mathbf{M}}= & {} {{\varvec{\alpha }}}^{T}\hat{\mathbf{M}}{{\varvec{\alpha }}} = \begin{bmatrix} \overline{\mathbf{M}}_{{\mathbf{x}_{r}}}&\overline{\mathbf{M}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}\\ \overline{\mathbf{M}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}^{T}&\overline{\mathbf{M}}_{{\mathbf{x}_{h}}} \end{bmatrix},\nonumber \\ \overline{\mathbf{C}}= & {} {{\varvec{\alpha }}}^{T}\hat{\mathbf{C}}{{\varvec{\alpha }}} = \begin{bmatrix} \overline{\mathbf{C}}_{{\mathbf{x}_{r}}}&\overline{\mathbf{C}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}\\ \overline{\mathbf{C}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}^{T}&\overline{\mathbf{C}}_{{\mathbf{x}_{h}}} \end{bmatrix},\nonumber \\ \overline{\mathbf{K}}= & {} {{\varvec{\alpha }}}^{T}\hat{\mathbf{K}}{{\varvec{\alpha }}} = \begin{bmatrix} \overline{\mathbf{K}}_{{\mathbf{x}_{r}}}&\overline{\mathbf{K}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}\\ \overline{\mathbf{K}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}^{T}&\overline{\mathbf{K}}_{{\mathbf{x}_{h}}} \end{bmatrix},\nonumber \\ \overline{\mathbf{A}}_{0}^{*}= & {} {{\varvec{\alpha }}}^{T}\hat{\mathbf{A}}_{0}^{*}, \nonumber \\ \overline{\mathbf{A}}_{k}= & {} {{\varvec{\alpha }}}^{T}\hat{\mathbf{A}}_{k}, \nonumber \\ \overline{\mathbf{B}}_{k}= & {} {{\varvec{\alpha }}}^{T}\hat{\mathbf{B}}_{k}. \end{aligned}$$
(27)

Specific expressions for the submatrices of \(\overline{\mathbf{M}}\) in Eq. (27) can be expressed as

$$\begin{aligned}&\overline{\mathbf{M}}_{{\mathbf{x}_{r}}} = \hat{\mathbf{M}}_{c1,c1},\nonumber \\&\overline{\mathbf{M}}_{{{\mathbf{x}_{r}}{\mathbf{x}_{h}}}} = \begin{bmatrix} \hat{\mathbf{M}}_{c1,c1}+\hat{\mathbf{M}}_{c1,c2}~~~&\hat{\mathbf{M}}_{c1,l} \end{bmatrix},\nonumber \\&\overline{\mathbf{M}}_{{\mathbf{x}_{h}}} = \begin{bmatrix} \hat{\mathbf{M}}_{c1,c1}+\hat{\mathbf{M}}_{c1,c2}^{T}+\hat{\mathbf{M}}_{c1,c2}+\hat{\mathbf{M}}_{c2,c2}~&\hat{\mathbf{M}}_{c1,l}+\hat{\mathbf{M}}_{c2,l}\\ \hat{\mathbf{M}}_{c1,l}^{T}+\hat{\mathbf{M}}_{c2,l}^{T}~~~&\mathbf{M}_{l,l} \end{bmatrix}.\nonumber \\ \end{aligned}$$
(28)

Note that \(\overline{\mathbf{C}}\) and \(\overline{\mathbf{K}}\) have the same expressions in their submatrices.

