Skip to main content
Log in

Interpolation of rotation and motion

  • Published:
Multibody System Dynamics Aims and scope Submit manuscript

Abstract

In Cosserat solids such as shear deformable beams and shells, the displacement and rotation fields are independent. The finite element implementation of these structural components within the framework of flexible multibody dynamics requires the interpolation of rotation and motion fields. In general, the interpolation process does not preserve fundamental properties of the interpolated field. For instance, interpolation of an orthogonal rotation tensor does not yield an orthogonal tensor, and furthermore, does not preserve the tensorial nature of the rotation field. Consequently, many researchers have been reluctant to apply the classical interpolation tools used in finite element procedures to interpolate these fields. This paper presents a systematic study of interpolation algorithms for rotation and motion. All the algorithms presented here preserve the fundamental properties of the interpolated rotation and motion fields, and furthermore, preserve their tensorial nature. It is also shown that the interpolation of rotation and motion is as accurate as the interpolation of displacement, a widely accepted tool in the finite element method. The algorithms presented in this paper provide interpolation tools for rotation and motion that are accurate, easy to implement, and physically meaningful.

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

Similar content being viewed by others

References

  1. Agrawal, O., Shabana, A.: Application of deformable-body mean axis to flexible multibody system dynamics. Comput. Methods Appl. Mech. Eng. 56(2), 217–245 (1986)

    Article  MATH  Google Scholar 

  2. Bathe, K.: Finite Element Procedures. Prentice Hall, Englewood Cliffs (1996)

    Google Scholar 

  3. Bauchau, O.: Flexible Multibody Dynamics. Springer, Dordrecht (2011)

    Book  MATH  Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

  5. Bauchau, O., Craig, J.: Structural Analysis with Application to Aerospace Structures. Springer, Dordrecht (2009)

    Google Scholar 

  6. Bauchau, O., Li, L.: Tensorial parameterization of rotation and motion. J. Comput. Nonlinear Dyn. 6(3), 031007 (2011)

    Article  Google Scholar 

  7. Bauchau, O., Trainelli, L.: The vectorial parameterization of rotation. Nonlinear Dyn. 32(1), 71–92 (2003)

    Article  MATH  MathSciNet  Google Scholar 

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

    Article  MATH  MathSciNet  Google Scholar 

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

    Article  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  11. Géradin, M., Cardona, A.: Kinematics and dynamics of rigid and flexible mechanisms using finite elements and quaternion algebra. Comput. Mech. 4, 115–135 (1989)

    Article  Google Scholar 

  12. Géradin, M., Cardona, A.: Flexible Multibody System: A Finite Element Approach. Wiley, New York (2001)

    Google Scholar 

  13. Hughes, T.: The Finite Element Method. Prentice Hall, Englewood Cliffs (1987)

    MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  15. Jelenić, G., Crisfield, M.: Geometrically exact 3D beam theory: implementation of a strain-invariant finite element for static and dynamics. Comput. Methods Appl. Mech. Eng. 171, 141–171 (1999)

    Article  MATH  Google Scholar 

  16. Malvern, L.: Introduction to the Mechanics of a Continuous Medium. Prentice Hall, Englewood Cliffs (1969)

    Google Scholar 

  17. Milenković, V.: Coordinates suitable for angular motion synthesis in robots. In: Proceedings of the Robot VI Conference, Detroit, MI, 2–4 March 1982 (1982). Paper MS82-217

    Google Scholar 

  18. 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  MATH  MathSciNet  Google Scholar 

  19. Romero, I., Armero, F.: An objective finite element approximation of the kinematics of geometrically exact rods and its use in the formulation of an energy–momentum conserving scheme in dynamics. Int. J. Numer. Methods Biomed. Eng. 54, 1683–1716 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  20. Shabana, A.: Flexible multibody dynamics: review of past and recent developments. Multibody Syst. Dyn. 1(2), 189–222 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  21. Shabana, A.: Uniqueness of the geometric representation in large rotation finite element formulations. J. Comput. Nonlinear Dyn. 5(4), 044501 (2010)

