Skip to main content
Log in

Nonlinear three-dimensional beam theory for flexible multibody dynamics

  • Published:
Multibody System Dynamics Aims and scope Submit manuscript

Abstract

In flexible multibody systems, it is convenient to approximate many structural components as beams or shells. Classical beam theories, such as Euler–Bernoulli beam theory, often form the basis of the analytical development for beam dynamics. The advantage of this approach is that it leads to a very simple kinematic representation of the problem: the beam’s section is assumed to remain plane and its displacement field is fully defined by three displacement and three rotation components. While such an approach is capable of capturing the kinetic energy of the system accurately, it cannot represent the strain energy adequately. This paper presents a different approach to the problem. Based on a finite element discretization of the cross-section, an exact solution of the theory of three-dimensional elasticity is developed. The proposed approach is based on the Hamiltonian formalism and leads to an expansion of the solution in terms of extremity and central solutions. Kinematically, the problem is decomposed into an arbitrarily large rigid-section motion and a warping field. The sectional strains associated with the rigid-section motion and the warping field are assumed to remain small. As a consequence of this kinematic decomposition, the governing equations of the problem fall into two distinct categories: the equations describing geometrically exact beams and those describing local deformations. The governing equations for geometrically exact beams are nonlinear, one-dimensional equations, whereas a linear, two-dimensional analysis provides the detailed distribution of three-dimensional stress and strain fields. Within the stated assumptions, the solutions presented here are the exact solution of three-dimensional elasticity for beams undergoing arbitrarily large motions.

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

Similar content being viewed by others

References

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

    Google Scholar 

  2. Giavotto, V., Borri, M., Mantegazza, P., Ghiringhelli, G., Carmaschi, V., Maffioli, G.C., Mussi, F.: Anisotropic beam theory and applications. Comput. Struct. 16(1–4), 403–413 (1983)

    Article  MATH  Google Scholar 

  3. Mielke, A.: Saint-Venant’s problem and semi-inverse solutions in nonlinear elasticity. Arch. Ration. Mech. Anal. 102, 205–229 (1988)

    Article  MATH  MathSciNet  Google Scholar 

  4. Mielke, A.: Normal hyperbolicity of center manifolds and Saint-Venant’s principle. Arch. Ration. Mech. Anal. 110, 353–372 (1990)

    Article  MATH  MathSciNet  Google Scholar 

  5. Mielke, A.: Hamiltonian and Lagrangian Flows on Center Manifolds with Applications to Elliptic Variational Problems. Lecture Notes in Mathematics, vol. 1489. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  6. Zhong, W.X.: A New Systematic Methodology for Theory of Elasticity. Dalian University of Technology Press, Dalian (1995)

    Google Scholar 

  7. Zhong, W.X., Xu, X.S., Zhang, H.W.: Hamiltonian system and the Saint-Venant problem in elasticity. Appl. Math. Mech. 17(9), 827–836 (1996)

    Article  MATH  MathSciNet  Google Scholar 

  8. Yao, W.A., Zhong, W.X., Lim, C.W.: Symplectic Elasticity. World Scientific, New Jersey (2009)

    Book  MATH  Google Scholar 

  9. Morandini, M., Chierichetti, M., Mantegazza, P.: Characteristic behavior of prismatic anisotropic beam via generalized eigenvectors. Int. J. Solids Struct. 47, 1327–1337 (2010)

    Article  MATH  Google Scholar 

  10. Berdichevsky, V.L.: On the energy of an elastic rod. Prikl. Mat. Meh. 45(4), 518–529 (1982)

    Google Scholar 

  11. Hodges, D.H.: A review of composite rotor blade modeling. AIAA J. 28(3), 561–565 (1990)

    Article  Google Scholar 

  12. Bauchau, O.A., Han, S.L.: Three-dimensional beam theory for flexible multibody dynamics. J. Comput. Nonlinear Dyn. 9(4), 041011 (2014). 12 pp.

    Article  Google Scholar 

  13. Simo, J.C., Vu-Quoc, L.: On the dynamics in space of rods undergoing large motions—a geometrically exact approach. Comput. Methods Appl. Mech. Eng. 66(1), 125–161 (1988)

    Article  MATH  MathSciNet  Google Scholar 

  14. Borri, M., Merlini, T.: A large displacement formulation for anisotropic beam analysis. Meccanica 21, 30–37 (1986)

    Article  MATH  Google Scholar 

  15. Danielson, D.A., Hodges, D.H.: A beam theory for large global rotation, moderate local rotation, and small strain. J. Appl. Mech. 55(1), 179–184 (1988)

    Article  Google Scholar 

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

    Book  MATH  Google Scholar 

  17. Volovoi, V.V., Hodges, D.H., Berdichevsky, V.L., Sutyrin, V.G.: Dynamic dispersion curves for non-homogeneous, anisotropic beams with cross sections of arbitrary geometry. J. Sound Vib. 215, 1101–1120 (1998)

    Article  Google Scholar 

  18. Lanczos, C.: The Variational Principles of Mechanics. Dover, New York (1970)

    MATH  Google Scholar 

  19. de Saint-Venant, J.C.-B.: Mémoire sur la torsion des prismes. Rec. Savants Étrang. 14, 233–560 (1855)

    Google Scholar 

  20. Simo, J.C.: The three-dimensional dynamic problem. Part I. A finite strain beam formulation. Comput. Methods Appl. Mech. Eng. 49(1), 55–70 (1985)

