Skip to main content

Energy-Momentum Integrators for Elastic Cosserat Points, Rigid Bodies, and Multibody Systems

  • Chapter
  • First Online:
Structure-preserving Integrators in Nonlinear Structural Dynamics and Flexible Multibody Dynamics

Part of the book series: CISM International Centre for Mechanical Sciences ((CISM,volume 565))

Abstract

The goal of this chapter is to present the development of energy-momentum (EM) schemes in the framework of discrete (or finite-dimensional) mechanical systems. EM integrators belong to the class of structure-preserving numerical methods and have been originally developed in the field of nonlinear solid and structural mechanics. EM schemes and energy dissipating variants thereof typically exhibit improved numerical stability and robustness when compared to standard integrators. Due to their superior numerical properties, EM schemes have soon been extended to more involved applications such as flexible multibody dynamics and coupled thermomechanical problems. In this chapter, we start the development of second-order EM schemes in the context of the Cosserat point (or pseudo-rigid body). The theory of a Cosserat point shares main structural properties with semi-discrete formulations of elastodynamics. Indeed, the Cosserat point can be directly linked to the 4-node tetrahedral finite element. Besides its usefulness in explaining main ingredients of EM schemes such as the algorithmic stress formula, the Cosserat point is ideally suited to perform the transition to rigid body dynamics. In particular, in the present work, the rigid body formulation is obtained by imposing the zero strain condition on the Cosserat point. This way the rigid body is treated as constrained mechanical system. Moreover, we show that the EM discretization of constrained mechanical systems can be derived in a straightforward way from the EM scheme for the Cosserat point. The resulting rigid body formulation is closely connected to natural coordinates. Eventually, we deal with the extension to multibody systems which can be done in a straightforward way due to the presence of holonomic constraints in the present rigid body formulation.

Funding for this work has been provided by the German Science Foundation under Grant BE 2285/10-1.

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 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
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

Notes

  1. 1.

    If the external loads or part of them can be derived from an associated potential energy function \(V_\mathrm {ext}\) their contribution to the balance of energy can be shifted to the left-hand side of (47) by replacing U in (49) with \(U+V_\mathrm {ext}\).

  2. 2.

    The present framework comprises as well domain decomposition problems (Hesch and Betsch 2010) and large deformation contact (Hesch and Betsch 2009, 2011a, b).

References

  • Antman, S. S. (2005). Nonlinear problems of elasticity (2nd ed.). Springer.

    Google Scholar 

  • Armero, F., & Petöcz, E. (1998). Formulation and analysis of conserving algorithms for frictionless dynamic contact/impact problems. Computer Methods in Applied Mechanics and Engineering, 158, 269–300.

    Article  MathSciNet  MATH  Google Scholar 

  • Armero, F., & Romero, I. (2001). On the formulation of high-frequency dissipative time-stepping algorithms for nonlinear dynamics. Part II: Second-order methods. Computer Methods in Applied Mechanics and Engineering, 190, 6783–6824.

    Article  MathSciNet  MATH  Google Scholar 

  • Armero, F., & Romero, I. (2003). Energy-dissipative momentum-conserving time-stepping algorithms for the dynamics of nonlinear Cosserat rods. Computational Mechanics, 31, 3–26.

    Article  MathSciNet  MATH  Google Scholar 

  • Ascher, U. M., & Petzold, L. R. (1998). Computer methods for ordinary differential equations and differential-algebraic equations. SIAM.

    Google Scholar 

  • Bauchau, O. A. (2011). Flexible multibody dynamics. Solid mechanics and its applications New York: Springer.

    Google Scholar 

  • Bauchau, O. A., & Bottasso, C. L. (1999). On the design of energy preserving and decaying schemes for flexible, nonlinear multi-body systems. Computer Methods in Applied Mechanics and Engineering, 169(1–2), 61–79.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P. (2006). Energy-consistent numerical integration of mechanical systems with mixed holonomic and nonholonomic constraints. Computer Methods in Applied Mechanics and Engineering, 195, 7020–7035.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P., & Leyendecker, S. (2006). The discrete null space method for the energy consistent integration of constrained mechanical systems. Part II: Multibody dynamics. International Journal for Numerical Methods in Engineering, 67(4), 499–552.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P., & Sänger, N. (2009a). On the use of geometrically exact shells in a conserving framework for flexible multibody dynamics. Computer Methods in Applied Mechanics and Engineering, 198, 1609–1630.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P., & Sänger, N. (2009b). A nonlinear finite element framework for flexible multibody dynamics: Rotationless formulation and energy-momentum conserving discretization. In C. L. Bottasso (Ed.), Multibody Dynamics: Computational Methods and Applications, Computational Methods in Applied Sciences (Vol. 12, pp. 119–141). Springer.

    Google Scholar 

  • Betsch, P., & Sänger, N. (2013). On the consistent formulation of torques in a rotationless framework for multibody dynamics. Computers & Structures, 127, 29–38.

    Article  Google Scholar 

  • Betsch, P., & Steinmann, P. (2002a). Conservation properties of a time FE method. Part III: Mechanical systems with holonomic constraints. International Journal for Numerical Methods in Engineering, 53, 2271–2304.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P., & Steinmann, P. (2002b). Frame-indifferent beam finite elements based upon the geometrically exact beam theory. International Journal for Numerical Methods in Engineering, 54, 1775–1788.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P., & Steinmann, P. (2002c). A DAE approach to flexible multibody dynamics. Multibody System Dynamics, 8, 367–391.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P., & Steinmann, P. (2003). Constrained dynamics of geometrically exact beams. Computational Mechanics, 31, 49–59.

    Article  MATH  Google Scholar 

  • Betsch, P., & Uhlar, S. (2007). Energy-momentum conserving integration of multibody dynamics. Multibody System Dynamics, 17(4), 243–289.

    Article  MathSciNet  MATH  Google Scholar 

  • Betsch, P., Hesch, C., Sänger, N., & Uhlar, S. (2010). Variational integrators and energy-momentum schemes for flexible multibody dynamics. Journal of Computational and Nonlinear Dynamics, 5(3), 031001/1-11.

    Article  Google Scholar 

  • Betsch, P., Siebert, R., & Sänger, N. (2012). Natural coordinates in the optimal control of multibody systems. Journal of Computational and Nonlinear Dynamics, 7(1), 011009/1-8.

    Article  Google Scholar 

  • Bottasso, C. L., & Croce, A. (2004). Optimal control of multibody systems using an energy preserving direct transcription method. Multibody System Dynamics, 12(1), 17–45.

