Skip to main content
Log in

Spectral collocation methods for the periodic solution of flexible multibody dynamics

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

Abstract

Many flexible multibody systems of practical interest exhibit a periodic response. This paper focuses on the implementation of the collocation version of the Fourier spectral method to determine periodic solutions of flexible multibody systems modeled via the finite element method. To facilitate the analysis and obtain governing equations presenting low-order nonlinearities, the motion formalism is adopted. Application of Fourier spectral methods requires global interpolation schemes that approximate the unknown fields over the entire period of response with exponential convergence characteristics. The classical spectral interpolation schemes were developed for linear fields and hence do not apply to the nonlinear configuration manifolds, such as \(\mathrm {SO}(3)\) or \(\mathrm {SE}(3)\), that are used to describe the kinematics of multibody systems. Furthermore, the configuration and velocity fields are related through nonlinear kinematic compatibility equations. Clearly, special procedures must be developed to adapt the Fourier spectral approach to flexible multibody systems. The spectral interpolation of motion is investigated; interpolation schemes based on the polar decomposition are proposed. Assembly of the linearized governing equations at all the grid points leads to the governing equations of the spectral method. Numerical examples illustrate the performance of the proposed approach.

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

Similar content being viewed by others

References

  1. Ascher, U.M., Mattheij, R.M.M., Russell, R.D.: Numerical Solution of Boundary Value Problems for Ordinary Differential Equations. Society for Industrial and Applied Mathematics, Philadelphia (1995)

    Book  MATH  Google Scholar 

  2. Canuto, C., Hussaini, M.Y., Quarteroni, A., Zang, T.A.: Spectral Methods in Fluid Dynamics. Springer, Berlin, Heidelberg (1988)

    Book  MATH  Google Scholar 

  3. Boyd, J.P.: Chebyshev & Fourier Spectral Methods. Springer, Berlin, Heidelberg (1989)

    Book  MATH  Google Scholar 

  4. Hesthaven, J.S., Gottlieb, S., Gottlieb, D.: Spectral Methods for Time-Dependent Problems. Cambridge University Press, Cambridge (2007)

    Book  MATH  Google Scholar 

  5. Krylov, N.M., Bogoliubov, N.N., Lefschetz, S.: Introduction to Nonlinear Mechanics. Princeton University Press, Princeton (1947)

    Google Scholar 

  6. Hayashi, C.: Nonlinear Oscillations in Physical Systems. Princeton University Press, Princeton (1964)

    MATH  Google Scholar 

  7. Nayfeh, A.H., Mook, D.T.: Nonlinear Oscillations. Wiley, New York (1979)

    MATH  Google Scholar 

  8. Kundert, K.S., White, J.K., Sangiovanni-Vincentelli, A.: Steady-State Methods for Simulating Analog and Microwave Circuits. Kluwer Academic Publishers, Boston, Dordrecht, London (1990)

    Book  MATH  Google Scholar 

  9. Cardona, A., Coune, T., Lerusse, A., Géradin, M.: A multiharmonic method for non-linear vibration analysis. Int. J. Numer. Methods Eng. 37(9), 1593–1608 (1994)

    Article  MATH  Google Scholar 

  10. Jaumouillé, 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)

    Article  Google Scholar 

  11. LaBryer, A., Attar, P.J.: A harmonic balance approach for large-scale problems in nonlinear structural dynamics. Comput. Struct. 88(17–18), 1002–1014 (2010)

    Article  Google Scholar 

  12. Detroux, T., Renson, L., Masset, L., Kerschen, G.: The harmonic balance method for bifurcation analysis of large-scale nonlinear mechanical systems. Comput. Methods Appl. Mech. Eng. 296, 18–38 (2015)

    Article  MathSciNet  Google Scholar 

  13. Von Groll, G., Ewins, D.J.: The harmonic balance method with arc-length continuation in rotor/stator contact problems. J. Sound Vib. 241(2), 223–233 (2001)

