Skip to main content
Log in

Modeling and analysis of multiple impacts in multibody systems under unilateral and bilateral constrains based on linear projection operators

  • Published:
Multibody System Dynamics Aims and scope Submit manuscript

Abstract

This paper presents a unifying dynamics formulation for nonsmooth multibody systems (MBSs) subject to changing topology and multiple impacts based on a linear projection operator. An oblique projection matrix ubiquitously yields all characteristic variables of such systems as follows: (i) the constrained acceleration before jump discontinuity from the projection of unconstrained acceleration, (ii) post-impact velocity from the projection of pre-impact velocity, (iii) impulse during impact from the projection of pre-impact momentum, (iv) generalized constraint force from the projection of generalized input force, and (v) post-impact kinetic energy from pre-impact kinetic energy based on projected inertia matrix. All solutions are presented in closed-form with elegant geometrical interpretations. The formulation is general enough to be applicable to MBSs subject to simultaneous multiple impacts with nonidentical restitution coefficients, changing topology, i.e., unilateral constraints becoming inactive or vice versa, or even when the overall constraint Jacobian becomes singular. Not only do the solutions always exist regardless of the constraint condition, but also the condition number for a generalized constraint inertia matrix is minimized in order to reduce numerical sensitivity in computation of the projection matrix to roundoff errors. The model is proven to be energetically consistent if a global restitution coefficient is assumed. In the case of nonidentical restitution coefficients, the set of energetically consistent restitution matrices is characterized by using a linear matrix inequality (LMI).

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

Similar content being viewed by others

References

  1. McClamroch, N.H., Wang, D.: Feedback stabilization and tracking in constrained robots. IEEE Trans. Autom. Control 33, 419–426 (1988)

    Article  MATH  Google Scholar 

  2. Garcia de Jalón, J., Bayo, E.: Kinematic and Dynamic Simulation of Multibody Systems: The Real-Time Challenge. Springer, New York (1994)

    Book  Google Scholar 

  3. Blajer, W., Schiehlen, W., Schirm, W.: A projective criterion to the coordinate partitioning method for multibody dynamics. Appl. Mech. 64, 86–98 (1994)

    MATH  Google Scholar 

  4. Aghili, F.: Control of redundant mechanical systems under equality and inequality constraints on both input and constraint forces. J. Comput. Nonlinear Dyn. 6(3), 031013 (2011)

    Article  Google Scholar 

  5. Aghili, F., Su, C.: Impact dynamics in robotic and mechatronic systems. In: 2017 International Conference on Advanced Mechatronic Systems (ICAMechS), pp. 163–167 (2017)

    Chapter  Google Scholar 

  6. Gottschlich, S.N., Kak, A.C.: A dynamic approach to high-precision parts mating. IEEE Trans. Syst. Man Cybern. 19(4), 797–810 (1989)

    Article  Google Scholar 

  7. Dupree, K., Liang, C.H., Hu, G., Dixon, W.E.: Adaptive Lyapunov-based control of a robot and mass–spring system undergoing an impact collision. IEEE Trans. Syst. Man Cybern., Part B, Cybern. 38(4), 1050–1061 (2008)

    Article  Google Scholar 

  8. Marhefka, D.W., Orin, D.E.: A compliant contact model with nonlinear damping for simulation of robotic systems. IEEE Trans. Syst. Man Cybern., Part A, Syst. Hum. 29(6), 566–572 (1999)

    Article  Google Scholar 

  9. Khulief, Y.A.: Modeling of impact in multibody systems: an overview. J. Comput. Nonlinear Dyn. 8(2), 021012 (2013)

    Article  Google Scholar 

  10. Brogliato, B.: Kinetic quasi-velocities in unilaterally constrained Lagrangian mechanics with impacts and friction. Multibody Syst. Dyn. 32(2), 175–216 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  11. Hunt, K.H., Crossley, F.R.E.: Coefficient of restitution interpreted as damping in vibroimpact. J. Appl. Mech. 42, 440–445 (1975)