Next, the CB-CMS method is used to reduce the model size by reducing all the DOFs in \(\mathbf{x}_{h}\) using a truncated set of normal modes. The Craig–Bampton transformation matrix \({{\varvec{\beta }}}\) can be expressed as

$$\begin{aligned} {{\varvec{\beta }}} = \begin{bmatrix} \mathbf{I}&\mathbf{0}\\ {{\varvec{\Psi }}}_{c}&{{\varvec{\Phi }}}_{n}\\ \end{bmatrix}, \end{aligned}$$
(29)

where \({{\varvec{\Psi }}_{c}}\) includes the constraint modes and \({{\varvec{\Phi }}}_{n}\) includes the fixed interface normal modes of the healthy structure. The constraint modes are a set of static displacements of the DOFs in \(\mathbf{x}_{h}\) that can be calculated by enforcing successive unit displacements of each relative DOF in \(\mathbf{x}_{r}\) while keeping all other relative DOFs fixed. The normal modes can be obtained by solving for the eigenvectors of (\(\overline{\mathbf{K}}_{\mathbf{x}_{h}}\), \(\overline{\mathbf{M}}_{\mathbf{x}_{h}}\)). The model size is reduced by truncating the normal modes over the frequency range of interest. The coordinates are reduced using the Craig–Bampton transformation

$$\begin{aligned} \overline{\mathbf{x}}= \begin{bmatrix} \hat{\mathbf{x}}_{r} \\ \mathbf{x}_{h} \end{bmatrix} ={{\varvec{\beta }}} \begin{bmatrix} \hat{\mathbf{x}}_{r} \\ {{\varvec{\eta }}} \end{bmatrix} ={{\varvec{\beta }}}{} \mathbf{u}, \end{aligned}$$
(30)

where all the contacting DOFs \(\hat{\mathbf{x}}_{r}\) are kept active in the reduced coordinate system \(\mathbf{u}\). The equations of motion can also be reduced to

$$\begin{aligned}&\mathbf{M}_{\mathrm{ROM}}{\ddot{\mathbf{u}}}(t) + \mathbf{C}_{\mathrm{ROM}}{\dot{\mathbf{u}}}(t) + \mathbf{K}_{\mathrm{ROM}}{\mathbf{u}}(t) = \mathbf{A}_{\mathrm{ROM},0}^{*} \nonumber \\&\quad + \sum _{k=1}^{N_k}\big [{\mathbf{A}}_{\mathrm{ROM},k}\cos (\alpha _{k}{t} + \tau _{k})\nonumber \\&\quad +\, {\mathbf{B}}_{\mathrm{ROM},k}\sin (\alpha _{k}{t} + \tau _{k})\big ], \end{aligned}$$
(31)

where

$$\begin{aligned} \mathbf{M}_{\mathrm{ROM}}= & {} {{\varvec{\beta }}}^{T}\overline{\mathbf{M}}{{\varvec{\beta }}} = \begin{bmatrix} {\mathbf{M}}_{\mathbf{x}_{r}}&{\mathbf{M}}_{\mathbf{x}_{r},{{\varvec{\eta }}}}\\ {\mathbf{M}}_{\mathbf{x}_{r},{{\varvec{\eta }}}}^{T}&{\mathbf{M}}_{{\varvec{\eta }}} \end{bmatrix}, \nonumber \\ \mathbf{C}_{\mathrm{ROM}}= & {} {{\varvec{\beta }}}^{T}\overline{\mathbf{C}}{{\varvec{\beta }}} = \begin{bmatrix} {\mathbf{C}}_{\mathbf{x}_{r}}&{\mathbf{C}}_{\mathbf{x}_{r},{{\varvec{\eta }}}}\\ {\mathbf{C}}_{\mathbf{x}_{r},{{\varvec{\eta }}}}^{T}&{\mathbf{C}}_{{\varvec{\eta }}} \end{bmatrix}, \nonumber \\ \mathbf{K}_{\mathrm{ROM}}= & {} {{\varvec{\beta }}}^{T}\overline{\mathbf{K}}{{\varvec{\beta }}} = \begin{bmatrix} {\mathbf{K}}_{\mathbf{x}_{r}}&{\mathbf{K}}_{\mathbf{x}_{r},{{\varvec{\eta }}}}\\ {\mathbf{K}}_{\mathbf{x}_{r},{{\varvec{\eta }}}}^{T}&{\mathbf{K}}_{{\varvec{\eta }}} \end{bmatrix}, \nonumber \\ \mathbf{A}_{\mathrm{ROM},0}^{*}= & {} {{\varvec{\beta }}}^{T}\overline{\mathbf{A}}_{0}^{*}, \nonumber \\ \mathbf{A}_{\mathrm{ROM},k}= & {} {{\varvec{\beta }}}^{T}\overline{\mathbf{A}}_{k}, \nonumber \\ \mathbf{B}_{\mathrm{ROM},k}= & {} {{\varvec{\beta }}}^{T}\overline{\mathbf{B}}_{k}. \end{aligned}$$
(32)