    Article  Google Scholar 

  14. Sanliturk, K.Y., Ewins, D.J.: Modeling two-dimensional friction contact and its application using harmonic balance method. J. Sound Vib. 193, 511–523 (1996)

    Article  MATH  Google Scholar 

  15. Hall, K.C., Thomas, J.P., Clark, W.S.: Computation of unsteady nonlinear flows in cascades using a harmonic balance technique. AIAA J. 40, 879–886 (2002)

    Article  Google Scholar 

  16. Thomas, J.P., Dowell, E.H., Hall, K.C.: Modeling viscous transonic limit cycle oscillation behavior using a harmonic balance approach. J. Aircr. 41, 1266–1274 (2004)

    Article  Google Scholar 

  17. Custer, C.H., Thomas, J.P., Dowell, E.H., Hall, K.C.: A nonlinear harmonic balance method for the CFD code OVERFLOW 2. In: Interational Forum on Aeroelasticity and Structural Dynamics (IFASD), Paper 2009-050, Seattle, WA, 7–10 June 2009

  18. Borri, M., Bottasso, C.L.: An intrinsic beam model based on a helicoidal approximation. Part I: formulation. Part II: linearization and finite element implementation. Int. J. Numer. Methods Eng. 37, 2267–2309 (1994)

    Article  MATH  Google Scholar 

  19. Merlini, T., Morandini, M.: The helicoidal modeling in computational finite elasticity. Part I: variational formulation. Part II: multiplicative interpolation. Int. J. Solids Struct. 41(18–19), 5351–5409 (2004)

    Article  MATH  Google Scholar 

  20. Sonneville, V., Cardona, A., Brüls, O.: Geometrically exact beam finite element formulated on the special Euclidean group SE(3). Comput. Methods Appl. Mech. Eng. 268(1), 451–474 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  21. Sonneville, V., Brüls, O.: A formulation on the special Euclidean group for dynamic analysis of multibody systems. J. Comput. Nonlinear Dyn. 9(4), 04002 (2014)

    MATH  Google Scholar 

  22. Weideman, J.A.C.: Numerical integration of periodic functions: a few examples. Am. Math. Mon. 1, 21–36 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  23. Trefethen, L.N., Weideman, J.A.C.: The exponentially convergent trapezoidal rule. SIAM Rev. 56(3), 385–458 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  24. Borri, M., Trainelli, L., Bottasso, C.L.: On representations and parameterizations of motion. Multibody Syst. Dyn. 4, 129–193 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  25. Bauchau, O.A., Choi, J.Y.: The vector parameterization of motion. Nonlinear Dyn. 33(2), 165–188 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  26. Bauchau, O.A.: Flexible Multibody Dynamics. Springer, Dordrecht, Heidelberg, London, New-York (2011)

    Book  MATH  Google Scholar 

  27. Han, S.L., Bauchau, O.A.: Manipulation of motion via dual entities. Nonlinear Dyn. 85(1), 509–524 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  28. Clifford, W.K.: Preliminary sketch of biquaternions. Proc. Lond. Math. Soc. s1–4(1), 381–395 (1871)

    Article  MathSciNet  MATH  Google Scholar 

  29. McCarthy, J.M.: An Introduction to Theoretical Kinematics. The MIT Press, Cambridge, MA (1990)

    Google Scholar 

  30. Martinez, J.M.R., Duffy, J.: The principle of transference: history, statement and proof. Mech. Mach. Theory 28(1), 165–177 (1993)

    Article  Google Scholar 

  31. Angeles, J.: The application of dual algebra to kinematic analysis. In: Angeles, J., Zakhariev, E. (eds.) Computational Methods in Mechanical Systems, vol. 161, pp. 3–31. Springer, Heidelberg (1998)