    Article  MathSciNet  MATH  Google Scholar 

  • Bottasso, C. L., & Trainelli, L. (2004). An attempt at the classification of energy decaying schemes for structural and multibody dynamics. Multibody System Dynamics, 12(2), 173–185.

    Article  MathSciNet  MATH  Google Scholar 

  • Bottasso, C. L., Borri, M., & Trainelli, L. (2001). Integration of elastic multibody systems by invariant conserving/dissipating algorithms. II. Numerical schemes and applications. Computer Methods in Applied Mechanics and Engineering, 190, 3701–3733.

    Article  MathSciNet  MATH  Google Scholar 

  • Bottasso, C. L., Bauchau, O. A., & Choi, J.-Y. (2002). An energy decaying scheme for nonlinear dynamics of shells. Computer Methods in Applied Mechanics and Engineering, 191(27–28), 3099–3121.

    Article  MathSciNet  MATH  Google Scholar 

  • Bottema, O., & Roth, B. (1979). Theoretical Kinematics. Amsterdam: North-Holland Publishing Company.

    MATH  Google Scholar 

  • Brank, B., Briseghella, L., Tonello, N., & Damjanic, F. B. (1998). On non-linear dynamics of shells: Implementation of energy-momentum conserving algorithm for a finite rotation shell model. International Journal for Numerical Methods in Engineering, 42, 409–442.

    Article  MathSciNet  MATH  Google Scholar 

  • Cohen, H., & Muncaster, R. G. (1988). The theory of Pseudo-rigid Bodies. New York: Springer.

    Book  MATH  Google Scholar 

  • Conde Martín, S., Betsch, P., & García Orden, J. C. (2016). A temperature-based thermodynamically consistent integration scheme for discrete thermo-elastodynamics. Communications in Nonlinear Science and Numerical Simulation, 32, 63–80.

    Article  Google Scholar 

  • Crisfield, M. A. (1997). Non-linear finite element analysis of solids and structures. Advanced topics. New York: Wiley.

    Google Scholar 

  • Crisfield, M. A., & Shi, J. (1994). A co-rotational element/time-integration strategy for non-linear dynamics. International Journal for Numerical Methods in Engineering, 37, 1897–1913.

    Article  MathSciNet  MATH  Google Scholar 

  • de García Jalón, J. (2007). Twenty-five years of natural coordinates. Multibody System Dynamics, 18(1), 15–33.

    Google Scholar 

  • de Jalón, J. G., & Bayo, E. (1994). Kinematic and dynamic simulation of multibody systems: The real-time challenge. Springer.

    Google Scholar 

  • Géradin, M. G., & Cardona, A. (2001). Flexible multibody dynamics: A finite element approach. Wiley.

    Google Scholar 

  • Gonzalez, O. (1996). Time integration and discrete Hamiltonian systems. Journal of Nonlinear Science, 6, 449–467.

    Article  MathSciNet  MATH  Google Scholar 

  • Gonzalez, O., & Simo, J. C. (1996). On the stability of symplectic and energy-momentum algorithms for non-linear Hamiltonian systems with symmetry. Computer Methods in Applied Mechanics and Engineering, 134, 197–222.

    Article  MathSciNet  MATH  Google Scholar 

  • Greenspan, D. (1984). Conservative numerical methods for \(\ddot{x}=f(x)\). Journal of Computational Physics, 56, 28–41.

    Article  MathSciNet  MATH  Google Scholar 

  • Groß, M., & Betsch, P. (2011). Galerkin-based energy-momentum consistent time-stepping algorithms for classical nonlinear thermo-elastodynamics. Mathematics and Computers in Simulation, 82(4), 718–770.

    Article  MathSciNet  MATH  Google Scholar 

  • Groß, M., & Betsch, P. (2010). Energy-momentum consistent finite element discretization of dynamic finite viscoelasticity. International Journal for Numerical Methods in Engineering, 81(11), 1341–1386.

    MathSciNet  MATH  Google Scholar 

  • Groß, M., Betsch, P., & Steinmann, P. (2005). Conservation properties of a time FE method. Part IV: Higher order energy and momentum conserving. International Journal for Numerical Methods in Engineering, 63, 1849–1897.

    Article  MathSciNet  MATH  Google Scholar 

  • Gurtin, M. E. (1981). An introduction to continuum mechanics. Academic Press.

    Google Scholar 

  • Hesch, C., & Betsch, P. (2011a). Transient 3d contact problems-NTS method: mixed methods and conserving integration. Computational Mechanics, 48(4), 437–449.

    Article  MathSciNet  MATH  Google Scholar 

  • Hesch, C., & Betsch, P. (2010). Transient three-dimensional domain decomposition problems: Frame-indifferent mortar constraints and conserving integration. International Journal for Numerical Methods in Engineering, 82(3), 329–358.

    MathSciNet  MATH  Google Scholar 

  • Hesch, C., & Betsch, P. (2011b). Transient three-dimensional contact problems: mortar method. Mixed methods and conserving integration. Computational Mechanics, 48(4), 461–475.

    Article  MathSciNet  MATH  Google Scholar 

  • Hesch, C., & Betsch, P. (2009). A mortar method for energy-momentum conserving schemes in frictionless dynamic contact problems. International Journal for Numerical Methods in Engineering, 77(10), 1468–1500.

    Article  MathSciNet  MATH  Google Scholar 

  • Hesch, C., & Betsch, P. (2011c). Energy-momentum consistent algorithms for dynamic thermomechanical problems—application to mortar domain decomposition problems. International Journal for Numerical Methods in Engineering, 86(11), 1277–1302.

    Article  MathSciNet  MATH  Google Scholar 

  • Hughes, T. J. R. (2000). The Finite element method. Dover Publications.

    Google Scholar 

  • Hughes, T. J. R., & Winget, J. (1980). Finite rotation effects in numerical integration of rate constitutive equations arising in large-deformation analysis. International Journal for Numerical Methods in Engineering, 15, 1862–1867.

    Article  MathSciNet  MATH  Google Scholar 

  • Hughes, T. J. R., Caughey, T. K., & Liu, W. K. (1978). Finite-element methods for nonlinear elastodynamics which conserve energy. Journal of Applied Mechanics, 45, 366–370.

    Article  MATH  Google Scholar 

  • Ibrahimbegović, A. (2009). Nonlinear solid mechanics. Solid mechanics and its applications (Vol. 160). Springer.

    Google Scholar 

  • Ibrahimbegović, A., Mamouri, S., Taylor, R. L., & Chen, A. J. (2000). Finite element method in dynamics of flexible multibody systems: Modeling of holonomic constraints and energy conserving integration schemes. Multibody System Dynamics, 4(2–3), 195–223.