    Article  Google Scholar 

  12. Goldsmith, W.: Impact: The Theory and Physical Behavior of Colliding Solids. Edward Arnold, London (1960)

    MATH  Google Scholar 

  13. Lankarani, H.M., Nikravesh, P.E.: A contact force model with hysteresis damping for impact analysis of multibody systems. J. Mech. Des. 112, 369–376 (1990)

    Article  Google Scholar 

  14. Hurmuzlu, Y., Chang, T.H.: Rigid body collisions of a special class of planar kinematic chains. IEEE Trans. Syst. Man Cybern. 22(5), 964–971 (1992)

    Article  MATH  Google Scholar 

  15. Mu, X., Wu, Q.: On impact dynamics and contact events for biped robots via impact effects. IEEE Trans. Syst. Man Cybern., Part B, Cybern. 36(6), 1364–1372 (2006)

    Article  Google Scholar 

  16. Liu, C., Zhao, Z., Brogliato, B.: Frictionless multiple impacts in multibody systems. I. Theoretical framework. Proc. R. Soc. A 464, 3193–3211 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  17. Zhao, Z., Liu, C., Brogliato, B.: Planar dynamics of a rigid body system with frictional impacts. II. Qualitative analysis and numerical simulations. Proc. R. Soc. A 465, 2267–2292 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  18. Yoshida, Y., Takeuchi, K., Miyamoto, Y., Sato, D., Nenchev, D.: Postural balance strategies in response to disturbances in the frontal plane and their implementation with a humanoid robot. IEEE Trans. Syst. Man Cybern. Syst. 44(6), 692–704 (2014)

    Article  Google Scholar 

  19. Schiehlen, W., Seifried, R., Eberhard, P.: Elastoplastic phenomena in multibody impact dynamics. Comput. Methods Appl. Mech. Eng. 195, 6874–6890 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  20. Zhang, X., Vu-Quoc, L.: Modeling the dependence of the coefficient of restitution on the impact velocity in elasto-plastic collisions. Int. J. Impact Eng. 27(3), 317–341 (2002)

    Article  Google Scholar 

  21. Najafabadi, S.M., Kovecses, J., Angeles, J.: Impacts in multibody systems: modeling and experiments. Multibody Syst. Dyn. 120, 163–176 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  22. Pena, F., Lourenco, P.B., Campos-Costa, A.: Experimental dynamic behavior of free-standing multi-block structures under seismic loadings. J. Earthq. Eng. 12(6), 953–979 (2008)

    Article  Google Scholar 

  23. Aghili, F., Buehler, M., Hollerbach, J.M.: Dynamics and control of direct-drive robots with positive joint torque feedback. In: IEEE Int. Conf. Robotics and Automation, vol. 2, pp. 1156–1161 (1997)

    Chapter  Google Scholar 

  24. Arponen, T.: Regularization of constraint singularities in multibody systems. Multibody Syst. Dyn. 6, 355–375 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  25. Blajer, W.: Augmented lagrangian formulation: geometrical interpretation and application to systems with singularities and redundancy. Multibody Syst. Dyn. 8, 141–159 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  26. Aghili, F., Piedbœuf, J.-C.: Simulation of motion of constrained multibody systems based on projection operator. Multibody Syst. Dyn. 10, 3–16 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  27. Muller, A.: A conservative elimination procedure for permanently redundant closure constraints in MBS models with relative coordinates. Multibody Syst. Dyn. 16, 309–330 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  28. Aghili, F.: A unified approach for inverse and direct dynamics of constrained multibody systems based on linear projection operator: applications to control and simulation. IEEE Trans. Robot. 21(5), 834–849 (2005)

    Article  Google Scholar 

  29. Mistry, M., Buchli, J., Schaal, S.: Inverse dynamics control of floating base systems using orthogonal decomposition. In: 2010 IEEE International Conference on Robotics and Automation (ICRA), pp. 3406–3412 (2010)