    Chapter  Google Scholar 

  32. Fischer, I.S.: Dual Number Methods in Kinematics, Statics and Dynamics. CRC Press, Boca Raton (1999)

    Google Scholar 

  33. Pennestrì, E., Stefanelli, R.: Linear algebra and numerical algorithms using dual numbers. Multibody Syst. Dyn. 18, 323–344 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  34. Condurache, D., Burlacu, A.: Dual tensors based solutions for rigid body motion parameterization. Mech. Mach. Theory 74, 390–412 (2014)

    Article  Google Scholar 

  35. Romero, I.: The interpolation of rotations and its application to finite element models of geometrically exact rods. Comput. Mech. 34(2), 121–133 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  36. Bauchau, O.A., Han, S.L.: Interpolation of rotation and motion. Multibody Syst. Dyn. 31(3), 339–370 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  37. Sonneville, V., Brüls, O., Bauchau, O.A.: Interpolation schemes for geometrically exact beams: a motion approach. International Journal of Numerical Methods in Engineering 112, 1129–1153 (2017)

    Article  MathSciNet  Google Scholar 

  38. Grioli, G.: Una proprietà di minimo nella cinematica delle deformazioni finite. Bollettino dell’Unione Matematica Italiana 2, 252–255 (1940)

    MATH  Google Scholar 

  39. Malvern, L.E.: Introduction to the Mechanics of a Continuous Medium. Prentice Hall Inc, Englewood Cliffs, New Jersey (1969)

    MATH  Google Scholar 

  40. Martins, L.C., Podio-Guidugli, P.: A variational approach to the polar decomposition theorem. Rendiconti Accademia Nazionale dei Lincei, Serie VIII 66, 487–493 (1979)

    MathSciNet  MATH  Google Scholar 

  41. Martins, L.C., Podio-Guidugli, P.: An elementary proof of the polar decomposition theorem. Am. Math. Mon. 87(4), 288–290 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  42. Lu, J., Papadopoulos, P.: On the direct determination of the rotation tensor from the deformation gradient. Math. Mech. Solids 2(1), 17–26 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  43. Bouby, C., Fortuné, D., Pietraszkiewicz, W., Vallée, C.: Direct determination of the rotation in the polar decomposition of the deformation gradient by maximizing a Rayleigh quotient. J. Appl. Math. Mech. 85(3), 155–162 (2005)

    MathSciNet  MATH  Google Scholar 

  44. Wahba, G.: A least squares estimate of satellite attitude. SIAM Rev. 7(3), 409–409 (1965)

    Article  Google Scholar 

  45. Davenport, P.: A vector approach to the algebra of rotations with applications. Technical Report X-546-65-437, NASA, November (1965)

  46. Markley, F.L., Mortari, D.: Quaternion attitude estimation using vector observations. J. Astronaut. Sci. 48(2/3), 359–380 (2000)

    Google Scholar 

  47. Cardona, A., Géradin, M.: A beam finite element non-linear theory with finite rotation. Int. J. Numer. Methods Eng. 26, 2403–2438 (1988)

    Article  MATH  Google Scholar 

  48. Ibrahimbegović, A., Frey, F., Kozar, I.: Computational aspects of vector-like parameterization of three-dimensional finite rotations. Int. J. Numer. Methods Eng. 38(21), 3653–3673 (1995)

    Article  MATH  Google Scholar 

  49. Crisfield, M.A., Jelenić, G.: Objectivity of strain measures in the geometrically exact three-dimensional beam theory and its finite-element implementation. Proc. R. Soc. Lond. Math. Phys. Eng. Sci. 455(1983), 1125–1147 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  50. Betsch, P., Steinmann, P.: Frame-indifferent beam element based upon the geometrically exact beam theory. Int. J. Numer. Methods Eng. 54, 1775–1788 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  51. Gawlik, E.S., Leok, M.: Embedding-based interpolation on the special orthogonal group. (2016). arXiv:1608.05738 [math.NA]

  52. Horn, B.: Closed-form solution of absolute orientation using unit quaternions. J. Opt. Soc. Am. A 4(4), 629–642 (1987)