    Article  MathSciNet  MATH  Google Scholar 

  • Johnson, E. R., & Murphey, T. D. (2009). Scalable variational integrators for constrained mechanical systems in generalized coordinates. IEEE Transactions on Robotics, 25(6), 1249–1261.

    Article  Google Scholar 

  • Koch, M. W., & Leyendecker, S. (2013). Energy momentum consistent force formulation for the optimal control of multibody systems. Multibody System Dynamics, 29, 381–401.

    Article  MathSciNet  Google Scholar 

  • Krenk, S. (2009). Non-linear modeling and analysis of solids and structures. Cambridge University Press.

    Google Scholar 

  • Kuhl, D., & Crisfield, M. A. (1999). Energy-conserving and decaying algorithms in non-linear structural mechanics. International Journal for Numerical Methods in Engineering, 45, 569–599.

    Article  MathSciNet  MATH  Google Scholar 

  • Kunkel, P., & Mehrmann, V. (2006). Differential-algebraic equations. European Mathematical Society.

    Google Scholar 

  • Laursen, T. A. (2002). Computational contact and impact mechanics. Springer.

    Google Scholar 

  • Laursen, T. A., & Chawla, V. (1997). Design of energy conserving algorithms for frictionless dynamic contact problems. International Journal for Numerical Methods in Engineering, 40, 863–886.

    Article  MathSciNet  MATH  Google Scholar 

  • Lens, E., & Cardona, A. (2007). An energy preserving/decaying scheme for nonlinearly constrained multibody systems. Multibody System Dynamics, 18(3), 435–470.

    Article  MathSciNet  MATH  Google Scholar 

  • Lens, E. V., Cardona, A., & Géradin, M. (2004). Energy preserving time integration for constrained multibody systems. Multibody System Dynamics, 11(1), 41–61.

    Article  MathSciNet  MATH  Google Scholar 

  • Lewis, D., & Simo, J. C. (1994). Conserving algorithms for the dynamics of Hamiltonian systems on Lie groups. Journal of Nonlinear Science, 4, 253–299.

    Article  MathSciNet  MATH  Google Scholar 

  • Leyendecker, S., Betsch, P., & Steinmann, P. (2006). Objective energy-momentum conserving integration for the constrained dynamics of geometrically exact beams. Computer Methods in Applied Mechanics and Engineering, 195, 2313–2333.

    Article  MathSciNet  MATH  Google Scholar 

  • Leyendecker, S., Betsch, P., & Steinmann, P. (2008a). The discrete null space method for the energy consistent integration of constrained mechanical systems. Part III: Flexible multibody dynamics. Multibody System Dynamics, 19(1–2), 45–72.

    Article  MathSciNet  MATH  Google Scholar 

  • Leyendecker, S., Marsden, J. E., & Ortiz, M. (2008b). Variational integrators for constrained dynamical systems. Zeitschrift für Angewandte Mathematik und Mechanik (ZAMM), 88(9), 677–708.

    Article  MathSciNet  MATH  Google Scholar 

  • Leyendecker, S., Ober-Blöbaum, S., Marsden, J. E., & Ortiz, M. (2010). Discrete mechanics and optimal control for constrained systems. Optimal Control Applications and Methods, 31(6), 505–528.

    Article  MathSciNet  MATH  Google Scholar 

  • Marsden, J. E., & Ratiu, T. S. (1999). Introduction to mechanics and symmetry (2nd ed.). Springer.

    Google Scholar 

  • McPhee, J. J., & Redmond, S. M. (2006). Modelling multibody systems with indirect coordinates. Computer Methods in Applied Mechanics and Engineering, 195, 6942–6957.

    Article  MathSciNet  MATH  Google Scholar 

  • Nordenholz, T. R., & O’Reilly, O. M. (1998). On steady motions of isotropic, elastic Cosserat points. IMA Journal of Applied Mathematics, 60, 55–72.

    Article  MathSciNet  MATH  Google Scholar 

  • Ober-Blöbaum, S., Junge, O., & Marsden, J. E. (2011). Discrete mechanics and optimal control: An analysis. ESAIM: Control, Optimisation and Calculus of Variations, 17(2), 322–352.

    Article  MathSciNet  MATH  Google Scholar 

  • Romero, I. (2009). Thermodynamically consistent time-stepping algorithms for non-linear thermomechanical systems. International Journal for Numerical Methods in Engineering, 79(6), 706–732.

    Article  MathSciNet  MATH  Google Scholar 

  • Romero, I. (2010). Algorithms for coupled problems that preserve symmetries and the laws of thermodynamics: Part II: Fractional step methods. Computer Methods in Applied Mechanics and Engineering, 199(33–36), 2235–2248.

    Article  MathSciNet  MATH  Google Scholar 

  • Romero, I. (2012). An analysis of the stress formula for energy-momentum methods in nonlinear elastodynamics. Computational Mechanics, 50, 603–610.

    Article  MathSciNet  MATH  Google Scholar 

  • Romero, I., & Armero, F. (2002a). 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. International Journal for Numerical Methods in Engineering, 54, 1683–1716.

    Article  MathSciNet  MATH  Google Scholar 

  • Romero, I., & Armero, F. (2002b). Numerical integration of the stiff dynamics of geometrically exact shells: an energy-dissipative momentum-conserving scheme. International Journal for Numerical Methods in Engineering, 54, 1043–1086.

    Article  MathSciNet  MATH  Google Scholar 

  • Rubin, M. B. (2000). Cosserat theories: shells, rods and points, solid mechanics and its applications (Vol. 79). Kluwer Academic Publishers.

    Google Scholar 

  • Simo, J. C., & Tarnow, N. (1992). The discrete energy-momentum method. Conserving algorithms for nonlinear elastodynamics. Zeitschrift für angewandte Mathematik und Physik (ZAMP), 43, 757–792.

    Article  MathSciNet  MATH  Google Scholar 

  • Simo, J. C., & Tarnow, N. (1994). A new energy and momentum conserving algorithm for the nonlinear dynamics of shells. International Journal for Numerical Methods in Engineering, 37, 2527–2549.

    Article  MathSciNet  MATH  Google Scholar 

  • Simo, J. C., & Wong, K. K. (1991). Unconditionally stable algorithms for rigid body dynamics that exactly preserve energy and momentum. International Journal for Numerical Methods in Engineering, 31, 19–52.

    Article  MathSciNet  MATH  Google Scholar 

  • Simo, J. C., Lewis, D., & Marsden, J. E. (1991). Stability of relative equilibria. Part I: The reduced energy-momentum method. Archive for Rational Mechanics and Analysis, 115, 15–59.

