Abstract

The paper presents a geometrical overview on an optimal control problem on a special Lie group. The Hamilton-Poisson realization of the dynamics offers us the possibility to study the system from mechanical geometry point of view.

1. Introduction

Recent work in nonlinear control has drawn attention to drift-free systems with fewer degrees than state variables. These arise naturally in problems of motion planning for wheeled robots subject to nonholonomic controls [1], models of kinematic drift effects in space subjects to appendage vibrations or articulations [1], the molecular dynamics [2], the autonomous underwater vehicle dynamics [3], the car's dynamics [4], and spacecraft dynamics [5]. The purpose of our paper is to study a class of left-invariant, drift-free optimal control problems on a specific Lie group 𝐺. The class of all control-affine left-invariant, drift-free optimal control problems on 𝐺 can be reduced to a class of two typical controllable left-invariant control systems on 𝐺. The left-invariant, drift-free optimal control problems involve finding a trajectory-control pair on 𝐺, which minimizes a cost function and satisfies the given dynamical constrains and boundary conditions in a fixed time. The problem is lifted to the cotangent bundle π‘‡βˆ—πΊ using the optimal Hamiltonian on π’’βˆ—, where the maximum principle yields the optimal control. The energy-Casimir method is used to give sufficient conditions for nonlinear stability of the equilibrium states. Around this equilibrium states, we are able to find the periodical orbits using Moser's theorem. In the last paragraph, we have studied the numerical integration via three methods: Lie-Trotter algorithm, Kahan's algorithms, and Runge-Kutta 4th method. Numerical simulations and a comparison between these three methods are presented too.

2. The Geometrical Picture of the Problem

Let 𝐺 be the Lie group given by𝐺=1π‘₯2π‘₯40π‘’βˆ’π‘₯1π‘₯3ξƒ­001βˆˆβ„³3(ℝ)∣π‘₯1,π‘₯2,π‘₯3,π‘₯4ξƒ°βˆˆβ„.(2.1)

Proposition 2.1. The Lie algebra 𝒒 of 𝐺 is generated by 𝐴1=0000βˆ’10000,𝐴2=,𝐴0100000003=000001000,𝐴4=,001000000(2.2) and the Lie algebra structure of 𝒒 is given by the following []β‹…,⋅𝐴1𝐴2𝐴3𝐴4𝐴10𝐴2βˆ’π΄30𝐴2βˆ’π΄20𝐴40𝐴3𝐴3βˆ’π΄400𝐴40000.(2.3)

Proposition 2.2. The minus-Lie-Poisson structure on π’’βˆ—β‰ƒ(𝑅4)βˆ—β‰ƒπ‘…4 is generated by the matrix Ξ βˆ’=⎑⎒⎒⎣0βˆ’π‘₯2π‘₯30π‘₯20βˆ’π‘₯40βˆ’π‘₯3π‘₯4⎀βŽ₯βŽ₯⎦000000.(2.4)
An easy computation leads one to the following.

Proposition 2.3. The following two systems are drift-free-left invariant controllable systems on 𝐺, namely: ̇𝐴𝑋=𝑋1𝑒1+𝐴2𝑒2+𝐴3𝑒3̇𝐴,(2.5)𝑋=𝑋1𝑒1+𝐴2𝑒2+𝐴3𝑒3+𝐴4𝑒4ξ€Έ,(2.6) where π‘‹βˆˆπΊ,𝐴𝑖 are the matrix defined above, and π‘’π‘–βˆˆπΆβˆž(ℝ,ℝ),𝑖=1,4.

Proof. Since the span of the set of Lie brackets generated by 𝐴1,𝐴2, and 𝐴3 coincides with 𝒒 the proposition is a consequence of a result due to Jurdjevic and Sussmann [6].

3. An Optimal Control Problem for the System (2.5)

Let 𝐽 be the cost function given by𝐽𝑒1,𝑒2,𝑒3ξ€Έ=12ξ€œπ‘‘π‘“0𝑐1𝑒21(𝑑)+𝑐2𝑒22(𝑑)+𝑐3𝑒23(𝑑)𝑑𝑑𝑐1>0,𝑐2>0,𝑐3>0.(3.1)

Then we have the following.