    Article  Google Scholar 

  53. Moakher, M.: Means and averaging in the group of rotations. SIAM J. Matrix Anal. Appl. 24(1), 1–16 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  54. Hartley, R., Trumpf, J., Dai, Y.C., Li, H.D.: Rotation averaging. Int. J. Comput. Vis. 103(3), 267–305 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  55. Bauchau, O.A., Han, S.L.: Three-dimensional beam theory for flexible multibody dynamics. J. Comput. Nonlinear Dyn. 9(4), 041011 (2014)

    Article  Google Scholar 

  56. Han, S.L., Bauchau, O.A.: Nonlinear three-dimensional beam theory for flexible multibody dynamics. Multibody Syst. Dyn. 34(3), 211–242 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  57. Han, S.L., Bauchau, O.A.: On Saint-Venant’s problem for helicoidal beams. J. Appl. Mech. 83(2), 021009 (2016)

    Article  Google Scholar 

  58. Bauchau, O.A., Betsch, P., Cardona, A., Gerstmayr, J., Jonker, B., Masarati, P., Sonneville, V.: Validation of flexible multibody dynamics beam formulations using benchmark problems. Multibody Syst. Dyn. 37(1), 29–48 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  59. Dimentberg, F.M.: The screw calculus and its applications. Technical Report AD 680993, Clearinghouse for Federal and Scientific Technical Information, Virginia, USA, April (1968)

  60. Golub, G.H., van Loan, C.F.: Matrix Computations, 2nd edn. The Johns Hopkins University Press, Baltimore (1989)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shilei Han.

Appendices

Appendix A: Dual numbers, vectors, matrices, and functions

Dual numbers were first introduced in the nineteenth century by Clifford [28]. Typically, they are written as , where a and \(a^{o}\) are referred to as the primal and dual parts, respectively, and parameter \(\epsilon \) is such that \(\epsilon ^{n} = 0\) for \(n \ge 2\). The product of two dual scalars now becomes . The zero and identity dual numbers are and , respectively. The inverse of a dual scalar is .

A dual vector is composed of two vectors of the same size . The inner product of two dual vectors is . The cross product of two dual vectors of size \(3\times 1\) is , where notation \(\tilde{(\cdot )}\) is the skew-symmetric dual matrix associated with . Dual vector is unit if , which is equivalent to \(\underline{a}^{ T} \underline{a}= 1\) and \(\underline{a}^\mathrm{{T}} \underline{a}^{o} = 0\).

Similarly, a dual matrix is composed of two matrices of the same size . The transpose of a dual matrix is . The identity dual matrix is defined as . The product of a dual matrix by a dual vector is found easily as . Dual matrix is orthogonal if , which is equivalent to \(\underline{\underline{A}}^\mathrm{{T}} \underline{\underline{A}}= \underline{\underline{I}}\) and .

Consider a dual function of a dual variable, . Function is assumed to be analytic [33, 59], i.e., it is of the following form: , where J(a) is the primal part of dual function and notation \((\cdot )'\) indicates a derivative with respect to a. The derivative of dual function with respect to dual variable is a dual number, . Two important observations can be made: (1) the primal part of an analytic function depends on the primal part of its dual variable only and (2) the dual part of an analytic function is a linear function of the dual part of its dual variable. Because the magnitude of a dual number is not defined, dual numbers cannot be compared. The primal part of a dual function, J(a), can reach a minimum or maximum, whereas its dual part, \(a^{o} J' (a)\), never reaches a minimum or maximum in a open set because it is a linear function of \(a^{o}\). Minimization of a dual function is not meaningful.

Appendix B: Identities of matrix and vectors

This section presents a set of useful identities that are used throughout the paper. The identities involve a dual matrix, , of size \(3\times 3\), and a dual vector, , of size \(3\times 1\).

(38a)
(38b)
(38c)
(38d)
(38e)
(38f)

These identities can be verified easily.