    Article  MathSciNet  MATH  Google Scholar 

  • Simo, J. C., Rifai, M. S., & Fox, D. D. (1992a). On a stress resultant geometrically exact shell model. Part VI: Conserving algorithms for non-linear dynamics. International Journal for Numerical Methods in Engineering, 34, 117–164.

    Article  MathSciNet  MATH  Google Scholar 

  • Simo, J. C., Tarnow, N., & Wong, K. K. (1992b). Exact energy-momentum conserving algorithms and symplectic schemes for nonlinear dynamics. Computer Methods in Applied Mechanics and Engineering, 100, 63–116.

    Article  MathSciNet  MATH  Google Scholar 

  • Tarnow, N. (1993). Energy and Momentum Conserving Algorithms for Hamiltonian Systems in the Nonlinear Dynamics of Solids. Ph.D. Dissertation, Sudam report no. 93–4. Stanford University.

    Google Scholar 

  • Truesdell, C., & Noll, W. (2004). The non-linear field theories of mechanics (3rd ed.). Springer (2004).

    Google Scholar 

Download references

Acknowledgments

Support for this research was provided by the Deutsche Forschungsgemeinschaft (DFG) under Grant BE 2285/10-1. This support is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Peter Betsch .

Editor information

Editors and Affiliations

A Appendix

A Appendix

1.1 A.1 Balance Laws

For comparison, the balance laws are directly derived from the variational equations (19) governing the motion of the pseudo-rigid body. To this end, we recast (19) in the form

$$\begin{aligned} \delta \overline{{\varvec{x}}} \cdot \left( M \ddot{\overline{{\varvec{x}}}} - {\varvec{f}}_\mathrm {ext} \right)&= 0 \end{aligned}$$
(130)
$$\begin{aligned} \text{ tr }\left( \delta {\varvec{F}}^T \big (\ddot{{\varvec{F}}}{\varvec{E}}_0 + 2{\varvec{F}} DU({\varvec{C}}) - {\varvec{M}}_\mathrm {ext}\big ) \right)&= 0 \end{aligned}$$
(131)

Applying the polar decomposition theorem to the deformation gradient, we get

$$\begin{aligned} {\varvec{F}} = {\varvec{R}}{\varvec{U}} \qquad \text{ and }\qquad \delta {\varvec{F}} = \delta {\varvec{R}}{\varvec{U}} + {\varvec{R}}\delta {\varvec{U}} \end{aligned}$$
(132)

Since \({\varvec{R}}{\varvec{R}}^T = {\varvec{I}}\), \(\delta {\varvec{R}}{\varvec{R}}^T + {\varvec{R}}\delta {\varvec{R}}^T = {\varvec{0}}\), and consequently

$$\begin{aligned} \widehat{{\varvec{\omega }}}_\delta = \delta {\varvec{R}}{\varvec{R}}^T \end{aligned}$$
(133)

is skew-symmetric. A straightforward calculation shows that (131) can be rewritten as

$$\begin{aligned} \text{ tr }\left( \delta {\varvec{U}} {\varvec{U}} \big ( {\varvec{F}}^{-1}\ddot{{\varvec{F}}}{\varvec{E}}_0 + 2 DU({\varvec{C}}) - {\varvec{F}}^{-1}{\varvec{M}}_\mathrm {ext}\big ) \right)&= 0 \end{aligned}$$
(134)
$$\begin{aligned} {\varvec{\omega }}_\delta \cdot \left( 2\text{ vect }\big (\ddot{{\varvec{F}}}{\varvec{E}}_0{\varvec{F}}^T\big ) - 2\text{ vect }\big ({\varvec{M}}_\mathrm {ext}{\varvec{F}}^T\big )\right)&= 0 \end{aligned}$$
(135)

Accordingly, the nine independent equations emanating from (131) have been converted to six independent equations (134) plus three independent equations (135).

In (135), \(\text{ vect }(\bullet )\) denotes the vector invariant of a second-order tensor defined by

$$\begin{aligned} \text{ vect }({\varvec{a}}\otimes {\varvec{b}})\times {\varvec{c}} = \text{ skew }({\varvec{a}}\otimes {\varvec{b}}){\varvec{c}} \end{aligned}$$

Since

$$\begin{aligned} \text{ skew }({\varvec{a}}\otimes {\varvec{b}}){\varvec{c}}&=\frac{1}{2}({\varvec{a}}\otimes {\varvec{b}} - {\varvec{b}}\otimes {\varvec{a}}) {\varvec{c}} \\&=\frac{1}{2}\left( ({\varvec{b}}\cdot {\varvec{c}}){\varvec{a}} - ({\varvec{a}}\cdot {\varvec{c}}){\varvec{b}}\right) \\&=\frac{1}{2}({\varvec{b}}\times {\varvec{a}})\times {\varvec{c}} \end{aligned}$$

we have

$$\begin{aligned} \text{ vect }({\varvec{a}}\otimes {\varvec{b}}) = \frac{1}{2}({\varvec{b}}\times {\varvec{a}}) \end{aligned}$$
(136)

Accordingly,

$$\begin{aligned} 2\text{ vect }\big (\ddot{{\varvec{F}}}{\varvec{E}}_0{\varvec{F}}^T\big )&= 2\text{ vect }\big (E_0^{ij}\ddot{{\varvec{d}}}_i\otimes {\varvec{d}}_j\big )= E_0^{ij}{\varvec{d}}_j\times \ddot{{\varvec{d}}}_i \end{aligned}$$
(137)
$$\begin{aligned} 2\text{ vect }\big ({\varvec{M}}_\mathrm {ext}{\varvec{F}}^T\big )&= 2\text{ vect }\big ({\varvec{f}}^i_\mathrm {ext}\otimes {\varvec{d}}_i\big )= {\varvec{d}}_i\times {\varvec{f}}^i_\mathrm {ext} = \overline{{\varvec{m}}}_\mathrm {ext} \end{aligned}$$
(138)

1.1.1 A.1.1 Balance of Angular Momentum

To get the balance law for angular momentum, substitute \(\delta {{\varvec{U}}} = {{\varvec{0}}}\) into (134), \({\varvec{\omega }}_\delta ={\varvec{\xi }}\) into (135), and \(\delta \overline{{\varvec{x}}}={\varvec{\xi }}\times \overline{{\varvec{x}}}\) into (130). Subsequent summation of the resulting equations yields

$$\begin{aligned} {\varvec{\xi }}\cdot \left( M\overline{{\varvec{x}}}\times \ddot{\overline{{\varvec{x}}}} + 2\text{ vect }\big (\ddot{{\varvec{F}}}{\varvec{E}}_0{\varvec{F}}^T\big ) -\overline{{\varvec{x}}}\times {\varvec{f}}_\mathrm {ext} - 2\text{ vect }\big ({\varvec{M}}_\mathrm {ext}{\varvec{F}}^T\big ) \right) = 0 \end{aligned}$$

or

$$\begin{aligned} {\varvec{\xi }}\cdot \left( \frac{d}{dt} {\varvec{j}} - {\varvec{m}}_\mathrm {ext} \right) = 0 \end{aligned}$$