    Chapter  Google Scholar 

  30. Mistry, M., Righetti, L.: Operational space control of constrained and underactuated systems. In: Proceedings of Robotics: Science and Systems, Los Angeles, CA, USA (2011)

    Google Scholar 

  31. Righetti, L., Buchli, J., Mistry, M., Schaal, S.: Inverse dynamics control of floating-base robots with external constraints: a unified view. In: 2011 IEEE International Conference on Robotics and Automation (ICRA), pp. 1085–1090 (2011)

    Chapter  Google Scholar 

  32. Ahmad, M., Ismail, K.A., Mat, F.: Impact models and coefficient of restitution: a review. J. Eng. Appl. Sci. 11(10), 6549–6555 (2016)

    Google Scholar 

  33. Ismail, K.A., Stronge, W.: Impact of viscoplastic bodies: dissipation and restitution. J. Appl. Mech. 75(6), 061011 (2008)

    Article  Google Scholar 

  34. Kangur, K., Kleis, I.: Experimental and theoretical determination of the coefficient of velocity restitution upon impact. Mech. Solids 23(5), 2–5 (1988)

    Google Scholar 

  35. Jackson, R.L., Green, I., Marghitu, D.B.: Predicting the coefficient of restitution of impacting elastic-perfectly plastic spheres. Nonlinear Dyn. 60(3), 217–229 (2010)

    Article  MATH  Google Scholar 

  36. Stronge, W.: Unraveling paradoxical theories for rigid body collisions. J. Appl. Mech. 58, 1049–1055 (1991)

    Article  MATH  Google Scholar 

  37. Stoianovici, D., Hurmuzlu, Y.: A critical study of the applicability of rigid-body collision theory. J. Appl. Mech. 63(2), 307–316 (1996)

    Article  Google Scholar 

  38. Lu, C.J., Kuo, M.C.: Coefficients of restitution based on a fractal surface model. J. Appl. Mech. 70(3), 339–345 (2003)

    Article  MATH  Google Scholar 

  39. amd, J.G.V., Braatz, R.D.: A tutorial on linear and bilinear matrix inequalities. J. Process Control 10, 363–385 (2000)

    Article  Google Scholar 

  40. Gahinet, P., Nemirovski, A., Laub, A.J., Chilali, M.: The LMI control toolbox. In: Proceedings of the 3rd European Control Conference, Rome, Italy, 1995, pp. 3206–3211 (1995)

    Google Scholar 

  41. Delebecque, R.N.F., Ghaoui, L.E.: LMITOOL: A Package for LMI Optimization in Scilab—User’s Guide (1995)

    Google Scholar 

  42. Golub, G.H., Loan, C.F.V.: Matrix Computations. Johns Hopkins University Press, Baltimore (1996)

    MATH  Google Scholar 

  43. Aghili, F.: Non-minimal order model of mechanical systems with redundant constraints for simulations and controls. IEEE Trans. Autom. Control 61(5), 1350–1355 (2016)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Farhad Aghili.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Time-derivative of a projection matrix

The Tikhonov regularization theorem [42] describes the pseudo-inverse as the following limit:

$$ \boldsymbol{A}^{+} = \lim_{\epsilon \rightarrow 0} \boldsymbol{A}^{T}\bigl(\boldsymbol{A} \boldsymbol{A} ^{T} + \epsilon \boldsymbol{I}\bigr)^{-1} . $$
(57)

