Skip to main content
Log in

Elimination approach toward normalization constraint for Euler parameters

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

Abstract

It is known that three-parameter representations of spatial rotation suffer from kinematic singularities. Euler parameters (unit quaternions) are a four-parameter solution for this issue. However, a normality constraint needs to be enforced while implementing this set of four parameters, otherwise the mass matrix in the equations of motion will be singular. This paper discusses an approach to enforce this normality constraint using a redefinition of the state space in terms of quasi-velocities, along with the standard elimination of dependent variables at the velocity and acceleration level. The approach is implemented on a 3D double pendulum example.

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

Similar content being viewed by others

Availability of data

All data generated from the simulations are included within this article.

References

  1. Andrle, M.S., Crassidis, J.L.: Geometric integration of quaternions. J. Guidance Control Dyn. 36(6), 1762–1767 (2013). https://doi.org/10.2514/1.58558

    Article  Google Scholar 

  2. Betsch, P.: On the use of Euler parameters in multibody dynamics. PAMM 6(1), 85–86 (2006). https://doi.org/10.1002/pamm.200610023

    Article  Google Scholar 

  3. Betsch, P., Siebert, R.: Rigid body dynamics in terms of quaternions: Hamiltonian formulation and conserving numerical integration. Int. J. Numer. Methods Eng. 79(4), 444–473 (2009). https://doi.org/10.1002/nme.2586

    Article  MathSciNet  MATH  Google Scholar 

  4. Bowling, A.: Vector Mechanics: A Systematic Approach. Aqualan Press, LLC, New York (2018)

    Google Scholar 

  5. Chou, J.C.K.: Quaternion kinematic and dynamic differential equations. IEEE Trans. Robot. Autom. 8(1), 53–64 (1992). https://doi.org/10.1109/70.127239

    Article  MathSciNet  Google Scholar 

  6. Crouch, P.E., Grossman, R.: Numerical integration of ordinary differential equations on manifolds. J. Nonlinear Sci. 3(1), 1–33 (1993). https://doi.org/10.1007/BF02429858

    Article  MathSciNet  MATH  Google Scholar 

  7. Diebel, J.: Representing attitude: Euler angles, unit quaternions, and rotation vectors. https://www.astro.rug.nl/software/kapteyn-beta/_downloads/attitude.pdf (2006). Accessed 10 Jan 2022

  8. Haghshenas-Jaryani, M., Bowling, A.: A new switching strategy for addressing Euler parameters in dynamic modeling and simulation of rigid multibody systems. Multibody Syst. Dyn. 30(2), 185–197 (2013)

    Article  MathSciNet  Google Scholar 

  9. Hurtado, J.E., Sinclair, A.J.: Lagrangian mechanics of overparameterized systems. Nonlinear Dyn. 66(1), 201–212 (2011). https://doi.org/10.1007/s11071-010-9921-1

    Article  MathSciNet  MATH  Google Scholar 

  10. Kane, T.R., Levinson, D.A.: Dynamics: Theory and Applications. McGraw Hill, New York (1985)

    Google Scholar 

  11. Möller, M., Glocker, C.: Rigid body dynamics with a scalable body, quaternions and perfect constraints. Multibody Syst. Dyn. 27(4), 437–454 (2012)

    Article  MathSciNet  Google Scholar 

  12. Munthe-Kaas, H.: Runge–Kutta methods on Lie groups. BIT Numer. Math. 38(1), 92–111 (1998). https://doi.org/10.1007/BF02510919

    Article  MathSciNet  MATH  Google Scholar 

  13. Nikravesh, P.E., Wehage, R.A., Kwon, O.K.: Euler parameters in computational kinematics and dynamics. Part 1. J. Mech. Transm. Autom. Des. 107(3), 358–365 (1985)

    Article  Google Scholar 

  14. Nikravesh, P.E., Wehage, R.A., Kwon, O.K.: Euler parameters in computational kinematics and dynamics. Part 2. J. Mech. Transm. Autom. Des. 107(3), 366–369 (1985)