The last equation has to hold for arbitrary \({\varvec{\xi }}\in \mathbb {R}^3\). Accordingly, one obtains \(d{{\varvec{j}}}/dt={\varvec{m}}_\mathrm {ext}\), where

$$\begin{aligned} {{\varvec{j}}}&= M\overline{{\varvec{x}}}\times \dot{\overline{{\varvec{x}}}} + 2\text{ vect }\big (\dot{{\varvec{F}}}{\varvec{E}}_0{\varvec{F}}^T\big ) \\ {{\varvec{m}}}_\mathrm {ext}&= \overline{{\varvec{x}}}\times {\varvec{f}}_\mathrm {ext} + 2\text{ vect }\big ({\varvec{M}}_\mathrm {ext}{\varvec{F}}^T\big ) \end{aligned}$$

denote, respectively, the total angular momentum and the resultant external torque with respect to the origin of the inertial frame of reference. Note that the same conclusions can be drawn by substituting \(\delta \overline{{\varvec{x}}}={\varvec{\xi }}\times \overline{{\varvec{x}}}\) into (130), and \(\delta {\varvec{F}} = \widehat{{\varvec{\xi }}}{\varvec{F}}\) into (131).

1.1.2 A.1.2 Balance of Energy

Suppose that an external force \({\varvec{f}}_\mathrm {ext}\in \mathbb {R}^3\) along with external director forces \({\varvec{f}}^i_\mathrm {ext}\in \mathbb {R}^3\), \(i=1,2,3\), are acting on the body under consideration. Recall that the external director forces \({\varvec{f}}^i_\mathrm {ext}\) can be linked to the second-order tensor \({\varvec{M}}_\mathrm {ext}\) via \({\varvec{M}}_\mathrm {ext} = {\varvec{f}}^i_\mathrm {ext} \otimes {\varvec{D}}_i\) (see Eq. (38) in Sect. 2.2). To define the external director forces we introduce 9 independent quantities \(\mathcal {M}^{ij}\) such that

$$\begin{aligned} {\varvec{\mathcal {M}}} = \mathcal {M}^{ij} {\varvec{D}}_i \otimes {\varvec{D}}_j \end{aligned}$$
(139)

and

$$\begin{aligned} {\varvec{M}}_\mathrm {ext} = {\varvec{F}}{\varvec{\mathcal {M}}} = \mathcal {M}^{ij} {\varvec{d}}_i \otimes {\varvec{D}}_j \end{aligned}$$
(140)

Note that the last equation implies

$$\begin{aligned} {\varvec{f}}^i_\mathrm {ext} = \mathcal {M}^{ki} {\varvec{d}}_k \end{aligned}$$
(141)

Now substitute \(\dot{\overline{{\varvec{x}}}}\) for \(\delta \overline{{\varvec{x}}}\) into (130) and \(\dot{{\varvec{F}}}\) for \(\delta {\varvec{F}}\) into (131). Subsequent summation of both equations yields

$$\begin{aligned} \dot{\overline{{\varvec{x}}}} \cdot \left( M \ddot{\overline{{\varvec{x}}}} - {\varvec{f}}_\mathrm {ext} \right) + \text{ tr }\left( \dot{{\varvec{F}}}^T \big (\ddot{{\varvec{F}}}{\varvec{E}}_0 + 2{\varvec{F}} DU({\varvec{C}}) - {\varvec{M}}_\mathrm {ext}\big ) \right) = 0 \end{aligned}$$
(142)

Taking into account the relationships

$$\begin{aligned} \dot{\overline{{\varvec{x}}}} \cdot M \ddot{\overline{{\varvec{x}}}}&= \frac{d}{dt} \left( \frac{1}{2} M\dot{\overline{{\varvec{x}}}}\cdot \dot{\overline{{\varvec{x}}}} \right) \\ \text{ tr }\left( \dot{{\varvec{F}}}^T\ddot{{\varvec{F}}}{\varvec{E}}_0 \right)&= \frac{d}{dt}\left( \frac{1}{2} \text{ tr }\left( \dot{{\varvec{F}}}^T\dot{{\varvec{F}}}{\varvec{E}}_0 \right) \right) \end{aligned}$$

we define the kinetic energy

$$\begin{aligned} T = \frac{1}{2} M \dot{\overline{{\varvec{x}}}}\cdot \dot{\overline{{\varvec{x}}}}+\frac{1}{2}\text{ tr }\left( \dot{{\varvec{F}}}{\varvec{E}}_0\dot{{\varvec{F}}}^T\right) \end{aligned}$$
(143)

Moreover,

$$\begin{aligned} \text{ tr }\left( \dot{{\varvec{F}}}^T{\varvec{F}} 2DU({\varvec{C}})\right)&= \text{ tr }\left( 2DU({\varvec{C}}) \text{ sym }(\dot{{\varvec{F}}}^T{\varvec{F}})\right) \\&= \text{ tr }\left( 2DU({\varvec{C}}) \frac{1}{2}(\dot{{\varvec{F}}}^T{\varvec{F}} + {\varvec{F}}^T\dot{{\varvec{F}}}) \right) \\&= \text{ tr }\left( 2DU({\varvec{C}}) \frac{1}{2} \dot{{\varvec{C}}} \right) \\&= \frac{d}{dt} U({\varvec{C}}) \end{aligned}$$

Now (142) can be recast in the form

$$\begin{aligned} \frac{d}{dt} E = P_{\mathrm {ext}} \end{aligned}$$
(144)

Here, E is the total mechanical energy given by

$$\begin{aligned} E = T + U \end{aligned}$$

where U denotes the total strain energy defined in (13). On the right hand side of balance equation (144)

$$\begin{aligned} P_{\mathrm {ext}} = {\varvec{f}}_\mathrm {ext}\cdot \dot{\overline{{\varvec{x}}}} + \text{ tr }\left( \dot{{\varvec{F}}}^T {\varvec{M}}_\mathrm {ext} \right) \end{aligned}$$

denotes the power of the external forces acting on the pseudo-rigid body. We next focus on the power of the director forces given by

$$\begin{aligned} \overline{P}_{\mathrm {ext}} = \text{ tr }\left( \dot{{\varvec{F}}}^T {\varvec{M}}_\mathrm {ext} \right) \end{aligned}$$