By differentiation of the above expression, one can verify that the time-derivative of the pseudo-inverse can be written in the following form:

$$\begin{aligned} \frac{d}{dt} \boldsymbol{A}^{+} & =\lim _{\epsilon \rightarrow 0} \dot{\boldsymbol{A}} ^{T}\bigl(\boldsymbol{A} \boldsymbol{A}^{T} + \epsilon \boldsymbol{I}\bigr)^{-1} - \boldsymbol{A}^{T} \bigl(\boldsymbol{A} \boldsymbol{A}^{T} + \epsilon \boldsymbol{I}\bigr)^{-1} \bigl[\dot{\boldsymbol{A}} \boldsymbol{A}^{T} + \boldsymbol{A} \dot{\boldsymbol{A}}^{T} \bigr] \bigl(\boldsymbol{A} \boldsymbol{A}^{T} + \epsilon \boldsymbol{I} \bigr)^{-1} \\ & = - \boldsymbol{A}^{+} \dot{\boldsymbol{A}} \boldsymbol{A}^{+} + \lim_{\epsilon \rightarrow 0} \dot{ \boldsymbol{A}}^{T}\bigl(\boldsymbol{A} \boldsymbol{A}^{T} + \epsilon \boldsymbol{I}\bigr)^{-1} - \boldsymbol{A}^{+} \boldsymbol{A} \dot{\boldsymbol{A}}^{T}\bigl(\boldsymbol{A} \boldsymbol{A}^{T} + \epsilon \boldsymbol{I}\bigr)^{-1} \\ &= -\boldsymbol{A}^{+} \dot{\boldsymbol{A}} \boldsymbol{A}^{+} + \lim_{\epsilon \rightarrow 0} \boldsymbol{P} \dot{ \boldsymbol{A}}^{T} \bigl(\boldsymbol{A} \boldsymbol{A}^{T} + \epsilon \boldsymbol{I}\bigr)^{-1}. \end{aligned}$$
(58)

On the other hand, using (58) in the time-derivative of the expression of the projection matrix \(\boldsymbol{P}= \boldsymbol{I} - \boldsymbol{A}^{+} \boldsymbol{A}\) yields

$$\begin{aligned} \dot{\boldsymbol{P}} =& - \frac{\mathrm{d}}{\mathrm{d}t} \boldsymbol{A}^{+} \boldsymbol{A} - \boldsymbol{A}^{+} \dot{ \boldsymbol{A}} \\ =& \boldsymbol{A}^{+} \dot{\boldsymbol{A}} \boldsymbol{A}^{+} \boldsymbol{A} + \lim_{\epsilon \rightarrow 0} \boldsymbol{P} \dot{ \boldsymbol{A}}^{T} \bigl(\boldsymbol{A} \boldsymbol{A} ^{T} + \epsilon \boldsymbol{I}\bigr)^{-1} \boldsymbol{A} - \boldsymbol{A}^{+} \dot{\boldsymbol{A}} \\ =& \boldsymbol{A}^{+} \dot{\boldsymbol{A}} (\boldsymbol{I} - \boldsymbol{P}) + \boldsymbol{P} \dot{\boldsymbol{A}} ^{T} \boldsymbol{A}^{+T} - \boldsymbol{A}^{+} \dot{\boldsymbol{A}} \\ =& \boldsymbol{\varLambda }+ \boldsymbol{\varLambda }^{T}, \end{aligned}$$

where \(\boldsymbol{\varLambda }=- \boldsymbol{A}^{+} \dot{\boldsymbol{A}} \boldsymbol{P}\). Note that identity \(\boldsymbol{P} \boldsymbol{A}^{+} = \boldsymbol{A}^{+T} \boldsymbol{P} = \boldsymbol{0}\) implies \(\boldsymbol{\varLambda }^{T} \boldsymbol{P} = \boldsymbol{0}\) and hence one can conclude \(\ddot{\boldsymbol{q}}_{\perp }= \dot{\boldsymbol{P}} \dot{\boldsymbol{q}}= \boldsymbol{\varLambda } \dot{\boldsymbol{q}} = \boldsymbol{\varOmega }\dot{\boldsymbol{q}}\) [43].