The explicit expressions for the submatrices of \(\mathbf{M}_{\mathrm{ROM}}\) in Eq. (32) can be expressed as

$$\begin{aligned}&{\mathbf{M}}_{\mathbf{x}_{r}} = \overline{\mathbf{M}}_{{\mathbf{x}_{r}}} + {{\varvec{\Psi }}}_{c}^{T}\overline{\mathbf{M}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}^{T}+ \overline{\mathbf{M}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}{{\varvec{\Psi }}}_{c}+{{\varvec{\Psi }}}_{c}^{T}\overline{\mathbf{M}}_{{\mathbf{x}_{h}}}{{\varvec{\Psi }}}_{c},\nonumber \\&{\mathbf{M}}_{{\mathbf{x}_{r}},{{\varvec{\eta }}}}=\overline{\mathbf{M}}_{{{\mathbf{x}_{r}},{\mathbf{x}_{h}}}}{{\varvec{\Phi }}}_{n}+{{\varvec{\Psi }}}_{c}^{T}\overline{\mathbf{M}}_{{\mathbf{x}_{h}}}{{\varvec{\Phi }}}_{n},\nonumber \\&{\mathbf{M}}_{{\varvec{\eta }}} = {{\varvec{\Phi }}}_{n}^{T}\overline{\mathbf{M}}_{{\mathbf{x}_{h}}}{{\varvec{\Phi }}}_{n}. \end{aligned}$$
(33)

Note that \(\mathbf{C}_{\mathrm{ROM}}\) and \(\mathbf{K}_{\mathrm{ROM}}\) have the same expressions in their submatrices. Moreover, \({\mathbf{M}}_{{\varvec{\eta }}} = {{\varvec{\Phi }}}_{n}^{T}\overline{\mathbf{M}}_{{\mathbf{x}_{h}}}{{\varvec{\Phi }}}_{n} = \mathbf{I}\) and \({\mathbf{K}}_{{\varvec{\eta }}} = {{\varvec{\Phi }}}_{n}^{T}\overline{\mathbf{K}}_{{\mathbf{x}_{h}}}{{\varvec{\Phi }}}_{n} = {{\varvec{\Lambda }}}\), where \({{\varvec{\Lambda }}}\) is the eigenvalue matrix of the healthy structure.

Appendix C: Expressions for modal responses

This appendix gives the solutions for the ordinary differential equation given in Eq. (15) subject to known initial conditions, for both underdamped and overdamped modes.

For underdamped modes, the modal response can be written as

$$\begin{aligned} {q}_{j}^{i}(t)= & {} {D}_{j}^{i} e^{-\zeta _{j}^{i} \omega _{j}^{i}t}\sin (\omega _{d,j}^{i}{t}+{\psi }_{j}^{i})\nonumber \\&+ \sum _{k=1}^{N_k}\big [{X}_{k,j}^{i}{\cos ({\alpha }_{k}{t} + {\tau }_{k}^{i} - {\theta }_{k,j}^{i})}\nonumber \\&+ {Y}_{k,j}^{i}{\sin ({\alpha }_{k}{t} + {\tau }_{k}^{i} - {\theta }_{k,j}^{i})}\big ], \end{aligned}$$
(34)