    Article  Google Scholar 

  22. Shabana, A.: General method for modeling slope discontinuity and T-sections using ANCF gradient deficient finite elements. J. Comput. Nonlinear Dyn. 6(2), 024502 (2011)

    Article  Google Scholar 

  23. Shabana, A., Mikkola, A.: Use of the finite element absolute nodal coordinate formulation in modeling slope discontinuity. J. Mech. Des. 125(2), 342–350 (2003)

    Article  Google Scholar 

  24. Shabana, A., Wehage, R.: A coordinate reduction technique for dynamic analysis of spatial substructures with large angular rotations. J. Struct. Mech. 11(3), 401–431 (1983)

    Article  Google Scholar 

  25. Simo, J., Vu-Quoc, L.: A three dimensional finite strain rod model. Part II: Computational aspects. Comput. Methods Appl. Mech. Eng. 58(1), 79–116 (1986)

    Article  MATH  Google Scholar 

  26. Stuelpnagel, J.: On the parameterization of the three-dimensional rotation group. SIAM Rev. 6(4), 422–430 (1964)

    Article  MATH  MathSciNet  Google Scholar 

  27. Wehage, R.: Quaternions and Euler parameters. A brief exposition. In: Haug, E. (ed.) Computer Aided Analysis and Optimization of Mechanical Systems Dynamics, pp. 147–180. Springer, Berlin (1984)

    Chapter  Google Scholar 

  28. Wiener, T.: Theoretical analysis of gimballess inertial reference equipment using delta-modulated instruments. Ph.D. thesis, Department of Aeronautical and Astronautical Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts (1962)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Olivier A. Bauchau.

Appendices

Appendix A: Notational conventions

When dealing with unit quaternions, \(\hat {e}\), the rotation tensor, \(\underline {\underline {R}}\), is expressed by the following matrix of size 4×4:

(A.1)

The following matrices are used to manipulate quaternions:

(A.2)

When dealing with bi-quaternions, \(\check {g}^{T} = \{ \hat {q}^{T}, \hat {e}^{T} \}\), the motion tensor, \(\underline {\underline {\mathcal {C}}}\), is expressed by the following matrix of size 8×8:

(A.3)

The following matrix is used to manipulate bi-quaternions:

(A.4)

When dealing with motion, generalized skew-symmetric matrices are of the following form:

(A.5)

where \(\widetilde {a}\) and \(\widetilde {b}\) are 3×3 skew-symmetric matrices. Matrix \(\underline {\underline {\mathcal {Z}}}\) is defined as follows:

(A.6)

where α and β are two arbitrary scalars.

Appendix B: Polar decomposition theorem

The polar decomposition theorem states that an invertible matrix, \(\underline {\underline {A}}\), can be decomposed into the product of an orthogonal matrix, \(\underline {\underline {R}}\), by a positive-definite, symmetric matrix, \(\underline {\underline {U}}\),

$$ \underline {\underline {A}}= \underline {\underline {R}}\, \underline {\underline {U}}= \underline {\underline {V}}\, \underline {\underline {R}}. $$
(B.1)

The second equality provides an alternative statement of the polar decomposition theorem, where matrix \(\underline {\underline {A}}\) is now decomposed as the product a positive-definite, symmetric matrix, \(\underline {\underline {V}}\), by the same orthogonal matrix, \(\underline {\underline {R}}\). Matrices \(\underline {\underline {R}}\), \(\underline {\underline {U}}\), and \(\underline {\underline {V}}\) are uniquely defined.