    Article  MATH  MathSciNet  Google Scholar 

  21. Simo, J.C., 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 

  22. Danielson, D.A., Hodges, D.H.: Nonlinear beam kinematics by decomposition of the rotation tensor. J. Appl. Mech. 54(2), 258–262 (1987)

    Article  MATH  Google Scholar 

  23. Atilgan, A.R., Hodges, D.H.: Unified nonlinear analysis for nonhomogeneous anisotropic beams with closed cross sections. AIAA J. 29(11), 1990–1999 (1991)

    Article  MATH  Google Scholar 

  24. Atilgan, A.R., Hodges, D.H., Fulton, M.V.: Nonlinear deformation of composite beams: unification of cross-sectional and elastica analyses. Appl. Mech. Rev. 44(11), S9–S15 (1991)

    Article  Google Scholar 

  25. Borri, M., Ghiringhelli, G.L., Merlini, T.: Linear analysis of naturally curved and twisted anisotropic beams. Compos. Eng. 2(5-7), 433–456 (1992)

    Article  Google Scholar 

  26. Timoshenko, S.P., Goodier, J.N.: Theory of Elasticity, 3rd edn. McGraw-Hill, New York (1970)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Olivier A. Bauchau.

Appendices

Appendix A: Hamiltonian matrices

Matrix , of size 2n×2n, is said to be Hamiltonian if it satisfies the following property

(73)

where the skew-symmetric matrix is defined as

(74)

Note the following properties of matrix :

(75a)
(75b)

In view of these properties, the definition (73) of Hamiltonian matrices can be recast as . Because matrix is skew-symmetric, the following property holds for any vector \(\underline{\mathcal{U}}\):

(76)

Property (73) implies that the most general form of a Hamiltonian matrix is

(77)

where matrices , , and are of size n×n, and matrices and are symmetric. Clearly, the transpose of a Hamiltonian matrix is also Hamiltonian.

1.1 A.1 Eigenvalues of Hamiltonian matrices

Let λ and μ be two eigenvalues of a Hamiltonian matrix and the associated eigenvectors are denoted \(\underline{\mathcal{U}}_{\,\lambda}\) and \(\underline{\mathcal{U}}_{\,\mu }\), respectively, i.e., and . Pre-multiplying the first equation by and the second by leads to and , respectively. Because matrix is symmetric, the right-hand sides of these two equations are identical, and subtraction yields . If λ+μ≠0, the following symplectic orthogonality results:

(78)

Let λ be an eigenvalue of a Hamiltonian matrix associated with eigenvector \(\underline{\mathcal{U}}_{\,+\lambda}\), i.e., . It then follows that and properties (73) and (75a) then imply . Clearly, if \(\underline{\mathcal{U}}_{\,+\lambda}\) is an eigenvector of associated with eigenvalue λ, is an eigenvector of associated with eigenvalue −λ. Because the spectra of eigenvalues of matrices and are identical, the eigenvalues of Hamiltonian matrices are symmetric about the imaginary axis, i.e., occur in pairs of opposite sign, ±λ. In summary,

(79a)
(79b)

where Eqs. (79a) and (79b) express identical properties for eigenvalues +λ and −λ, respectively. Vectors \(\underline{\mathcal{U}}_{\,+\lambda }\) and can also be interpreted as the right and left eigenvectors of matrix , both associated with eigenvalue +λ.

It is convenient to define matrix , whose columns store the right eigenvectors associated with eigenvalues −λ and +λ. These vectors enjoy the following properties

(80)

Indeed, property (76) implies and vectors \(\underline{\mathcal{U}}_{\,+\lambda}\) and \(\underline{\mathcal{U}}_{\,-\lambda}\) can be normalized to enforce . The following pseudo-orthogonality statement in the space of Hamiltonian matrix then follows

(81)

Appendix B: Symplectic matrices

Matrix , of size 2n×2n, is said to be symplectic if it satisfies the following property

(82)

where the skew-symmetric matrix is defined by Eq. (74). The inverse of a symplectic matrix is expressed easily as

(83)

Indeed, using definition (82) leads to .

Recasting Eq. (82) as and using Eq. (83) to express the inverse of the symplectic matrix leads to , and finally,

(84)

This means that the transpose of a symplectic matrix is itself symplectic. It is proved easily that if matrix is symplectic, its inverse is also symplectic. Furthermore, the product of two symplectic matrices is also symplectic. Note that matrices and are symplectic.

A close connection exists between symplectic and Hamiltonian matrices. First, their definitions are closely related: and , for Hamiltonian and symplectic matrices, respectively. Next, consider matrix , defined through the following transformation, . Using Eq. (83) leads to , which implies that matrix is symmetric if matrix is itself symmetric. Consequently, if matrix is Hamiltonian, so is matrix . It follows that the transformation of a Hamiltonian matrix by a symplectic matrix yields a Hamiltonian matrix, i.e.,

(85)

Consider the following partition of a symplectic matrix:

(86)

where matrices , , , and are of size n×n. The definition of a symplectic matrix then implies

(87a)
(87b)
(87c)

The last two properties imply that products and form symmetric matrices.

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. Nonlinear three-dimensional beam theory for flexible multibody dynamics. Multibody Syst Dyn 34, 211–242 (2015). https://doi.org/10.1007/s11044-014-9433-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11044-014-9433-8

Keywords

Navigation