where \(\omega _{d,j}^{i}\) is the damped frequency corresponding to the natural frequency \(\omega _{j}^{i}\); \(\theta _{k,j}^{i}=\tan ^{-1}\Big [\frac{2\zeta _{j}^{i} \omega _{j}^{i} {\alpha }_{k}}{({\omega }_{j}^{i})^2-{\alpha }_{k}^2}\Big ]\); \({X}_{k,j}^{i}= \frac{G_{k,j}^{i}}{\root \of {[({\omega }_{k,j}^{i})^2 - {\alpha }_{k}^2]^2+(2\zeta _{j}^{i} \omega _{j}^{i} {\alpha }_{k})^2}}\) and \({Y}_{k,j}^{i}= \frac{H_{k,j}^{i}}{\root \of {[({\omega }_{k,j}^{i})^2 - {\alpha }_{k}^2]^2+(2\zeta _{j}^{i} \omega _{j}^{i} {\alpha }_{k})^2}}\) are the steady-state amplitudes; \({D}_{j}^{i}\) is the transient response amplitude; and \({\psi }_{j}^{i}\) is the phase angle of the transient response. Note that \({D}_{j}^{i}\) and \({\psi }_{j}^{i}\) depend on the initial modal positions \({q}_{0,j}^{i}\) and velocity \(\dot{q}_{0,j}^{i}\) that can be calculated using

$$\begin{aligned} {q}_{0,j}^{i}= & {} ({{\varvec{\Phi }}}_{j}^{i})^{T}{} \mathbf{M}_{\mathrm{ROM}}({\tilde{\mathbf{u}}}_{0}^{i} - {\varvec{\delta }}^{i}),\nonumber \\ \dot{q}_{0,j}^{i}= & {} ({{\varvec{\Phi }}}_{j}^{i})^{T}\mathbf{M}_{\mathrm{ROM}}{\dot{\tilde{\mathbf{u}}}}{_{0}^{i}}. \end{aligned}$$
(35)

The explicit expressions for \({D}_{j}^{i}\) and \({\psi }_{j}^{i}\) can be obtained by enforcing the initial conditions in Eq. (34) and its derivative, namely,

$$\begin{aligned} {q}_{j}^{i}(0)= & {} {q}_{0,j}^{i}, \nonumber \\ \dot{q}_{j}^{i}(0)= & {} \dot{q}_{0,j}^{i}. \end{aligned}$$
(36)

These transient parameters can be expressed as

$$\begin{aligned} {\psi }_{j}^{i}= & {} \tan ^{-1}\bigg \{{{\omega }_{d,j}^{i}\Big [{q_{0,j}^{i} - \sum _{k=1}^{N}(X_{k,j}^{i}{{\mathbb {C}}}_{k,j}^{i} + Y_{k,j}^{i}{{\mathbb {S}}}_{k,j}^{i})}\Big ]}\Big / \nonumber \\&\Big [{{\dot{q}}_{0,j}^{i} + {q}_{0,j}^{i}{\zeta }_{j}^{i}{\omega }_{j}^{i} - {\sum _{k=1}^{N}}\Big (({\zeta }_{j}^{i}{\omega }_{j}^{i}{X_{k,j}^{i}} + {\alpha }_{k}{Y_{k,j}^{i}}){{\mathbb {C}}}_{k,j}^{i}}\nonumber \\&+ {({\zeta }_{j}^{i}{\omega }_{j}^{i}{Y_{k,j}^{i}} - {\alpha }_{k}{X_{k,j}^{i}}){{\mathbb {S}}}_{k,j}^{i}\Big )}\Big ]\bigg \}\nonumber \\ {D}_{j}^{i}= & {} \frac{{q}_{0,j}^{i} - \sum _{k=1}^{N}(X_{k,j}^{i}{{\mathbb {C}}}_{k,j}^{i} + Y_{k,j}^{i}{{\mathbb {S}}}_{k,j}^{i})}{\sin {\psi }_{j}^{i}}, \end{aligned}$$
(37)

where

$$\begin{aligned} {{\mathbb {C}}}_{k,j}^{i}= & {} \cos ({\tau }_{k}^{i} - {\theta }_{k,j}^{i}),\nonumber \\ {{\mathbb {S}}}_{k,j}^{i}= & {} \sin ({\tau }_{k}^{i} - {\theta }_{k,j}^{i}). \end{aligned}$$
(38)