Appendix B: Properties of \(\boldsymbol{M}_{c}\)

Consider a nonzero vector \(\boldsymbol{a} \in \mathbb{R}^{n}\) and its orthogonal decomposition components \(\boldsymbol{a}_{\parallel }=\boldsymbol{P} \boldsymbol{a}\) and \(\boldsymbol{a}_{\perp }=(\boldsymbol{I} - \boldsymbol{P}) \boldsymbol{a}\). Then, one can say that

$$ \boldsymbol{a}^{T} {\boldsymbol{M}}_{c} \boldsymbol{a} = \boldsymbol{a}^{T}_{\parallel } \boldsymbol{M} \boldsymbol{a}_{\parallel } + \nu \| \boldsymbol{a}_{\perp } \|^{2} >0. $$
(59)

Notice that both terms \(\boldsymbol{a}^{T}_{\parallel } \boldsymbol{M} \boldsymbol{a}_{\parallel }>0\) and \(\| \boldsymbol{a}_{\perp } \|^{2}>0\) are positive semi-definite. Moreover, for a given nonzero vector \(\boldsymbol{a}\), if \(\boldsymbol{a}_{\perp }= \boldsymbol{0}\) then \(\boldsymbol{a}^{T}_{\parallel } \neq 0\), and vice versa. This means that the sum of the two orthogonal terms must be positive definite, and so must be the constraint inertia matrix \({\boldsymbol{M}}_{c}\).

By definition we have

$$\begin{aligned} \boldsymbol{\varOmega }\dot{\boldsymbol{q}} = (\boldsymbol{I} - \boldsymbol{P}) \boldsymbol{\varOmega } \dot{\boldsymbol{q}.} \end{aligned}$$

Since \(\boldsymbol{M}_{c}\) is always invertible, i.e., \(\boldsymbol{M}_{c}^{-1}\boldsymbol{M} _{c} = \boldsymbol{I}\), the above equation can be equivalently written as

$$\begin{aligned} \boldsymbol{\varOmega }\dot{\boldsymbol{q}} & = \boldsymbol{M}_{c}^{-1} \boldsymbol{M}_{c} (\boldsymbol{I} - \boldsymbol{P}) \boldsymbol{ \varOmega }\dot{\boldsymbol{q}} \\ & = \boldsymbol{M}_{c}^{-1} \nu (\boldsymbol{I} - \boldsymbol{P}) \boldsymbol{\varOmega }\dot{\boldsymbol{q}} \\ & = \nu \boldsymbol{M}_{c}^{-1}\boldsymbol{\varOmega }\dot{ \boldsymbol{q},} \end{aligned}$$

which proves (9e).

From definition we have \(\boldsymbol{M}_{c} \boldsymbol{P} = \boldsymbol{M}\), and hence \(\boldsymbol{M} \boldsymbol{M}_{c}^{-1} = \boldsymbol{M}_{c} \boldsymbol{P} \boldsymbol{M}_{c}^{-1} = \boldsymbol{M} _{c} \boldsymbol{M}_{c}^{-1} \boldsymbol{P} = \boldsymbol{P}\), which proves relationship (9d).

Now, consider the characteristic equation of the constraint mass matrix

$$\begin{aligned} \bigl( \boldsymbol{P} \boldsymbol{M} \boldsymbol{P} + \nu (\boldsymbol{I} - \boldsymbol{P}) \bigr) \boldsymbol{x} - \lambda \boldsymbol{x} =0. \end{aligned}$$

Clearly, \(\lambda =\nu \) is the eigenvalue for all orthogonal eigenvectors which span \(\mathcal{N}^{\perp }(\boldsymbol{A})\) because \(\lambda =\nu \) means \((\boldsymbol{P} \boldsymbol{M} \boldsymbol{P} - \boldsymbol{P})\boldsymbol{x} =\boldsymbol{0} \quad \forall \boldsymbol{x} \in {\mathcal{N}}^{\perp }(\boldsymbol{A})\). The remaining set of orthogonal eigenvectors must lie in \(\mathcal{N}(\boldsymbol{A})\) that correspond to the nonzero eigenvalues of \(\boldsymbol{P} \boldsymbol{M} \boldsymbol{P}\):

$$\begin{aligned} \boldsymbol{P} \boldsymbol{M} \boldsymbol{P} \boldsymbol{x} - \lambda \boldsymbol{x} =\boldsymbol{0} \qquad \lambda \neq 0 \quad \forall \boldsymbol{x} \in {\mathcal{N}}. \end{aligned}$$

Therefore, the set of all eigenvalues of the positive definite matrix \({\boldsymbol{M}}_{c}\) is the union of the above sets corresponding to the eigenvectors in \(\mathcal{N}\) and \(\mathcal{N}^{\perp }\), i.e.,

$$ \lambda ({\boldsymbol{M}}_{c})=: \bigl\{ \underbrace{ \nu , \dots , \nu }_{r}, \; \underbrace{ \lambda _{\stackrel{\mathrm{min}}{\neq 0}} ( \boldsymbol{P} \boldsymbol{M} \boldsymbol{P}), \dots , \lambda _{\mathrm{max}}( \boldsymbol{P} \boldsymbol{M} \boldsymbol{P})} _{n-r} \bigr\} , $$
(60)

where \(\{ \lambda _{\stackrel{\mathrm{min}}{\neq 0}}(\boldsymbol{P} \boldsymbol{M} \boldsymbol{P}), \dots , \lambda _{\mathrm{max}}(\boldsymbol{P} \boldsymbol{M} \boldsymbol{P}) \}\) are all nonzero eigenvalues of \(\boldsymbol{P} \boldsymbol{M} \boldsymbol{P}\). According to (60), the condition number of \(\boldsymbol{M}_{c}\), which is simply the ratio of the largest to smallest eigenvalues, is

$$ \mbox{cond}({\boldsymbol{M}}_{c}) = \frac{\max (\nu , \lambda _{\mathrm{max}}( \boldsymbol{P} \boldsymbol{M} \boldsymbol{P}) )}{\min (\nu , \lambda _{\stackrel{\mathrm{min}}{\neq 0} }(\boldsymbol{P} \boldsymbol{M} \boldsymbol{P}))} . $$
(61)

Clearly, the RHS of (61) is at its minimum if \(\nu \) is selected to be within the lower- and upper-bounds defined in (9f).

By virtue of (60), the Singular Value Decomposition of \(\boldsymbol{M}_{c}\) takes the form

$$ \boldsymbol{M}_{c} = \begin{bmatrix} \boldsymbol{V}_{1} & \boldsymbol{V}_{2} \end{bmatrix} \begin{bmatrix} \nu \boldsymbol{I} & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{\varSigma } \end{bmatrix} \begin{bmatrix} \boldsymbol{V}_{1}^{T} \\ \boldsymbol{V}_{2}^{T} \end{bmatrix} , $$
(62)

where matrix \(\boldsymbol{\varSigma }=\mbox{diag}\{ \lambda _{\stackrel{\mathrm{min}}{\neq 0}}(\boldsymbol{P} \boldsymbol{M} \boldsymbol{P}), \dots ,\lambda _{\mathrm{max}}(\boldsymbol{P} \boldsymbol{M} \boldsymbol{P})\}\) contains the nonzero singular values, \(\boldsymbol{V}=[\boldsymbol{V}_{1} \;\; \boldsymbol{V}_{2}]\) is a unitary matrix so that \(\mbox{span}(\boldsymbol{V}_{1}) \equiv {\mathcal{N}} ^{\perp }\) and \(\mbox{span}(\boldsymbol{V}_{2}) \equiv {\mathcal{N}}\), i.e., \(\boldsymbol{P} = \boldsymbol{V}_{2} \boldsymbol{V}_{2}^{T}\), \(\boldsymbol{V}_{2}^{T} \boldsymbol{V}_{2} = \boldsymbol{I}\), and \(\boldsymbol{V}_{1}^{T} \boldsymbol{V}_{2} =\boldsymbol{0}\). Thus

$$\begin{aligned} \boldsymbol{M}_{c}^{-1} \boldsymbol{P} &= \begin{bmatrix} \boldsymbol{V}_{1} & \boldsymbol{V}_{2} \end{bmatrix} \begin{bmatrix} \nu ^{-1} \boldsymbol{I} & \boldsymbol{0} \\ \boldsymbol{0} & \boldsymbol{\varSigma }^{-1} \end{bmatrix} \begin{bmatrix} \boldsymbol{V}_{1}^{T} \\ \boldsymbol{V}_{2}^{T} \end{bmatrix} \boldsymbol{V}_{2} \boldsymbol{V}_{2}^{T} \\ &= \boldsymbol{V}_{2} \boldsymbol{\varSigma }^{-1} \boldsymbol{V}_{2}^{T} = \boldsymbol{M}_{o}^{+}. \end{aligned}$$

Appendix C: Properties of \(\boldsymbol{S}\)

Using (9b) in the following derivations yields

$$\begin{aligned} \boldsymbol{S} \boldsymbol{P} &= \boldsymbol{P} - \boldsymbol{M}_{c}^{-1} \boldsymbol{P} \boldsymbol{M} \boldsymbol{P} \\ &= \boldsymbol{P} - \boldsymbol{M}_{c}^{-1} \boldsymbol{M}_{c} \boldsymbol{P} \\ & = \boldsymbol{P} - \boldsymbol{P} = \boldsymbol{0}, \end{aligned}$$

which proves identity (11a). It follows that

$$\begin{aligned} (\boldsymbol{I} - \boldsymbol{P}) \boldsymbol{S}^{T} = \boldsymbol{S}^{T} - \boldsymbol{P} \boldsymbol{S}^{T} = \boldsymbol{S}^{T}. \end{aligned}$$

On the other hand, by definition we have

$$\begin{aligned} (\boldsymbol{I} - \boldsymbol{P}) \boldsymbol{S} &= \boldsymbol{I} - \boldsymbol{P} - \boldsymbol{M}_{c}^{-1} \boldsymbol{P} \boldsymbol{M} + \boldsymbol{P} \boldsymbol{M}_{c}^{-1} \boldsymbol{P} \boldsymbol{M} \\ &= \boldsymbol{I} - \boldsymbol{P} - \boldsymbol{M}_{c}^{-1} \boldsymbol{P} \boldsymbol{M} + \boldsymbol{M}_{c}^{-1} \boldsymbol{P}^{2} \boldsymbol{M} \\ & = \boldsymbol{I} - \boldsymbol{P}, \end{aligned}$$

which proves identity (11c).

Finally,

$$\begin{aligned} \boldsymbol{M} \boldsymbol{S} &= \boldsymbol{M} - \boldsymbol{M} \boldsymbol{M}_{c}^{-1} \boldsymbol{P} \boldsymbol{M} \\ &= \boldsymbol{M} - \boldsymbol{M} \boldsymbol{P} \boldsymbol{M}_{c}^{-1} \boldsymbol{M} \\ &= \bigl(\boldsymbol{I} - \boldsymbol{M} \boldsymbol{P} \boldsymbol{M}_{c}^{-1} \bigr) \boldsymbol{M} \\ & = \boldsymbol{S}^{T} \boldsymbol{M}. \end{aligned}$$

On the other hand, using the above result in the following derivation yields

$$ \boldsymbol{S}^{T} \boldsymbol{M} \boldsymbol{S} = \boldsymbol{S}^{T} \boldsymbol{S}^{T} \boldsymbol{M} = \boldsymbol{S}^{T} \boldsymbol{M}, $$

which proves identity (11d).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aghili, F. Modeling and analysis of multiple impacts in multibody systems under unilateral and bilateral constrains based on linear projection operators. Multibody Syst Dyn 46, 41–62 (2019). https://doi.org/10.1007/s11044-018-09658-w

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11044-018-09658-w

Keywords

Navigation