Appendix C: The polar decomposition theorem

The polar decomposition theorem can be stated as follows.

Theorem 1

(Polar decomposition theorem) An invertible matrix, \(\underline{\underline{G}}\in \mathbb {R}^{3\times 3}\), can be decomposed into the product of a rotation tensor, \(\underline{\underline{R}}\in \mathrm {SO}(3)\), by a symmetric matrix, \(\underline{\underline{S}}\), as \(\underline{\underline{G}}= \underline{\underline{R}}\, \underline{\underline{S}}\). Matrices \(\underline{\underline{R}}\) and \(\underline{\underline{S}}\) are defined uniquely if matrix \([ \mathrm {tr}(\underline{\underline{S}}) \underline{\underline{I}}- \underline{\underline{S}}]\) is required to be positive definite.

Proof

Let the spectral decomposition of positive-definite matrix \(\underline{\underline{G}}^\mathrm{{T}} \underline{\underline{G}}\) be \(\underline{\underline{U}}^\mathrm{{T}} \mathrm {diag}(\lambda _{1}, \lambda _{2}, \lambda _{3})\underline{\underline{U}}\), where positive eigenvalues, \(\lambda _{i}\), \(i = 1, 2, 3\), satisfy \(\lambda _{1} \le \lambda _{2} \le \lambda _{3}\). In view of identity \(\underline{\underline{S}}^\mathrm{{T}} \underline{\underline{S}}= (\underline{\underline{R}}^\mathrm{{T}} \underline{\underline{G}})^\mathrm{{T}} (\underline{\underline{R}}^\mathrm{{T}} \underline{\underline{G}}) = \underline{\underline{G}}^\mathrm{{T}} \underline{\underline{G}}\), symmetric matrix \(\underline{\underline{S}}\) can be chosen as \(\underline{\underline{U}}^\mathrm{{T}}\mathrm {diag}(\pm \sqrt{\lambda _{1}},\pm \sqrt{\lambda _{2}},\pm \sqrt{\lambda _{3}})\underline{\underline{U}}\). Of these eight choices, two only, \(\underline{\underline{U}}^\mathrm{{T}} \mathrm {diag}(\pm \sqrt{\lambda _{1}}, \sqrt{\lambda _{2}}, \sqrt{\lambda _{3}}) \underline{\underline{U}}\), render matrix \([ \mathrm {tr}(\underline{\underline{S}})\underline{\underline{I}}- \underline{\underline{S}}]\) positive definite. The sign of the lowest eigenvalue is determined by sign of \(\det (\underline{\underline{G}}) = \det (\underline{\underline{R}}) \det (\underline{\underline{S}}) = \det (\underline{\underline{S}})\): Choose the positive or negative sign if \(\mathrm {det}(\underline{\underline{G}})\) is positive or negative, respectively. \(\square \)

Remark 1

Polar decomposition theorem (1) differs slightly from the traditional polar decomposition theorem used in continuum mechanics [39, 60]. The proof above shows that eight different symmetric matrices \(\underline{\underline{S}}\) satisfy multiplicative decomposition \(\underline{\underline{G}}= \underline{\underline{R}}\, \underline{\underline{S}}\). The solution is made unique by imposing an additional condition: In the traditional and present versions of the theorem, matrices \(\underline{\underline{S}}\) and \([ \mathrm {tr}(\underline{\underline{S}}) \underline{\underline{I}}- \underline{\underline{S}}]\) are required to be positive definite, respectively. When \(\mathrm {det}(\underline{\underline{G}}) > 0\), the two theorems are identical.

Given an invertible matrix \(\underline{\underline{G}}\in \mathbb {R}^{3\times 3}\), the following question is asked: Which rotation tensor \(\underline{\underline{R}}\in \mathrm {SO}(3)\) is as close as possible to \(\underline{\underline{G}}\)? The problem is stated as

$$\begin{aligned} \min _{\underline{\underline{R}}\in \mathrm {SO}(3)} J(\underline{\underline{R}}) = \Vert \underline{\underline{R}}- \underline{\underline{G}}\Vert _{F}^{2}, \end{aligned}$$
(39)

where the closeness of the two matrices is defined as the square of the Frobenius norm of their difference.