For overdamped modes, the modal response can be expressed as

$$\begin{aligned} {q}_{j}^{i}(t)= & {} {E}_{1,j}^{i} e^{s_{1,j}^{i}t} + {E}_{2,j}^{i} e^{s_{2,j}^{i}t} \nonumber \\&+ \sum _{k=1}^{N_k}\big [{X}_{k,j}^{i}{\cos ({\alpha }_{k}{t} + {\tau }_{k}^{i} - {\theta }_{k,j}^{i})}\nonumber \\&+\, {Y}_{k,j}^{i}{\sin ({\alpha }_{k}{t} + {\tau }_{k}^{i} - {\theta }_{k,j}^{i})}\big ], \end{aligned}$$
(39)

where \({s_{1,j}^{i}}\) and \({s_{2,j}^{i}}\) are the roots of the characteristic polynomial of the system that can be expressed as \({s_{1,j}^{i}} = {\omega }_{j}^{i}\big [-\zeta _{j}^{i} +\root \of {({\zeta _{j}^{i}})^2-1}\big ]\) and \({s_{2,j}^{i}} = {\omega }_{j}^{i}\big [-\zeta _{j}^{i} -\root \of {({\zeta _{j}^{i}})^2-1}\big ]\), respectively. \({E}_{1,j}^{i}\) and \({E}_{2,j}^{i}\) are the transient response amplitudes that can also be obtained by enforcing the initial conditions in Eq. (39) and its derivative. These parameters can be expressed as

$$\begin{aligned} {E}_{1,j}^{i}= & {} \frac{-s_{2,j}^{i}{q_{0,j}^{i}} + \dot{q}_{0,j}^{i}}{s_{1,j}^{i} - s_{2,j}^{i}}\nonumber \\&+ \frac{1}{s_{1,j}^{i} - s_{2,j}^{i}} \Bigg \{ {\sum _{k=1}^{N_k}} \big [{\alpha }_{k}({{X}_{k,j}^{i}} {{{\mathbb {S}}}_{k,j}^{i}} - {{Y}_{k,j}^{i}} {{{\mathbb {C}}}_{k,j}^{i}})\big ]\nonumber \\&+\,s_{2,j}^{i} {\sum _{k=1}^{N_k}} ({{X}_{k,j}^{i}} {{{\mathbb {C}}}_{k,j}^{i}} + {{Y}_{k,j}^{i}} {{{\mathbb {S}}}_{k,j}^{i}}) \Bigg \}, \nonumber \\ {E}_{2,j}^{i}= & {} \frac{s_{1,j}^{i}{q_{0,j}^{i}} - \dot{q}_{0,j}^{i}}{s_{1,j}^{i} - s_{2,j}^{i}}\nonumber \\&- \frac{1}{s_{1,j}^{i} - s_{2,j}^{i}} \Bigg \{ {\sum _{k=1}^{N_k}} \big [{\alpha }_{k}({{X}_{k,j}^{i}} {{{\mathbb {S}}}_{k,j}^{i}} - {{Y}_{k,j}^{i}} {{{\mathbb {C}}}_{k,j}^{i}})\big ]\nonumber \\&+\,s_{1,j}^{i} {\sum _{k=1}^{N_k}} ({{X}_{k,j}^{i}} {{{\mathbb {C}}}_{k,j}^{i}} + {{Y}_{k,j}^{i}} {{{\mathbb {S}}}_{k,j}^{i}}) \Bigg \}. \end{aligned}$$
(40)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tien, MH., D’Souza, K. Transient dynamic analysis of cracked structures with multiple contact pairs using generalized HSNC. Nonlinear Dyn 96, 1115–1131 (2019). https://doi.org/10.1007/s11071-019-04844-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-019-04844-7

Keywords

Navigation