The polar decomposition theorem will be generalized by the following statement:

$$ \underline {\underline {\mathcal {T}}}= \underline {\underline {\mathcal {C}}}\, \underline {\underline {\mathcal {U}}}= \underline {\underline {\mathcal {V}}}\, \underline {\underline {\mathcal {C}}}, $$
(B.2)

where \(\underline {\underline {\mathcal {C}}}\) is the motion tensor defined by Eq. (4.1). Matrix \(\underline {\underline {\mathcal {T}}}\) is an arbitrary matrix of the following form:

(B.3)

where \(\underline {\underline {T}}\) is an invertible matrix and matrices \(\underline {\underline {\mathcal {U}}}\) and \(\underline {\underline {\mathcal {V}}}\) are

(B.4a)
(B.4b)

Matrices \(\underline {\underline {U}}\) and \(\underline {\underline {V}}\) are positive-definite and symmetric matrices; matrices \(\underline {\underline {F}}\) and \(\underline {\underline {G}}\) are symmetric.

The generalization of the polar decomposition theorem expressed by Eq. (B.2) is proved easily. The first equality of Eq. (B.2) implies \(\underline {\underline {T}}= \underline {\underline {R}}\, \underline {\underline {U}}\); this means that matrices \(\underline {\underline {R}}\) and \(\underline {\underline {U}}\) correspond to the polar decomposition of \(\underline {\underline {T}}\), see Eq. (B.1) and, therefore, matrix \(\underline {\underline {R}}\) is orthogonal, matrix \(\underline {\underline {U}}\) positive-definite and symmetric, and both are defined uniquely. Next, the first equality of Eq. (B.2) also implies \(\underline {\underline {W}}= \underline {\underline {R}}\, \underline {\underline {F}}+ \widetilde {u}\underline {\underline {R}}\, \underline {\underline {U}}\), which can be recast as \(\widetilde {u}^{*} \underline {\underline {U}}+ \underline {\underline {F}}= \underline {\underline {R}}^{T} \underline {\underline {W}}\), where \(\underline {u}^{*} = \underline {\underline {R}}^{T} \underline {u}\). Because matrix \(\underline {\underline {F}}\) is symmetric, \((\widetilde {u}^{*} \underline {\underline {U}}) - (\widetilde {u}^{*} \underline {\underline {U}})^{T} = (\underline {\underline {R}}^{T} \underline {\underline {W}}) - (\underline {\underline {R}}^{T} \underline {\underline {W}})^{T}\), which leads to a linear system of equations for \(\underline {u}^{*}\),

$$ \bigl[ \mathrm {tr}(\underline {\underline {U}}) \underline {\underline {I}}- \underline {\underline {U}}\bigr] \underline {u}^{*}= 2 \mathrm {axial}\bigl(\underline {\underline {R}}^{T} \underline {\underline {W}}\bigr). $$
(B.5)

Because matrices \(\underline {\underline {U}}\), \(\underline {\underline {R}}\), and \(\underline {\underline {W}}\) are known, a unique solution exists \(\underline {u}^{*}\), leading to \(\underline {u}= \underline {\underline {R}}\underline {u}^{*}\), provided that the determinant of the system does not vanish. It is shown easily that \(\mathrm {det}[ \mathrm {tr}(\underline {\underline {U}}) \underline {\underline {I}}- \underline {\underline {U}}] = (\lambda_{2}^{2} + \lambda_{3}^{2}) (\lambda_{1}^{2} + \lambda_{3}^{2}) (\lambda_{1}^{2} + \lambda_{2}^{2}) \), where \(\lambda_{i}^{2}\), i=1,2,3 are the positive eigenvalues of positive-definite matrix \(\underline {\underline {U}}\), and hence, a solution of system (B.5) always exists. Finally, matrix \(\underline {\underline {F}}= \underline {\underline {R}}^{T} (\underline {\underline {W}}- \widetilde {u}\underline {\underline {T}})\), which is symmetric because \(\underline {u}\) was determined by the solution of system (B.5). This proves that all the quantities appearing in decomposition (B.2) can be found unequivocally, hence proving the stated generalization of the polar decomposition theorem. The second equality in Eq. (B.2 can be proved in a similar manner.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Bauchau, O.A., Han, S. Interpolation of rotation and motion. Multibody Syst Dyn 31, 339–370 (2014). https://doi.org/10.1007/s11044-013-9365-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11044-013-9365-8

Keywords

Navigation