Taking into account (140), the last equation can be rewritten as

$$\begin{aligned} \overline{P}_{\mathrm {ext}}&= \text{ tr }\left( \dot{{\varvec{F}}}^T{\varvec{F}}{\varvec{\mathcal {M}}} \right) \\&= \text{ tr }\left( \big (\overline{{\varvec{\mathcal {M}}}} + \widetilde{{\varvec{\mathcal {M}}}}\big )\dot{{\varvec{F}}}^T{\varvec{F}} \right) \end{aligned}$$

In the last equation

$$\begin{aligned} \overline{{\varvec{\mathcal {M}}}}&= \text{ sym }\left( {\varvec{\mathcal {M}}} \right) \\ \widetilde{{\varvec{\mathcal {M}}}}&= \text{ skew }\left( {\varvec{\mathcal {M}}} \right) \end{aligned}$$

have been introduced. Now

$$\begin{aligned} \text{ tr }\left( \overline{{\varvec{\mathcal {M}}}}\dot{{\varvec{F}}}^T{\varvec{F}}\right) =\text{ tr }\left( \overline{{\varvec{\mathcal {M}}}}\text{ sym }\left( \dot{{\varvec{F}}}^T{\varvec{F}} \right) \right) =\frac{1}{2}\text{ tr }\left( \overline{{\varvec{\mathcal {M}}}}\dot{{\varvec{C}}}\right) \end{aligned}$$

Furthermore,

$$\begin{aligned} \text{ tr }\left( \widetilde{{\varvec{\mathcal {M}}}}\dot{{\varvec{F}}}^T{\varvec{F}}\right) =\text{ tr }\left( \widetilde{{\varvec{\mathcal {M}}}}\text{ skew }\left( \dot{{\varvec{F}}}^T{\varvec{F}} \right) \right) \end{aligned}$$

Applying the polar decomposition \({\varvec{F}} = {\varvec{R}}{\varvec{U}}\) along with \(\dot{{\varvec{F}}} = \dot{{\varvec{R}}}{\varvec{U}} + {\varvec{R}}\dot{{\varvec{U}}}\) and \(\widehat{{\varvec{\omega }}} = \dot{{\varvec{R}}}{\varvec{R}}^T\) (cf. (132) and (133) on p. 47), we get

$$\begin{aligned} \text{ tr }\left( \widetilde{{\varvec{\mathcal {M}}}}\text{ skew }\left( \dot{{\varvec{F}}}^T{\varvec{F}} \right) \right) = {\varvec{\omega }}\cdot 2\text{ vect }\left( {\varvec{F}}\widetilde{{\varvec{\mathcal {M}}}}{\varvec{F}}^T\right) +\text{ tr }\left( \widetilde{{\varvec{\mathcal {M}}}}\dot{{\varvec{U}}}^T{\varvec{U}} \right) \end{aligned}$$

Altogether the power of the external forces can be written in the form

$$\begin{aligned} P_{\mathrm {ext}}&= {\varvec{f}}_\mathrm {ext}\cdot \dot{\overline{{\varvec{x}}}} + \frac{1}{2}\text{ tr }\left( \overline{{\varvec{\mathcal {M}}}}\dot{{\varvec{C}}}\right) + {\varvec{\omega }}\cdot 2\text{ vect }\left( {\varvec{F}}\widetilde{{\varvec{\mathcal {M}}}}{\varvec{F}}^T\right) + \text{ tr }\left( \widetilde{{\varvec{\mathcal {M}}}}\dot{{\varvec{U}}}^T{\varvec{U}} \right) \\&= {\varvec{f}}_\mathrm {ext}\cdot \dot{\overline{{\varvec{x}}}} + \frac{1}{2}\overline{\mathcal {M}}^{ij}\dot{d}_{ij} + {\varvec{\omega }}\cdot \overline{{\varvec{m}}}_\mathrm {ext} + \text{ tr }\left( \widetilde{{\varvec{\mathcal {M}}}}\dot{{\varvec{U}}}^T{\varvec{U}} \right) \end{aligned}$$

Here, \(\overline{{\varvec{m}}}_\mathrm {ext}\) can be identified as the resultant external torque relative to the center of mass that has been introduced in (71). In particular, we have

$$\begin{aligned} \overline{{\varvec{m}}}_\mathrm {ext}&= 2\text{ vect }\left( {\varvec{F}}\widetilde{{\varvec{\mathcal {M}}}}{\varvec{F}}^T\right) \\&= \widetilde{\mathcal {M}}^{ji}{\varvec{d}}_i\times {\varvec{d}}_j \\&= \varepsilon _{ijk}\widetilde{\mathcal {M}}^{ji} {\varvec{d}}^k \\&= {\varvec{d}}_i\times {\varvec{f}}^i_\mathrm {ext} \end{aligned}$$

In the last equation use has been made of (141). Moreover,

$$\begin{aligned} \varepsilon _{ijk} = ({\varvec{d}}_i\times {\varvec{d}}_j)\cdot {\varvec{d}}_k = e_{ijk}\sqrt{d} \end{aligned}$$

where \(d=\det (d_{ij})\) (or \(\sqrt{d}=({\varvec{d}}_1\times {\varvec{d}}_2)\cdot {\varvec{d}}_3\)) and \(e_{ijk}\) denotes the alternating symbol.

1.2 A.2 Application of External Torques

It can be observed from the above treatment that the application of external torques \(\overline{{\varvec{m}}}_\mathrm {ext}\) relative to the center of mass is linked to the skew-symmetric tensor \(\widetilde{{\varvec{\mathcal {M}}}}=\widetilde{\mathcal {M}}^{ij}{\varvec{D}}_i\otimes {\varvec{D}}_j\). In particular, given the covariant components of the external torque, \(m_k={\varvec{d}}_k\cdot \overline{{\varvec{m}}}_\mathrm {ext}\), we obtain

$$\begin{aligned} m_k = \varepsilon _{ijk}\widetilde{\mathcal {M}}^{ji} \end{aligned}$$

from which it follows that

$$\begin{aligned} \widetilde{\mathcal {M}}^{23}&=-\widetilde{\mathcal {M}}^{32}=-\frac{m_1}{2\sqrt{d}} \\ \widetilde{\mathcal {M}}^{31}&=-\widetilde{\mathcal {M}}^{13}=-\frac{m_2}{2\sqrt{d}} \\ \widetilde{\mathcal {M}}^{12}&=-\widetilde{\mathcal {M}}^{21}=-\frac{m_3}{2\sqrt{d}} \end{aligned}$$

or

$$\begin{aligned} \widetilde{\mathcal {M}}^{ij} = \frac{e^{jik} m_k}{2\sqrt{d}} \end{aligned}$$
(145)

where \(e^{ijk}=e_{ijk}\) again denotes the alternating symbol. Accordingly, using the above formulas for \(\widetilde{\mathcal {M}}^{ij}\) in terms of the torque components \(m_k\), the corresponding director forces can be calculated via

$$\begin{aligned} {\varvec{f}}^j_\mathrm {ext} = \widetilde{\mathcal {M}}^{ij} {\varvec{d}}_i \end{aligned}$$
(146)

or

$$\begin{aligned} {\varvec{f}}^j_\mathrm {ext} = \frac{e^{ijk}}{2\sqrt{d}}\left( {\varvec{d}}_j\otimes {\varvec{d}}_k\right) \overline{{\varvec{m}}}_\mathrm {ext} \end{aligned}$$
(147)