Theorem 2

(Closest rotation tensor) The rotation tensor that satisfies minimization problem (39) is that provided by polar decomposition Theorem (1).

Proof

Defining \(\underline{\underline{S}}= \underline{\underline{R}}^\mathrm{{T}} \underline{\underline{G}}\), the objective function becomes \(J = \mathrm {tr}[ (\underline{\underline{R}}- \underline{\underline{G}})^\mathrm{{T}} (\underline{\underline{R}}- \underline{\underline{G}}) ] = \mathrm {tr}[ \underline{\underline{I}}+ \underline{\underline{G}}^\mathrm{{T}} \underline{\underline{G}}- \underline{\underline{S}}- \underline{\underline{S}}^\mathrm{{T}} ] = \mathrm {tr}(\underline{\underline{I}}+ \underline{\underline{G}}^\mathrm{{T}} \underline{\underline{G}}) - 2 \mathrm {tr}(\underline{\underline{S}})\) and its variation is \(\delta J = - 2 \mathrm {tr}(\delta \underline{\underline{S}}) = 2 \mathrm {tr}(\tilde{\delta \psi }\, \underline{\underline{S}})\), where \(\tilde{\delta \psi }= \underline{\underline{R}}^\mathrm{{T}} \delta \underline{\underline{R}}\). Finally, trace identity (38e) yields \(\delta J = - 4 \underline{\delta \psi }^\mathrm{{T}} \mathrm {axial}(\underline{\underline{S}}) = 0\), which implies that matrix \(\underline{\underline{S}}\) must be symmetric. For the objective function to reach its minimum, its Hessian must be positive definite. Taking a second-order variation yields \(\delta ^{2} J = - 2 \underline{\delta \psi }^\mathrm{{T}} \mathrm {axial}(\delta \underline{\underline{S}})\), leading to \(\delta ^{2} J = 2 \underline{\delta \psi }^\mathrm{{T}} [\mathrm {tr}{(\underline{\underline{S}})} \underline{\underline{I}}- \underline{\underline{S}}] \underline{\delta \psi }\), where matrix \([\mathrm {tr}{(\underline{\underline{S}})} \underline{\underline{I}}- \underline{\underline{S}}]\) is the Hessian. Clearly, polar decomposition theorem (1) provides uniquely defined rotation tensor \(\underline{\underline{R}}\) and matrix \(\underline{\underline{S}}\) that satisfy the two required conditions: \(\underline{\underline{S}}\) is symmetric and \([\mathrm {tr}{(\underline{\underline{S}})} \underline{\underline{I}}- \underline{\underline{S}}]\) is positive definite.

The polar decomposition theorem is now generalized to dual matrices.

Theorem 3

(Dual polar decomposition theorem) An invertible dual matrix, , can be decomposed into the product of a dual orthogonal matrix, , by a symmetric dual matrix, , as . Matrices and are defined uniquely if it is also required that matrix \([ \mathrm {tr}(\underline{\underline{S}}) \underline{\underline{I}}- \underline{\underline{S}}]\) be positive definite, where \(\underline{\underline{S}}\) is the primal part of dual matrix .