    Article  Google Scholar 

  15. O’Reilly, O.M., Varadi, P.C.: Hoberman’s sphere, Euler parameters and Lagrange equations. J. Elast. 56(2), 171–180 (1999). https://doi.org/10.1023/A:1007624027030

  16. Senan, N.A.F., O’Reilly, O.M.: On the use of quaternions and Euler–Rodrigues symmetric parameters with moments and moment potentials. Int. J. Eng. Sci. 47(4), 595–609 (2009). https://doi.org/10.1016/j.ijengsci.2008.12.008

  17. Shabana, A.A.: Euler parameters kinetic singularity. Proc. Inst. Mech. Eng. Part K J. Multi-body Dyn. 228(3), 307–313 (2014)

  18. Sherif, K., Nachbagauer, K., Steiner, W.: On the rotational equations of motion in rigid body dynamics when using Euler parameters. Nonlinear Dyn. 81(1), 343–352 (2015). https://doi.org/10.1007/s11071-015-1995-3

    Article  MathSciNet  MATH  Google Scholar 

  19. Shivarama, R., Fahrenthold, E.P.: Hamilton’s equations with Euler parameters for rigid body dynamics modeling. J. Dyn. Syst. Meas. Control 126(1), 124–130 (2004). https://doi.org/10.1115/1.1649977

  20. Sveier, A., Sjøberg, A.M., Egeland, O.: Applied Runge–Kutta–Munthe–Kaas integration for the quaternion kinematics. J. Guidance Control Dyn. 42(12), 2747–2754 (2019). https://doi.org/10.2514/1.G004578

    Article  Google Scholar 

  21. Udwadia, F.E., Schutte, A.D.: An alternative derivation of the quaternion equations of motion for rigid-body rotational dynamics. J. Appl. Mech. (2010). https://doi.org/10.1115/1.4000917

    Article  Google Scholar 

  22. Udwadia, F.E., Schutte, A.D.: A unified approach to rigid body rotational dynamics and control. Proc. R. Soc. A Math. Phys. Eng. Sci. 468(2138), 395–414 (2012). https://doi.org/10.1098/rspa.2011.0233

    Article  MathSciNet  MATH  Google Scholar 

  23. Vadali, S.: On the Euler parameter constraint. Aerospace Sciences Meetings. American Institute of Aeronautics and Astronautics (1988). https://doi.org/10.2514/6.1988-670

  24. Wehage, R.A., Haug, E.J.: Generalized coordinate partitioning for dimension reduction in analysis of constrained dynamic systems. J. Mech. Des. 104(1), 247–255 (1982)

    Google Scholar 

  25. Yu, W., Pan, Z.: Dynamical equations of multibody systems on Lie groups. Adv. Mech. Eng. 7(3), 1687814015575959 (2015). https://doi.org/10.1177/1687814015575959

  26. Zorić, N.D., Lazarević, M.P., Simonović, A.M.: Multi-body kinematics and dynamics in terms of quaternions: Langrange formulation in covariant form: Rodriguez approach. FME Trans. 38(1), 19–28 (2010)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alan Bowling.

Ethics declarations

Conflict of interest

The authors have no conflicts of interest to declare that are relevant to the content of this article.

Additional information

Publisher's Note

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

Appendices

Appendix 1: Double-pendulum equations of motion

The equations of motion for the double pendulum example are computed using Kane’s method. Because the final EOMs are too large, this section does not derive complete EOMs. Rather, it provides important equations in a format discussed in [4], which when evaluated result in the EOMs for the double pendulum.

Rigid Bodies: There are two rigid bodies, the two cylindrical rods of double pendulum.

Inertial Reference Frame and Point: The inertial reference point is N, and N \(= \left( {\widehat{\mathbf {N}}_{1}},{\widehat{\mathbf {N}}_{2}},{\widehat{\mathbf {N}}_{3}} \right) \) is the inertial reference frame shown in Fig. 2.

Other Points and Frames: Points A and B are body-attached points, and they are also the center of gravity of bodies A and B, respectively. Point C represents the spherical joint between two rods. A \(= \left( {\widehat{\mathbf {A}}_{1}},{\widehat{\mathbf {A}}_{2}},{\widehat{\mathbf {A}}_{3}} \right) \) and B \(= \left( {\widehat{\mathbf {B}}_{1}},{\widehat{\mathbf {B}}_{2}},{\widehat{\mathbf {B}}_{3}} \right) \) are body-attached frames shown in Fig. 2 (Tables 2, 3 and 4).

Location Descriptions: The location descriptions provide a formal way of specifying the position of every point that comprises the rigid bodies in the system.

$$\begin{aligned} L_A = \left\{ {\mathbf {P}_{NA}}, {}^{}_{A}{\,}^{N}_{}R \right\} + geom. L_B = \left\{ {\mathbf {P}_{CB}}, {}^{}_{B}{\,}^{A}_{}R \right\} + geom. \end{aligned}$$
(23)
$$\begin{aligned} {\mathbf {P}_{NA}} = {\mathbf {P}_{AC}} = l_A {\widehat{\mathbf {A}}_{1}} {\mathbf {P}_{NC}} = 2l_A {\widehat{\mathbf {A}}_{1}} {\mathbf {P}_{CB}} = l_B {\widehat{\mathbf {B}}_{1}} \end{aligned}$$
(24)
$$\begin{aligned} {}^{}_{A}{\,}^{N}_{}R = \frac{1}{e_{A_0}^2 + e_{A_1}^2 + e_{A_2}^2 + e_{A_3}^2} \begin{bmatrix} e_{A_0}^2 + e_{A_1}^2 - e_{A_2}^2 - e_{A_3}^2 &{} 2e_{A_1}e_{A_2} - 2e_{A_0}e_{A_3} &{} 2e_{A_0}e_{A_2} + 2e_{A_1}e_{A_3} \\ 2e_{A_1}e_{A_2} + 2e_{A_0}e_{A_3} &{} e_{A_0}^2 - e_{A_1}^2 + e_{A_2}^2 - e_{A_3}^2 &{} 2e_{A_2}e_{A_3} - 2e_{A_0}e_{A_1} \\ 2e_{A_1}e_{A_3} - 2e_{A_0}e_{A_2} &{} 2e_{A_2}e_{A_3} + 2e_{A_0}e_{A_1} &{} e_{A_0}^2 - e_{A_1}^2 - e_{A_2}^2 + e_{A_3}^2 \end{bmatrix} \end{aligned}$$
(25)
$$\begin{aligned} {}^{}_{B}{\,}^{A}_{}R = \frac{1}{e_{B_0}^2 + e_{B_1}^2 + e_{B_2}^2 + e_{B_3}^2} \begin{bmatrix} e_{B_0}^2 + e_{B_1}^2 - e_{B_2}^2 - e_{B_3}^2 &{} 2e_{B_1}e_{B_2} - 2e_{B_0}e_{B_3} &{} 2e_{B_0}e_{B_2} + 2e_{B_1}e_{B_3} \\ 2e_{B_1}e_{B_2} + 2e_{B_0}e_{B_3} &{} e_{B_0}^2 - e_{B_1}^2 + e_{B_2}^2 - e_{B_3}^2 &{} 2e_{B_2}e_{B_3} - 2e_{B_0}e_{B_1} \\ 2e_{B_1}e_{B_3} - 2e_{B_0}e_{B_2} &{} 2e_{B_2}e_{B_3} + 2e_{B_0}e_{B_1} &{} e_{B_0}^2 - e_{B_1}^2 - e_{B_2}^2 + e_{B_3}^2 \end{bmatrix} \end{aligned}$$
(26)

Coordinates: Eight coordinates appear in the location descriptions, where \(e_{A_{0-3}}\) and \(e_{B_{0-3}}\) are Euler parameters representing body A’s and body B’s orientation, respectively. The vector of generalized coordinates, \(\mathbf{q}\), is defined as:

$$\begin{aligned} \mathbf {q} = [ \mathbf {e}_A^T \ \ \mathbf {e}_B^T ]^T = [ e_{A_0} \ e_{A_1} \ e_{A_2} \ e_{A_3} \ e_{B_0} \ e_{B_1} \ e_{B_2} \ e_{B_3} ]^T \end{aligned}$$

Constraints: The normalization constraints associated with both the Euler parameter sets are given as

$$\begin{aligned} q_1^2 + q_2^2 + q_3^2 + q_4^2&= 1&,&q_5^2 + q_6^2 + q_7^2 + q_8^2&= 1 \end{aligned}$$
Table 2 Euler parameters for approach in Ref.  [8]
Table 3 Euler parameters for the proposed approach
Table 4 Euler parameters when the rotation matrices are integrated directly

Degrees of Freedom (DOFs):

8 coordinates - 2 constraints = 6 DOFs

Velocity: The angular velocity of body A and body B is

$$\begin{aligned} {^\mathrm{N} \varvec{\omega }^\mathrm{A}}&= 2 \left( u_1 {\widehat{\mathbf {A}}_{1}} + u_2 {\widehat{\mathbf {A}}_{2}} + u_3 {\widehat{\mathbf {A}}_{3}} \right) \nonumber \\&= 2( q_1 \dot{q}_2 - q_2 \dot{q}_1 - q_3 \dot{q}_4 + q_4 \dot{q}_3 ) {\widehat{\mathbf {A}}_{1}} \nonumber \\&\quad + 2( q_1 \dot{q}_3 + q_2 \dot{q}_4 - q_3 \dot{q}_1 - q_4 \dot{q}_2 ) {\widehat{\mathbf {A}}_{2}} \nonumber \\&\quad + 2( q_1 \dot{q}_4 - q_2 \dot{q}_3 + q_3 \dot{q}_2 - q_4 \dot{q}_1 ) {\widehat{\mathbf {A}}_{3}} \end{aligned}$$
(27)
$$\begin{aligned} {^\mathrm{A} \varvec{\omega }^\mathrm{B}}&= 2 \left( u_5 {\widehat{\mathbf {B}}_{1}} + u_6 {\widehat{\mathbf {B}}_{2}} + u_7 {\widehat{\mathbf {B}}_{3}} \right) \nonumber \\&= 2( q_5 \dot{q}_6 - q_6 \dot{q}_5 - q_7 \dot{q}_8 + q_8 \dot{q}_7 ) {\widehat{\mathbf {B}}_{1}} \nonumber \\&\quad + 2( q_5 \dot{q}_7 + q_6 \dot{q}_8 - q_7 \dot{q}_5 - q_8 \dot{q}_6 ) {\widehat{\mathbf {B}}_{2}} \nonumber \\&\quad + 2( q_5 \dot{q}_8 - q_6 \dot{q}_7 + q_7 \dot{q}_6 - q_8 \dot{q}_5 ) {\widehat{\mathbf {B}}_{3}} \end{aligned}$$
(28)
$$\begin{aligned} {^\mathrm{N} \varvec{\omega }^\mathrm{B}}&= {}^{}_{B}{\,}^{A}_{}R^T {^\mathrm{N} \varvec{\omega }^\mathrm{A}} + {^\mathrm{A} \varvec{\omega }^\mathrm{B}} \end{aligned}$$
(29)

The translational velocity of the mass centers A and B and point C is given as

$$\begin{aligned} {\mathbf {V}_{A}}&= \frac{d{\mathbf {P}_{NA}}}{dt} = {^\mathrm{N} \varvec{\omega }^\mathrm{A}} \times {\mathbf {P}_{NA}} \end{aligned}$$
(30)
$$\begin{aligned} {\mathbf {V}_{C}}&= \frac{d{\mathbf {P}_{NC}}}{dt} = {^\mathrm{N} \varvec{\omega }^\mathrm{A}} \times {\mathbf {P}_{NC}} \end{aligned}$$
(31)
$$\begin{aligned} {\mathbf {V}_{B}}&= \frac{d{\mathbf {P}_{NB}}}{dt} = {}^{}_{B}{\,}^{A}_{}R^T {\mathbf {V}_{C}} + {^\mathrm{N} \varvec{\omega }^\mathrm{B}} \times {\mathbf {P}_{CB}} \end{aligned}$$
(32)

The quasi-velocities required to compute partial derivatives in Kane’s equations of motion are \(\mathbf {u}_A\) and \(\mathbf {u}_B\) which are already defined in Sect. 3.

Acceleration: The angular acceleration of body A and body B is

$$\begin{aligned} {^\mathrm{N}\dot{\varvec{\omega }}^\mathrm{A}}&= 2( q_1 \ddot{q}_2 - q_2 \ddot{q}_1 - q_3 \ddot{q}_4 + q_4 \ddot{q}_3 ) {\widehat{\mathbf {A}}_{1}} \nonumber \\&\quad + 2( q_1 \ddot{q}_3 + q_2 \ddot{q}_4 - q_3 \ddot{q}_1 - q_4 \ddot{q}_2 ) {\widehat{\mathbf {A}}_{2}} \nonumber \\&\quad + 2( q_1 \ddot{q}_4 - q_2 \ddot{q}_3 + q_3 \ddot{q}_2 - q_4 \ddot{q}_1 ) {\widehat{\mathbf {A}}_{3}} \end{aligned}$$
(33)
$$\begin{aligned} {^\mathrm{A}\dot{\varvec{\omega }}^\mathrm{B}}&= 2( q_5 \ddot{q}_6 - q_6 \ddot{q}_5 - q_7 \ddot{q}_8 + q_8 \ddot{q}_7 ) {\widehat{\mathbf {B}}_{1}} \nonumber \\&\quad + 2( q_5 \ddot{q}_7 + q_6 \ddot{q}_8 - q_7 \ddot{q}_5 - q_8 \ddot{q}_6 ) {\widehat{\mathbf {B}}_{2}} \nonumber \\&\quad + 2( q_5 \ddot{q}_8 - q_6 \ddot{q}_7 + q_7 \ddot{q}_6 - q_8 \ddot{q}_5 ) {\widehat{\mathbf {B}}_{3}} \end{aligned}$$
(34)
$$\begin{aligned} {^\mathrm{N}\dot{\varvec{\omega }}^\mathrm{B}}&= {}^{}_{B}{\,}^{A}_{}R^T {^\mathrm{N}\dot{\varvec{\omega }}^\mathrm{A}} + {^\mathrm{A}\dot{\varvec{\omega }}^\mathrm{B}} + {^\mathrm{N} \varvec{\omega }^\mathrm{B}} \times {^\mathrm{A} \varvec{\omega }^\mathrm{B}} \end{aligned}$$
(35)

The translational acceleration of mass centers A and B and point C is given as

$$\begin{aligned} {\dot{\mathbf {V}}_{A}}&= \frac{d{\mathbf {V}_{A}}}{dt} = {^\mathrm{N}\dot{\varvec{\omega }}^\mathrm{A}} \times {\mathbf {P}_{NA}} + {^\mathrm{N} \varvec{\omega }^\mathrm{A}} \times \left( {^\mathrm{N} \varvec{\omega }^\mathrm{A}} \times {\mathbf {P}_{NA}} \right) \end{aligned}$$
(36)
$$\begin{aligned} {\dot{\mathbf {V}}_{C}}&= \frac{d{\mathbf {V}_{C}}}{dt} = {^\mathrm{N}\dot{\varvec{\omega }}^\mathrm{A}} \times {\mathbf {P}_{NC}} + {^\mathrm{N} \varvec{\omega }^\mathrm{A}} \times \left( {^\mathrm{N} \varvec{\omega }^\mathrm{A}} \times {\mathbf {P}_{NC}} \right) \end{aligned}$$
(37)
$$\begin{aligned} {\dot{\mathbf {V}}_{B}}&= \frac{d{\mathbf {V}_{B}}}{dt} = {}^{}_{B}{\,}^{A}_{}R^T {\dot{\mathbf {V}}_{C}} + {^\mathrm{N}\dot{\varvec{\omega }}^\mathrm{B}} \times {\mathbf {P}_{CB}} \nonumber \\&+ {^\mathrm{N} \varvec{\omega }^\mathrm{B}} \times \left( {^\mathrm{N} \varvec{\omega }^\mathrm{B}} \times {\mathbf {P}_{CB}} \right) \end{aligned}$$
(38)

Mass Properties: Mass of body A and body B is assumed to be \(m_A\) and \(m_B\), respectively, while the spherical joint is assumed to be massless. The inertia matrix of body A and body B can be given as

$$\begin{aligned} {I_\mathrm{AA}}&= \begin{bmatrix} \frac{1}{2}m_A r^2 &{} 0 &{} 0 \\ 0 &{} \frac{1}{12}m_A(3r^2+L^2) &{} 0 \\ 0 &{} 0 &{} \frac{1}{12}m_A(3r^2+L^2) \end{bmatrix} \end{aligned}$$
(39)
$$\begin{aligned} {I_\mathrm{BB}}&= \begin{bmatrix} \frac{1}{2}m_B r^2 &{} 0 &{} 0 \\ 0 &{} \frac{1}{12}m_B(3r^2+L^2) &{} 0 \\ 0 &{} 0 &{} \frac{1}{12}m_B(3r^2+L^2) \end{bmatrix} \end{aligned}$$
(40)

Forces and Moments: No external moments are acting on the bodies. Only the gravitational force acts on both the bodies. Note that both resultant force and moment vectors given below are expressed in inertial frame. Moments were taken about the mass center of each body.

$$\begin{aligned} {\mathbf {F}_{A}}&= -m_A g {\widehat{\mathbf {N}}_{3}},&{\mathbf {F}_{B}}&= -m_B g {\widehat{\mathbf {N}}_{3}} \end{aligned}$$
(41)
$$\begin{aligned} {\mathbf {M}_{AA}}&= \mathbf {0},&{\mathbf {M}_{BB}}&= \mathbf {0} \end{aligned}$$
(42)

Equations of Motion: The equations of motion can be computed using the Kane’s equations given below. All the necessary terms are already defined in this appendix.

$$\begin{aligned} \begin{aligned} 0&= F_i - F_i^*\\ F_i&= \sum _{K=1}^{bodies} \left[ {\mathbf {F}_{K}} \cdot \frac{\partial {\mathbf {V}_{K}}}{\partial u_i} + {\mathbf {M}_{KK}} \cdot \frac{\partial {^\mathrm{N} \varvec{\omega }^\mathrm{K}}}{\partial u_i} \right] \\ F_i^*&= \sum _{K=1}^{bodies} \left[ m_K {\dot{\mathbf {V}}_{K}} \cdot \frac{\partial {\mathbf {V}_{K}}}{\partial u_i} + {\dot{\mathbf {H}}_{KK}} \cdot \frac{\partial {^\mathrm{N} \varvec{\omega }^\mathrm{K}}}{\partial u_i} \right] \end{aligned} \end{aligned}$$
(43)

where \(i= \{ 1,2,3,5,6,7 \}\) and,

$$\begin{aligned} {\dot{\mathbf {H}}_{KK}} = {I_\mathrm{KK}} {^\mathrm{N}\dot{\varvec{\omega }}^\mathrm{K}} + {^\mathrm{N} \varvec{\omega }^\mathrm{K}} \times \left( {I_\mathrm{KK}} {^\mathrm{N} \varvec{\omega }^\mathrm{K}} \right) \end{aligned}$$
(44)

Appendix 2: Online constraint embedding method

The virtual work done by the system can be calculated from (43) as follows:

$$\begin{aligned} 0 = \delta W= & {} \sum _{i = 1}^n \sum _{K = 1}^{bodies} \\&\left[ {\mathbf {F}_{K}} \cdot \frac{\partial {\mathbf {V}_{K}}}{\partial u_i} - m_K {\dot{\mathbf {V}}_{K}} \cdot \frac{\partial {\mathbf {V}_{K}}}{\partial u_i} \right. \\&\left. + {\mathbf {M}_{KK}} \cdot \frac{\partial {^\mathrm{N} \varvec{\omega }^\mathrm{K}}}{\partial u_i} - {\dot{\mathbf {H}}_{KK}} \cdot \frac{\partial {^\mathrm{N} \varvec{\omega }^\mathrm{K}}}{\partial u_i} \right] \delta q_i \end{aligned}$$

The equation above can also be expressed as

$$\begin{aligned} 0 = \sum _{i = 1}^n (F_i - F_i^*) \delta q_i \end{aligned}$$
(45)

Here, the \(F_i\) and \(F_i^*\) are generated for both the dependent and independent generalized coordinates. Note that any holonomic constraint can be expressed in a form that is linear in the virtual displacements. Let’s say that we are solving for a system that has five generalized coordinates and two constraints. Thus, the relationship between dependent and independent virtual displacements can be given as

$$\begin{aligned} \begin{bmatrix} \delta q_{D_4} \\ \delta q_{D_5} \end{bmatrix} = \begin{bmatrix} C_{41} &{} C_{42} &{} C_{43} \\ C_{51} &{} C_{52} &{} C_{53} \end{bmatrix} \begin{bmatrix} \delta q_{I_1} \\ \delta q_{I_2} \\ \delta q_{I_3} \end{bmatrix} \end{aligned}$$
(46)

where the subscripts ‘\(D\)’ and ‘\(I\)’ denote a dependent or independent virtual displacement, respectively. Substituting the equation above in Eq. (45) yields

$$\begin{aligned} 0&= (F_1 - F_1^*)\delta q_1 \ + \ (F_2 - F_2^*)\delta q_2 \nonumber \\&\quad \, + \ (F_3 - F_3^*)\delta q_3 \quad \ + \ (F_4 - F_4^*)\delta q_4 \ \nonumber \\&\quad \,+ \ (F_5 - F_5^*)\delta q_5 \nonumber \\&= (F_1 - F_1^*)\delta q_1 \ + \ (F_2 - F_2^*)\delta q_2 \ \nonumber \\&\quad \ + \ (F_3 - F_3^*)\delta q_3 + \ (F_4 - F_4^*)(C_{41} \delta q_1 \nonumber \\&\quad \ + \ C_{42} \delta q_2 \ + \ C_{43} \delta q_3) \nonumber \\&\quad \ + \ (F_5 - F_5^*)(C_{51} \delta q_1 \ + \ C_{52} \delta q_2 \nonumber \\&\quad \ + \ C_{53} \delta q_3) \nonumber \\&= (F_1 - F_1^*\ + \ C_{41} (F_4 - F_4^*) \nonumber \\&\quad \ + \ C_{51} (F_5 - F_5^*))\delta q_1 \nonumber \\&\quad \ + \ (F_2 - F_2^*\ + \ C_{42} (F_4 - F_4^*) \nonumber \\&\quad \ + \ C_{52} (F_5 - F_5^*))\delta q_2 \nonumber \\&\quad \ + \ (F_3 - F_3^*\ + \ C_{43} (F_4 - F_4^*) \nonumber \\&\quad \ + \ C_{53} (F_5 - F_5^*))\delta q_3 \end{aligned}$$
(47)

Because the virtual displacements in the equation above are independent, the only way for the virtual work to be equal to zero for any values they may take is when their coefficients are zero for all time. Thus, it can be shown that

$$\begin{aligned} \bar{F}_i - \bar{F}_i^*&= 0 \quad \bar{F}_i = F_i + \sum _{j = p+1}^{p+q} C_{ji} F_j, \nonumber \\ \bar{F}_i^*&= F_i^*+ \sum _{j = p+1}^{p+q} C_{ji} F_j^*\end{aligned}$$
(48)

where \(p\) and \(q\) are the number of independent and dependent generalized coordinates, respectively. And, \(i = \{1,2,3,...,p\}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Joshi, V., Guy, A. & Bowling, A. Elimination approach toward normalization constraint for Euler parameters. Nonlinear Dyn 108, 1479–1493 (2022). https://doi.org/10.1007/s11071-022-07236-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11071-022-07236-6

Keywords

Navigation