To summarize, the action of an external torque \(\overline{{\varvec{m}}}_\mathrm {ext}\) relative to the center of mass can be realized by applying external director forces of the form

$$\begin{aligned} \begin{bmatrix} {\varvec{f}}^1_\mathrm {ext} \\ {\varvec{f}}^2_\mathrm {ext} \\ {\varvec{f}}^3_\mathrm {ext} \end{bmatrix} = \frac{1}{2\sqrt{d}} \begin{bmatrix} {\varvec{d}}_2\otimes {\varvec{d}}_3 - {\varvec{d}}_3\otimes {\varvec{d}}_2 \\ {\varvec{d}}_3\otimes {\varvec{d}}_1 - {\varvec{d}}_1\otimes {\varvec{d}}_3 \\ {\varvec{d}}_1\otimes {\varvec{d}}_2 - {\varvec{d}}_2\otimes {\varvec{d}}_1 \end{bmatrix} \overline{{\varvec{m}}}_\mathrm {ext} \end{aligned}$$
(148)

Remark A.1

Formula (148) can be viewed as an extension to flexible Cosserat points of the method proposed in Betsch and Sänger (2013). In this work the consistent application of torques has been dealt with in the context of rigid body dynamics formulated in terms of directors (or direction cosines). The formula proposed in Betsch and Sänger (2013) is given by

$$\begin{aligned} {\varvec{f}}^j_\mathrm {ext} = \frac{1}{2} \overline{{\varvec{m}}}_\mathrm {ext}\times {\varvec{d}}^j \end{aligned}$$
(149)

The equivalence of (149) to (148) can be shown by a direct calculation:

$$\begin{aligned} {\varvec{f}}^j_\mathrm {ext}&= \frac{1}{2} \overline{{\varvec{m}}}_\mathrm {ext}\times {\varvec{d}}^j \\&= \frac{1}{2} m_k {\varvec{d}}^k\times {\varvec{d}}^j \\&= \frac{1}{2} m_k d^{-\frac{1}{2}} e^{kji}{\varvec{d}}_i \\&= \widetilde{\mathcal {M}}^{ij} {\varvec{d}}_i \end{aligned}$$

where (145) has been used.

1.2.1 A.2.1 Fully Actuated Cosserat Point

If the Cosserat point shall be fully actuated, the 9 independent quantities \(\mathcal {M}^{ij}\) in (139) can be employed as control inputs. According to (141) this approach determines the external director forces

$$\begin{aligned} {\varvec{f}}^i_\mathrm {ext} = \mathcal {M}^{ji} {\varvec{d}}_j \end{aligned}$$

If required the external torque associated with the control inputs can be extracted via

$$\begin{aligned} \overline{{\varvec{m}}}_\mathrm {ext} = {\mathcal {M}}^{ji}{\varvec{d}}_i\times {\varvec{d}}_j \end{aligned}$$

Note that due to the presence of the cross product the skew-symmetric part of \(\mathcal {M}^{ji}\), that is, \(\widetilde{\mathcal {M}}^{ji}=(\mathcal {M}^{ji}-\mathcal {M}^{ij})/2\), is automatically extracted. The above result coincides with

$$\begin{aligned} \overline{{\varvec{m}}}_\mathrm {ext} = m_k{\varvec{d}}^k \quad \text{ where }\quad m_k = \sqrt{d}\,e_{ijk}{\mathcal {M}}^{ji} \end{aligned}$$

Again the skew-symmetric part of \(\mathcal {M}^{ji}\) is extracted due to the presence of the alternating symbol.

1.3 A.3 Iteration Matrix of the EM Integrator

Consider St. Venant-Kirchhoff material with strain energy density

$$\begin{aligned} W({\varvec{G}}) = \frac{\lambda }{2}\left( \text{ tr } {\varvec{G}}\right) ^2 + \mu \text{ tr }\left( {\varvec{G}}^2\right) \end{aligned}$$

where the Green–Lagrangean strain tensor is given by

$$\begin{aligned} {\varvec{G}} = \frac{1}{2}\left( {\varvec{C}} - {\varvec{I}}\right) = \gamma _{ij} {\varvec{D}}^i\otimes {\varvec{D}}^j \end{aligned}$$

Note that the components \(\gamma _{ij}=\frac{1}{2}(d_{ij}-\delta _{ij})\) have been introduced in (66). According to (11), the second Piola-Kirchhoff stress tensor is given by

$$\begin{aligned} \begin{array}{rcl} {\varvec{S}} &{}=&{} 2 DW({\varvec{C}}) \\ &{}=&{} DW({\varvec{G}}) \\ &{}=&{} \lambda \left( \text{ tr }{\varvec{G}}\right) {\varvec{I}} + 2\mu {\varvec{G}} \end{array} \end{aligned}$$
(150)

Moreover, the fourth-order elasticity tensor assumes the form

$$\begin{aligned} \begin{array}{rcl} {\varvec{\mathsf {C}}} &{}=&{} 4 D^2W({\varvec{C}}) \\ &{}=&{} D^2W({\varvec{G}}) \\ &{}=&{} \lambda {\varvec{I}}\otimes {\varvec{I}} + 2\mu {\varvec{\mathsf {I}}} \end{array} \end{aligned}$$
(151)

Since we have assumed that the director triad \(\{{\varvec{D}}_i\}\) in the reference configuration is orthonormal, it suffices to consider the Cartesian components of \({\varvec{S}}\) and \({\varvec{\mathsf {C}}}\). Accordingly, we have

$$\begin{aligned} S_{ij} = \lambda \gamma _{kk}\delta _{ij} + 2\mu \gamma _{ij} \end{aligned}$$
(152)

and

$$\begin{aligned} \mathsf {C}_{ijkl} = \lambda \delta _{ij}\delta _{kl} + \mu \left( \delta _{ik}\delta _{jl} + \delta _{il}\delta _{jk}\right) \end{aligned}$$