Proof

To prove the theorem, dual matrices and will be constructed and the solution will be shown to be unique. First, dual identity is expanded as \((\underline{\underline{G}}+ \epsilon \underline{\underline{G}}^{o}) = (\underline{\underline{R}}+ \epsilon \underline{\underline{R}}^{o}) (\underline{\underline{S}}+ \epsilon \underline{\underline{S}}^{o})\), which implies

$$\begin{aligned}&\underline{\underline{G}}= \underline{\underline{R}}\, \underline{\underline{S}}, \end{aligned}$$
(40a)
$$\begin{aligned}&\underline{\underline{G}}^{o} = \underline{\underline{R}}\, \underline{\underline{S}}^{o} + \underline{\underline{R}}^{o} \underline{\underline{S}}. \end{aligned}$$
(40b)

Equation (40a) expresses Theorem (1), i.e., rotation tensor \(\underline{\underline{R}}\) and symmetric matrix \(\underline{\underline{S}}\) are defined uniquely. Equation (40b) implies \(\underline{\underline{R}}^{ T} \underline{\underline{G}}^{o} = \underline{\underline{S}}^{o} + (\underline{\underline{R}}^{ T} \underline{\underline{R}}^{o}) \underline{\underline{S}}\), where matrix \(\underline{\underline{R}}^{ T} \underline{\underline{R}}^{o} = \tilde{z}\) is antisymmetric because motion tensor is orthogonal. Because matrix \(\underline{\underline{S}}^{o}\) must be symmetric, \(\mathrm {axial}(\underline{\underline{S}}^{o}) = \underline{0}\), and extracting the axial part of this equation yields \(\mathrm {axial}(\underline{\underline{R}}^{ T} \underline{\underline{G}}^{o}) = \mathrm {axial}[\tilde{z}\underline{\underline{S}}]\). Identity (38a) now yields \([\mathrm {tr}{(\underline{\underline{S}})} \underline{\underline{I}}- \underline{\underline{S}}] \underline{z}= 2 \mathrm {axial}(\underline{\underline{R}}^{ T} \underline{\underline{G}}^{o} )\), a linear system that can be solved to find \(\underline{z}\). The dual parts of motion tensor and symmetric matrix are found as \(\underline{\underline{R}}^{o} = \underline{\underline{R}}\tilde{z}\) and \(\underline{\underline{S}}^{o} = \underline{\underline{R}}^{ T} \underline{\underline{G}}^{o} - \tilde{z}\underline{\underline{S}}\), respectively. \(\square \)

Given an invertible dual matrix , the following question is asked: Which motion tensor is as close as possible to ? By analogy with Eq. (39), the following minimization problem is introduced

(41)

As discussed earlier, the minimization of a function of dual numbers is devoid of meaning. Therefore, notation “” is now defined as follows.

Definition 1

(Minimization of a dual function) The minimization of a dual function of dual variables, , implies the satisfaction of two conditions: (1) the variation of the objective function must vanish, , and (2) the primal part of the objective function must achieve a minimum.

As discussed in “Appendix A,” the primal part of a dual function can reach a minimum, whereas its dual part never reaches a minimum in a open set because it is a linear function of the dual part of its variable. Definition (1) does not require the minimization of the dual part of the function, bypassing the problem. Note that the vanishing of the variation of the primal part of the objective function is a prerequisite for its minimization. An alternative statement of definition (1) reads: The minimization of a function of dual variables implies the minimization of its primal part and the stationarity of its dual part.

Theorem 4

(Closest motion tensor) The motion tensor that satisfies dual minimization problem (41) is that provided by dual polar decomposition Theorem (3).

Proof

Defining , the objective function becomes and its variation is , where . Finally, trace identity (38e) yields , which implies that dual matrix must be symmetric. The objective function is evaluated as . The minimization of its primal part is expressed by minimization problem (39), which is solved by the polar decomposition theorem; see Theorem (2). Clearly, dual polar decomposition theorem (3) provides uniquely defined motion tensor and matrix that satisfy the two required conditions: is symmetric and \([\mathrm {tr}{(\underline{\underline{S}})} \underline{\underline{I}}- \underline{\underline{S}}]\) is positive definite.

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Han, S., Bauchau, O.A. Spectral collocation methods for the periodic solution of flexible multibody dynamics. Nonlinear Dyn 92, 1599–1618 (2018). https://doi.org/10.1007/s11071-018-4149-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-018-4149-6

Keywords

Navigation