Proposition 3.1. The controls that minimize 𝐽 and steer the system (2.5) from 𝑋=𝑋0 at 𝑑=0 to 𝑋=𝑋𝑓 at 𝑑=𝑑𝑓 are given by 𝑒1=1𝑐1π‘₯1,𝑒2=1𝑐2π‘₯2,𝑒3=1𝑐3π‘₯3,(3.2) where π‘₯𝑖’s are solutions of Μ‡π‘₯11=βˆ’π‘2π‘₯22+1𝑐3π‘₯23,Μ‡π‘₯2=1𝑐1π‘₯1π‘₯2βˆ’1𝑐3π‘₯3π‘₯4,Μ‡π‘₯31=βˆ’π‘1π‘₯1π‘₯3+1𝑐2π‘₯2π‘₯4,Μ‡π‘₯4=0.(3.3)

Proof. Let us consider the optimal Hamiltonian given by 𝐻π‘₯1,π‘₯2,π‘₯2,π‘₯4ξ€Έ=12π‘₯21𝑐1+π‘₯22𝑐2+π‘₯23𝑐3ξƒͺ.(3.4) It is in fact the controlled Hamiltonian 𝐻opt given by 𝐻opt=π‘₯1𝑒1+π‘₯2𝑒2+π‘₯3𝑒3βˆ’12𝑐1𝑒21+𝑐2𝑒22+𝑐3𝑒23ξ€Έ,(3.5) which is reduced to π’’βˆ— via Poisson reduction. Then the optimal controls are given by 𝑒1=1𝑐1π‘₯1,𝑒2=1𝑐2π‘₯2,𝑒1=1𝑐3π‘₯3,(3.6) where π‘₯𝑖’s are solutions of the reduced Hamilton's equations given by ξ€Ίπ‘₯1π‘₯2π‘₯3π‘₯4𝑑=Ξ βˆ’β‹…βˆ‡π»,(3.7) which is nothing else than the required (3.3).

Remark 3.2. It is easy to see from (3.3) that π‘₯4=constant, and so the dynamics (3.3) can be put in the equivalent form Μ‡π‘₯11=βˆ’π‘2π‘₯22+1𝑐3π‘₯23,Μ‡π‘₯2=1𝑐1π‘₯1π‘₯2βˆ’π‘˜π‘3π‘₯3,Μ‡π‘₯31=βˆ’π‘1π‘₯1π‘₯3+π‘˜π‘2π‘₯2.(3.8)

The goal of our paper is to study some geometrical and dynamical properties for the system (3.8).

Proposition 3.3. The dynamics (3.8) has the following Hamilton-Poisson realization: ℝ3ξ€Έ,,Ξ ,𝐻(3.9) where Π=0βˆ’π‘₯2π‘₯2π‘₯20βˆ’π‘˜βˆ’π‘₯3ξƒ­π‘˜0(3.10) and the Hamiltonian 𝐻π‘₯1,π‘₯2,π‘₯3ξ€Έ=12π‘₯21𝑐1+π‘₯22𝑐2+π‘₯23𝑐3ξƒͺ.(3.11)

Proof. Indeed, it is not hard to see that the dynamics (3.8) can be put in the equivalent form ξ€ΊΜ‡π‘₯1,Μ‡π‘₯2,Μ‡π‘₯3𝑑=Ξ β‹…βˆ‡π»,(3.12) as required. Moreover, the function 𝐢 given by 𝐢=π‘˜π‘₯1+π‘₯2π‘₯3(3.13) is a Casimir of our configuration. Indeed, (βˆ‡πΆ)𝑑Π=0,(3.14) as desired.

Remark 3.4. The phase curves of the dynamics (3.8) are intersections of π‘₯21𝑐1+π‘₯22𝑐2+π‘₯23𝑐3=const.,(3.15) with π‘˜π‘₯1+π‘₯2π‘₯3=const.,(3.16) see Figure 1.

Proposition 3.5. The dynamics (3.8) has an infinite number of Hamilton-Poisson realizations.

Proof. An easy computation shows us that the triples ℝ3,{β‹…,β‹…}π‘Žπ‘,𝐻𝑐𝑑,(3.17) where {𝑓,𝑔}π‘Žπ‘=βˆ’βˆ‡πΆπ‘Žπ‘β‹…(βˆ‡π‘“Γ—βˆ‡π‘”),(βˆ€)𝑓,π‘”βˆˆπΆβˆžξ€·β„3ξ€Έ,𝐢,β„π‘Žπ‘=π‘ŽπΆ+𝑏𝐻,𝐻𝑐𝑑=𝑐𝐢+𝑑𝐻,π‘Ž,𝑏,𝑐,π‘‘βˆˆβ„,π‘Žπ‘‘βˆ’π‘π‘=1,(3.18) define Hamilton-Poisson realizations of the dynamics (3.8), as required.

Remark 3.6. The above proposition tells us in fact that (3.8) is unchanged, so the trajectories of motion in ℝ3 remain the same when 𝐻 and 𝐢 are replaced by linear combinations of 𝐻 and 𝐢 with coefficients which form a real matrix with det one.

4. Stability and Periodical Orbits

It is not hard to see that the equilibrium states of our dynamics (3.8) are𝑒𝑀1𝑒=(𝑀,0,0),π‘€βˆˆβ„,𝑀2=ξƒ©βˆ’π‘˜π‘1βˆšπ‘2𝑐3ξ‚™,βˆ’π‘2𝑐3ξƒͺ𝑒𝑀,𝑀,π‘€βˆˆβ„,𝑀3=ξƒ©π‘˜π‘1βˆšπ‘2𝑐3,𝑐2𝑐3ξƒͺ𝑀,𝑀,π‘€βˆˆβ„.(4.1)

Let 𝐴 be the matrix of the linear part of the system (3.8), that is,βŽ›βŽœβŽœβŽœβŽœβŽœβŽœβŽœβŽ2A=0βˆ’π‘2π‘₯22𝑐3π‘₯31𝑐1π‘₯21𝑐1π‘₯1βˆ’π‘˜π‘3βˆ’1𝑐1π‘₯3π‘˜π‘2βˆ’1𝑐1π‘₯1⎞⎟⎟⎟⎟⎟⎟⎟⎠,(4.2) then the characteristic roots of 𝐴(𝑒𝑀1), respectively, 𝐴(𝑒𝑀2), respectively, and 𝐴(𝑒𝑀3) are given byπœ†1=0,πœ†2,3=±𝑐2𝑐3𝑀2βˆ’π‘1π‘˜2𝑐1βˆšπ‘2𝑐3,(4.3) respectively,πœ†1=0,πœ†2,3=Β±2π‘–βˆšπ‘1𝑐3𝑀,(4.4) respectively,πœ†1=0,πœ†2,3=Β±2π‘–βˆšπ‘1𝑐3𝑀,(4.5) so we can conclude with the following.

Proposition 4.1. The equilibrium states 𝑒𝑀1,π‘€βˆˆβ„, are spectrally stable if π‘€βˆˆ(βˆ’(𝑐1βˆšπ‘˜/𝑐2𝑐3),(𝑐1βˆšπ‘˜/𝑐2𝑐3)).

Proposition 4.2. The equilibrium states 𝑒𝑀2 and 𝑒𝑀3,π‘€βˆˆβ„, are spectrally stable for any π‘€βˆˆπ‘….

We can now pass to discuss the nonlinear stability of the equilibrium states 𝑒𝑀1,𝑒𝑀2, and 𝑒𝑀3, π‘€βˆˆβ„.

Proposition 4.3. The equilibrium states 𝑒𝑀1,π‘€βˆˆβ„βˆ—, are nonlinearly stable if π‘€βˆˆ(βˆ’(𝑐1βˆšπ‘˜/𝑐2𝑐3),(𝑐1βˆšπ‘˜/𝑐2𝑐3)).

Proof. We will make the proof using energy-Casimir method (see [7]). Let π»πœ‘π‘₯=𝐻+πœ‘(𝐢)=212𝑐1+π‘₯222𝑐2+π‘₯232𝑐3ξ€·+πœ‘π‘˜π‘₯1+π‘₯2π‘₯3ξ€Έ(4.6) be the energy-Casimir function, where πœ‘βˆΆπ‘…β†’π‘… is a smooth real-valued function defined on 𝑅.
Now, the first variation of π»πœ‘ is given by π›Ώπ»πœ‘=π‘₯1𝑐1𝛿π‘₯1+π‘₯2𝑐2𝛿π‘₯2+π‘₯3𝑐3𝛿π‘₯3ξ€·+Μ‡πœ‘π‘˜π‘₯1+π‘₯2π‘₯3ξ€Έβ‹…ξ€·π‘˜π›Ώπ‘₯1+π‘₯2𝛿π‘₯3+π‘₯3𝛿π‘₯2ξ€Έ,(4.7) where Μ‡πœ‘=πœ•πœ‘πœ•ξ€·π‘˜π‘₯1+π‘₯2π‘₯3ξ€Έ.(4.8) This equals zero at the equilibrium of interest if and only if Μ‡πœ‘(π‘˜π‘€)=βˆ’π‘€π‘˜π‘1.(4.9) The second variation of π»πœ‘ is given by 𝛿2π»πœ‘=1𝑐1𝛿π‘₯1ξ€Έ2+1𝑐2𝛿π‘₯2ξ€Έ2+1𝑐3𝛿π‘₯3ξ€Έ2ξ€·+Μˆπœ‘β‹…π‘˜π›Ώπ‘₯1+π‘₯2𝛿π‘₯3+π‘₯3𝛿π‘₯2ξ€Έ2+2Μ‡πœ‘β‹…π›Ώπ‘₯2𝛿π‘₯3.(4.10) Since π‘€βˆˆ(βˆ’(𝑐1βˆšπ‘˜/𝑐2𝑐3),(𝑐1βˆšπ‘˜/𝑐2𝑐3)) and if we choose the function πœ‘ such that π‘€Μ‡πœ‘(π‘˜π‘€)=βˆ’π‘˜π‘1,Μˆπœ‘(π‘˜π‘€)>0,(4.11) we can conclude that the second variation of π»πœ‘ at the equilibrium of interest is positive definite and, thus, 𝑒𝑀1 are nonlinearly stable.

Similar arguments lead us to the following result.

Proposition 4.4. The equilibrium states 𝑒𝑀2,π‘€βˆˆβ„βˆ— and 𝑒𝑀3,π‘€βˆˆβ„βˆ— are nonlinearly stable for any π‘€βˆˆβ„βˆ—.

In order to find the periodical orbits around the equilibrium states 𝑒𝑀1, we make use of the property that the dynamics described by a Hamilton-Poisson system takes place on the symplectic leaves of the Poisson configuration manifold, to prove the existence of periodic orbits by looking for periodic orbits of the symplectic Hamiltonian completely integrable system obtained by the restriction of the Lorenz system to the regular coadjoint orbits of πΊβˆ—. This procedure will be implemented around nonlinearly stable equilibrium states. The procedure is the following: we consider the system restricted to a regular coadjoint orbit of πΊβˆ— that contains a nonlinearly stable equilibrium, and then we will get the existence of periodic solutions for the restricted system. These periodic solutions are periodic solutions also for the unrestricted system.

Proposition 4.5. Near 𝑒𝑀1=(𝑀,0,0), π‘€βˆˆ(βˆ’(𝑐1βˆšπ‘˜/𝑐2𝑐3),(𝑐1βˆšπ‘˜/𝑐2𝑐3)), the reduced dynamics has, for each sufficiently small value of the reduced energy, at least 1 periodic solution whose period is close to 2πœ‹π‘1βˆšπ‘2𝑐3ξ”π‘˜2𝑐21βˆ’π‘2𝑐3𝑀2.(4.12)

Proof. Indeed, we have successively the following:(i) the restriction of our dynamics (3.8) to the coadjoint orbit π‘˜π‘₯1+π‘₯2π‘₯3=π‘˜π‘€(4.13) gives rise to a classical Hamiltonian system,(ii) the matrix of the linear part of the reduced dynamics has purely imaginary roots. More exactly πœ†2,3=Β±π‘–π‘˜2𝑐21βˆ’π‘2𝑐3𝑀2𝑐1βˆšπ‘2𝑐3,(4.14)(iii)consider the following: 𝑒spanβˆ‡πΆπ‘€1ξ€Έξ€Έ=𝑉0,(4.15) where 𝑉0𝐴𝑒=ker𝑀1100ξƒͺξ€Έξ€Έ=span,(4.16)(iv)the reduced Hamiltonian has a local minimum at the equilibrium state 𝑒𝑀1 (see the proof of Proposition 4.3).
Then our assertion follows via the Moser-Weinstein theorem with zero eigenvalue; see [8] for details.

Remark 4.6. The existence of the periodical orbits around the equilibrium points 𝑒2 and 𝑒3 remains an open problem, the Moser-Weinstein theorem with zero eigenvalue being inconclusive.

5. Lax Formulation and Numerical Integration of the Dynamics (3.8)

Proposition 5.1. The dynamics (3.8) allows a formulation in terms of Lax pairs.

Proof. Let us take the following: 𝐿=0𝑙12𝑙13βˆ’π‘™120𝑙23βˆ’π‘™13βˆ’π‘™230ξƒ­,𝑙12=2βˆšπ‘3π‘₯2𝑐2βˆšπ‘1+2π‘₯3βˆšπ‘1𝑐2+√2ξ‚€βˆšπ‘2βˆšβˆ’22βˆ’π‘1𝑐2𝑐2βˆšπ‘18βˆšπ‘˜βˆ’π‘3𝑐2βˆšπ‘1,𝑙132√=βˆ’π‘3π‘₯1βˆšπ‘1βˆ’ξ”2𝑐3ξ€·2βˆ’π‘1𝑐2𝑐2βˆšπ‘1π‘₯2βˆ’ξƒŽ2ξ€·2βˆ’π‘1𝑐2𝑐1𝑐2π‘₯3+𝑐2√2βˆ’π‘1𝑐2βˆ’4𝑐2βˆšπ‘1𝑐24ξ”π‘˜βˆ’2𝑐3ξ€·2βˆ’π‘1𝑐2𝑐2βˆšπ‘1,𝑙232=βˆ’π‘3ξ€·2βˆ’π‘1𝑐2𝑐1βˆšπ‘2π‘₯1+ξƒŽ2𝑐3𝑐2π‘₯2+√2π‘₯3ξƒŽ+42𝑐3𝑐2ξƒ¬βˆ’π‘˜,𝐡=0𝑏12𝑏13βˆ’π‘120𝑏23βˆ’π‘13βˆ’π‘230ξƒ­,𝑏12=𝑐2ξ‚€2√2π‘₯3ξ‚βˆš+π‘˜βˆ’22βˆ’π‘1𝑐2βˆšπ‘˜βˆ’42𝑐2𝑐32𝑐2βˆšπ‘1𝑐3,𝑏13=βˆšβˆ’22𝑐2βˆšπ‘2𝑐3π‘₯1++2π‘˜π‘2ξ€·2βˆ’π‘1𝑐2ξ€Έξ‚€βˆšβˆ’4𝑐32+π‘₯2ξ€Έβˆš+2𝑐2π‘˜ξ‚4𝑐2βˆšπ‘1𝑐3,𝑏23√=βˆ’2βˆ’π‘1𝑐2𝑐1√2π‘₯1+π‘₯2βˆ’βˆšπ‘22√2𝑐3π‘˜+2,(5.1) then, using MATHEMATICA 7.0, we can put the system (3.8) in the equivalent form Μ‡[],𝐿=𝐿,𝐡(5.2) as desired.

We will discuss now the numerical integration of the dynamics (3.8) via the Kahan integrator, Lie-Trotter integrator [9], and also via Runge-Kutta 4th steps integrator, and we will point out some properties of Kahan and Lie-Trotter integrators.

It is easy to see that for the equations (3.8), Kahan's integrator can be written in the following form: π‘₯1𝑛+1βˆ’π‘₯𝑛1β„Ž=βˆ’π‘2π‘₯2𝑛+1π‘₯𝑛2+β„Žπ‘3π‘₯3𝑛+1π‘₯𝑛3,π‘₯2𝑛+1βˆ’π‘₯𝑛2=β„Ž2𝑐1ξ€·π‘₯1𝑛+1π‘₯𝑛2+π‘₯2𝑛+1π‘₯𝑛1ξ€Έβˆ’β„Žπ‘˜2𝑐3ξ€·π‘₯𝑛3+π‘₯3𝑛+1ξ€Έ,π‘₯3𝑛+1βˆ’π‘₯𝑛3β„Ž=βˆ’2𝑐1ξ€·π‘₯1𝑛+1π‘₯𝑛3+π‘₯3𝑛+1π‘₯𝑛1ξ€Έ+β„Žπ‘˜2𝑐2ξ€·π‘₯𝑛2+π‘₯2𝑛+1ξ€Έ.(5.3)

Using MATHEMATICA 8.0, we can prove the following proposition which shows the incompatibility of the Kahan's integrator with the Poisson structure of the system (3.8).

Proposition 5.2. Kahan's integrator (5.3) has the following properties:(i)it is not Poisson preserving;(ii)it does not preserve the Casimir 𝐢 of the Poisson configuration (ℝ3,Ξ );(iii)it does not preserve the Hamiltonian 𝐻 of the system (3.8).

We will discuss now the numerical integration of the dynamics (3.8) via the Lie-Trotter integrator.

To begin with, let us observe that the Hamiltonian vector field 𝑋𝐻 splits as follows:𝑋𝐻=𝑋𝐻1+𝑋𝐻2+𝑋𝐻3,(5.4) where𝐻1=12𝑐1π‘₯21,𝐻2=12𝑐3π‘₯22,𝐻3=12𝑐3π‘₯23.(5.5) Their corresponding integral curves are, respectively, given byπ‘₯1(π‘₯𝑑)2π‘₯(𝑑)3ξƒ­(𝑑)=𝐴𝑖π‘₯1(π‘₯0)2π‘₯(0)3ξƒ­(0),𝑖=1,3,(5.6) where𝐴1=1000𝑒(π‘Ž/𝑐1)𝑑000π‘’βˆ’(π‘Ž/𝑐1)𝑑,π‘Ž=π‘₯1𝐴(0),2=βŽ‘βŽ’βŽ’βŽ’βŽ’βŽ£π‘1βˆ’π‘20π‘˜π‘‘0010𝑐2⎀βŽ₯βŽ₯βŽ₯βŽ₯⎦,𝑑1𝑏=π‘₯2𝐴(0),3=βŽ‘βŽ’βŽ’βŽ’βŽ£π‘10𝑐3π‘‘π‘˜01βˆ’π‘3π‘‘βŽ€βŽ₯βŽ₯βŽ₯⎦,001𝑐=π‘₯3(0).(5.7) Then the Lie-Trotter integrator is given byπ‘₯1𝑛+1=π‘₯𝑛1βˆ’π‘π‘2𝑑π‘₯𝑛2+𝑐𝑐3𝑑+π‘˜π‘π‘2𝑐3𝑑2ξ‚Άπ‘₯𝑛3,π‘₯2𝑛+1=𝑒(π‘Ž/𝑐1)𝑑π‘₯𝑛2βˆ’π‘˜π‘3𝑑𝑒(π‘Ž/𝑐1)𝑑π‘₯𝑛3,π‘₯3𝑛+1=π‘˜π‘2𝑑𝑒(βˆ’π‘Ž/𝑐1)𝑑π‘₯𝑛2βˆ’π‘˜2𝑐2𝑐3𝑑2𝑒(βˆ’π‘Ž/𝑐1)𝑑π‘₯𝑛3.(5.8)

Now, a direct computation or using MATHEMATICA leads us to the following.

Proposition 5.3. The Lie-Trotter integrator (5.8) has the following properties:(i) it preserves the Poisson structure Ξ ;(ii)it preserves the Casimir 𝐢 of the Poisson configuration (ℝ3,Ξ );(iii)it does not preserve the Hamiltonian 𝐻 of the system (3.8);(iv)its restriction to the coadjoint orbit (π’ͺπ‘˜,πœ”π‘˜), where π’ͺπ‘˜=π‘₯ξ€½ξ€·1,π‘₯2,π‘₯3ξ€Έβˆˆβ„3βˆ£π‘˜π‘₯1+π‘₯2π‘₯3ξ€Ύ=const.(5.9)and πœ”π‘˜ is the Kirillov-Kostant-Souriau symplectic structure on π’ͺπ‘˜, gives rise to a symplectic integrator.

Remark 5.4. If we compare these methods, see Figures 2, 3, and 4, with the 4th-step Runge-Kutta method, we can see that Lie-Trotter integrator and Kahan's integrator give us a weak approximation of our dynamics. However, Kahan's integrator and the Lie-Trotter integrator have the advantage of being more easily implemented.

6. Conclusion

The paper analyses a drift-free left invariant controllable system on a special Lie group. The Hamilton-Poisson realization of the system allows us to study the system from the mechanical geometry point of view. This means that we can use specific tools as energy-Casimir method for nonlinear stability, the Moser's theorem to find the periodical orbits, and Poisson integrators to make the numerical integration of the dynamics. In addition, we use non-Poisson integrators (Kahan's integrator and Runge-Kutta 4th-step integrator) to make a comparison between the obtained results. Numerical simulations via MATHEMATICA 8.0 are presented too. Similar problems have been studied on the Lie groups 𝑆𝑂(3),𝑆𝑂(4) (see [10]), on the Heisenberg Lie groups 𝐻(3) and 𝐻(4), or 𝑆𝐸(2,ℝ)×𝑆𝑂(3) (see [11]).

Acknowledgments

This paper was supported by the project β€œDevelopment and support of multidisciplinary postdoctoral programmes in major technical areas of national strategy for Research-Development-Innovation” 4D-POSTDOC, Contract no. POSDRU/89/1.5/S/52603, and the project was cofunded by the European Social Fund through Sectorial Operational Programme Human Resources Development 2007–2013.