We next deal with the linearization of the internal director forces \({\varvec{f}}^i_\mathrm {int}=S^{ij}{\varvec{d}}_j\). First consider the time-continuous case where, according to the product rule of differentiation, we get

$$\begin{aligned} \Delta {\varvec{f}}^i_\mathrm {int}=\Delta S^{ij}{\varvec{d}}_j + S^{ij}\Delta {\varvec{d}}_j \end{aligned}$$
(153)

With regard to (152)

$$\begin{aligned} \begin{array}{rcl} \Delta S_{ij} &{}=&{} \lambda \Delta \gamma _{kk}\delta _{ij} + 2\mu \Delta \gamma _{ij} \\ &{}=&{} \lambda ({\varvec{d}}_k\cdot \Delta {\varvec{d}}_k) \delta _{ij} + \mu ({\varvec{d}}_i\cdot \Delta {\varvec{d}}_j + {\varvec{d}}_j\cdot \Delta {\varvec{d}}_i) \end{array} \end{aligned}$$
(154)

Now, a straightforward calculation yields

$$\begin{aligned} \Delta {\varvec{f}}^i_\mathrm {int}=\left( {\varvec{K}}^{ij}_\mathrm {mat} + {\varvec{K}}^{ij}_\mathrm {geo}\right) \Delta {\varvec{d}}_j \end{aligned}$$
(155)

where the contributions to the iteration matrix have been split into a material part \({\varvec{K}}^{ij}_\mathrm {mat}\) and a geometric part \({\varvec{K}}^{ij}_\mathrm {geo}\). The material part is given by

$$\begin{aligned} {\varvec{K}}^{ij}_\mathrm {mat} = \lambda {\varvec{d}}_i\otimes {\varvec{d}}_j + \mu {\varvec{d}}_j\otimes {\varvec{d}}_i + \mu {\varvec{d}}_k\otimes {\varvec{d}}_k \, \delta _{ij} \end{aligned}$$

and the geometric part assumes the form

$$\begin{aligned} {\varvec{K}}^{ij}_\mathrm {geo} = S_{ij}{\varvec{I}} \end{aligned}$$

The symmetry of the iteration matrix follows from the properties \({\varvec{K}}^{ij}_\mathrm {mat}=({\varvec{K}}^{ji}_\mathrm {mat})^T\) and \({\varvec{K}}^{ij}_\mathrm {geo}=({\varvec{K}}^{ji}_\mathrm {geo})^T\). Similar to (153), in the discrete case the EM scheme (63) leads to

$$\begin{aligned} \Delta \left. ({\varvec{f}}^i_\mathrm {int})\right| _{n+\frac{1}{2}}=\Delta {S}^{ij}_A{{\varvec{d}}}_{j_{n+\frac{1}{2}}} + {S}^{ij}_A\Delta {{\varvec{d}}}_{j_{n+\frac{1}{2}}} \end{aligned}$$

Similar to (154), the algorithmic stress formula (65) leads to

$$\begin{aligned} \begin{array}{rcl} \Delta {S}^{ij}_A &{}=&{} \lambda \Delta \gamma _{kk_{n+\frac{1}{2}}}\delta _{ij} + 2\mu \Delta \gamma _{ij_{n+\frac{1}{2}}} \\ &{}=&{} \frac{\lambda }{2}({\varvec{d}}_{k_{n+1}}\cdot \Delta {\varvec{d}}_{k_{n+1}}) \delta _{ij} + \frac{\mu }{2}({\varvec{d}}_{i_{n+1}}\cdot \Delta {\varvec{d}}_{j_{n+1}} + {\varvec{d}}_{j_{n+1}}\cdot \Delta {\varvec{d}}_{i_{n+1}}) \end{array} \end{aligned}$$
(156)

Altogether the discrete counterpart of (155) is given by the consistent linearization

$$\begin{aligned} \Delta \left. ({\varvec{f}}^i_\mathrm {int})\right| _{n+\frac{1}{2}}=\left( \left. {\varvec{K}}^{ij}_\mathrm {mat}\right| _{n+\frac{1}{2}} + \left. {\varvec{K}}^{ij}_\mathrm {geo}\right| _{n+\frac{1}{2}}\right) \Delta {\varvec{d}}_{j_{n+1}} \end{aligned}$$

where the material part is given by

$$\begin{aligned} \left. {\varvec{K}}^{ij}_\mathrm {mat}\right| _{n+\frac{1}{2}} = \frac{1}{2}\left( \lambda {\varvec{d}}_{i_{n+\frac{1}{2}}}\otimes {\varvec{d}}_{j_{n+1}} + \mu {\varvec{d}}_{j_{n+\frac{1}{2}}}\otimes {\varvec{d}}_{i_{n+1}} + \mu {\varvec{d}}_{k_{n+\frac{1}{2}}}\otimes {\varvec{d}}_{k_{n+1}} \, \delta _{ij}\right) \end{aligned}$$

and the geometric part assumes the form

$$\begin{aligned} \left. {\varvec{K}}^{ij}_\mathrm {geo}\right| _{n+\frac{1}{2}} = \frac{1}{2}{S}^{ij}_A{\varvec{I}} \end{aligned}$$

It is obvious that in the discrete setting the material part destroys the symmetry of the iteration matrix, for

$$\begin{aligned} \left. \left( {\varvec{K}}^{ij}_\mathrm {mat}\right) \right| _{n+\frac{1}{2}}\ne \left. \left( {\varvec{K}}^{ji}_\mathrm {mat}\right) ^T\right| _{n+\frac{1}{2}} \end{aligned}$$

We finally remark that due to definition (13) of the total strain energy of the Cosserat point, namely \(U({\varvec{C}}) = V_0 W({\varvec{C}})\), the above stress components \(S^{ij}\) should be replaced by \(\overline{S}^{ij}=V_0 S^{ij}\).

Rights and permissions

Reprints and permissions

Copyright information

© 2016 CISM International Centre for Mechanical Sciences

About this chapter

Cite this chapter

Betsch, P. (2016). Energy-Momentum Integrators for Elastic Cosserat Points, Rigid Bodies, and Multibody Systems. In: Betsch, P. (eds) Structure-preserving Integrators in Nonlinear Structural Dynamics and Flexible Multibody Dynamics. CISM International Centre for Mechanical Sciences, vol 565. Springer, Cham. https://doi.org/10.1007/978-3-319-31879-0_2

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-31879-0_2

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-319-31877-6

  • Online ISBN: 978-3-319-31879-0

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics