On local normal forms of completely integrable systems

Similar documents
Dynamical systems on Leibniz algebroids

On a Hamiltonian version of a 3D Lotka-Volterra system

Unimodularity and preservation of measures in nonholonomic mechanics

Chap. 1. Some Differential Geometric Tools

Week 6: Differential geometry I

SYMPLECTIC MANIFOLDS, GEOMETRIC QUANTIZATION, AND UNITARY REPRESENTATIONS OF LIE GROUPS. 1. Introduction

[#1] R 3 bracket for the spherical pendulum

The dynamical rigid body with memory

Hamiltonian Dynamics from Lie Poisson Brackets

Hamiltonian Dynamics

LECTURE 3 MATH 261A. Office hours are now settled to be after class on Thursdays from 12 : 30 2 in Evans 815, or still by appointment.

Symplectic and Poisson Manifolds

LIE ALGEBROIDS AND POISSON GEOMETRY, OLIVETTI SEMINAR NOTES

Definition 5.1. A vector field v on a manifold M is map M T M such that for all x M, v(x) T x M.

Hamiltonian Systems of Negative Curvature are Hyperbolic

M3-4-5 A16 Notes for Geometric Mechanics: Oct Nov 2011

Hamiltonian flows, cotangent lifts, and momentum maps

Course Summary Math 211

About Integrable Non-Abelian Laurent ODEs

LECTURE 16: LIE GROUPS AND THEIR LIE ALGEBRAS. 1. Lie groups

arxiv: v1 [math-ph] 5 May 2015

A MARSDEN WEINSTEIN REDUCTION THEOREM FOR PRESYMPLECTIC MANIFOLDS

2 Constructions of manifolds. (Solutions)

ON THE PARALLEL PLANES IN VECTOR BUNDLES

Lecture I: Constrained Hamiltonian systems

Hamilton-Jacobi theory on Lie algebroids: Applications to nonholonomic mechanics. Manuel de León Institute of Mathematical Sciences CSIC, Spain

Contact and Symplectic Geometry of Monge-Ampère Equations: Introduction and Examples

LECTURE 8: THE MOMENT MAP

GEOMETRIC QUANTIZATION

Cocycles and stream functions in quasigeostrophic motion

Poisson Manifolds Bihamiltonian Manifolds Bihamiltonian systems as Integrable systems Bihamiltonian structure as tool to find solutions

Cost-extended control systems on SO(3)

Geometrical mechanics on algebroids

THEODORE VORONOV DIFFERENTIABLE MANIFOLDS. Fall Last updated: November 26, (Under construction.)

EXERCISES IN POISSON GEOMETRY

Vectors. January 13, 2013

Invariant Nonholonomic Riemannian Structures on Three-Dimensional Lie Groups

Numerical Algorithms as Dynamical Systems

Superintegrable 3D systems in a magnetic field and Cartesian separation of variables

BACKGROUND IN SYMPLECTIC GEOMETRY

arxiv:math-ph/ v1 25 Feb 2002

1. Tangent Vectors to R n ; Vector Fields and One Forms All the vector spaces in this note are all real vector spaces.

Problems in Linear Algebra and Representation Theory

A little taste of symplectic geometry

On homogeneous Randers spaces with Douglas or naturally reductive metrics

SUBTANGENT-LIKE STATISTICAL MANIFOLDS. 1. Introduction

MANIN PAIRS AND MOMENT MAPS

On Quadratic Hamilton-Poisson Systems

arxiv: v1 [math-ph] 31 Jan 2015

A LITTLE TASTE OF SYMPLECTIC GEOMETRY: THE SCHUR-HORN THEOREM CONTENTS

,, rectilinear,, spherical,, cylindrical. (6.1)

Symmetry Reductions of (2+1) dimensional Equal Width. Wave Equation

Recent Progress in the Integration of Poisson Systems via the Mid Point Rule and Runge Kutta Algorithm

ON CONTRAVARIANT PRODUCT CONJUGATE CONNECTIONS. 1. Preliminaries

4(xo, x, Po, P) = 0.

Group Actions and Cohomology in the Calculus of Variations

HANDOUT #12: THE HAMILTONIAN APPROACH TO MECHANICS

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM

Lecture Notes a posteriori for Math 201

Linear Algebra: Matrix Eigenvalue Problems

Superintegrability in a non-conformally-at space

3.1. Derivations. Let A be a commutative k-algebra. Let M be a left A-module. A derivation of A in M is a linear map D : A M such that

Reminder on basic differential geometry

Topics in Representation Theory: Lie Groups, Lie Algebras and the Exponential Map

Formal Groups. Niki Myrto Mavraki

Lecture 2: Controllability of nonlinear systems

B.7 Lie Groups and Differential Equations

The Erwin Schrodinger International Boltzmanngasse 9. Institute for Mathematical Physics A-1090 Wien, Austria

The Geometry of Euler s equation. Introduction

Stabilization and Passivity-Based Control

CS 468 (Spring 2013) Discrete Differential Geometry

CS 468. Differential Geometry for Computer Science. Lecture 13 Tensors and Exterior Calculus

Dierential geometry for Physicists

[2] (a) Develop and describe the piecewise linear Galerkin finite element approximation of,

Lie Matrix Groups: The Flip Transpose Group

DIFFERENTIAL GEOMETRY CLASS NOTES INSTRUCTOR: F. MARQUES. September 25, 2015

Math 360 Linear Algebra Fall Class Notes. a a a a a a. a a a

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

Math 554 Qualifying Exam. You may use any theorems from the textbook. Any other claims must be proved in details.

AN ALMOST KÄHLER STRUCTURE ON THE DEFORMATION SPACE OF CONVEX REAL PROJECTIVE STRUCTURES

Quasi-invariant Measures on Path Space. Denis Bell University of North Florida

Real Analysis Prelim Questions Day 1 August 27, 2013

Math 4377/6308 Advanced Linear Algebra

satisfying the following condition: If T : V V is any linear map, then µ(x 1,,X n )= det T µ(x 1,,X n ).

Higher order Koszul brackets

Solutions of M3-4A16 Assessed Problems # 3 [#1] Exercises in exterior calculus operations

PHYS 705: Classical Mechanics. Rigid Body Motion Introduction + Math Review

On universality of critical behaviour in Hamiltonian PDEs

Stability Subject to Dynamical Accessibility

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M =

The geometry of hydrodynamic integrability

Linear Algebra (Review) Volker Tresp 2017

An introduction to Birkhoff normal form

arxiv: v1 [math.sg] 8 Sep 2017

Ruijsenaars type deformation of hyperbolic BC n Sutherland m

Vortex Equations on Riemannian Surfaces

A Note on Poisson Symmetric Spaces

THEODORE VORONOV DIFFERENTIAL GEOMETRY. Spring 2009

LECTURE 26: THE CHERN-WEIL THEORY

Chap. 3. Controlled Systems, Controllability

Transcription:

On local normal forms of completely integrable systems ENCUENTRO DE Răzvan M. Tudoran West University of Timişoara, România Supported by CNCS-UEFISCDI project PN-II-RU-TE-2011-3-0103 GEOMETRÍA DIFERENCIAL, ROSARIO, 02 August 2012

Table of contents 1 Abstract 2 Completely integrable systems as Hamilton-Poisson systems 3 Local normal form of a completely integrable system 4 Complete integrability versus symmetry 5 Examples 6 Euler s equations versus normal form of 3d quadratic completely integr 7 Bibliography

Abstract The purpose of this talk is to show that a differential system on R n which admits a set of n 1 independent conservation laws defined on an open subset Ω R n, is essentially equivalent on an open and dense subset U Ω, with the linear differential system u 1 = u 1, u 2 = u 2,..., u n = u n. Consequently, the core of the (local) complete integrability is actually the integrability of the simplest linear differential equation, namely u = u. Another consequence of this result is that locally, a completely integrable differential system admits symmetries.

Completely integrable systems Let us consider a differential system on R n : ẋ 1 = X 1 (x 1,...,x n ) ẋ 2 = X 2 (x 1,...,x n ) ẋ n = X n (x 1,...,x n ), (1) where X 1,X 2,...,X n C (R n,r) are arbitrary real functions. Suppose that C 1,...,C n 2,C n 1 :Ω R n R are n 1 independent C integrals of motion of (1) defined on a nonempty open subset Ω R n.

Completely integrable systems Since C 1,...,C n 2,C n 1 :Ω R n R are integrals of motion of the vector field X = X 1 x1 + + X n xn X(R n ), we obtain that for each i {1,...,n 1} L X C i = C i (x),x (x) = n j=1 xj C i ẋ j =0, for every x =(x 1,...,x n ) Ω, where, stand for the canonical inner product on R n, and respectively stand for the gradient with respect to,.

Completely integrable systems Hence, the vector field X can be written as the vector field ( C 1 C n 1 ) multiplied by a real function (rescaling function), where stand for the Hodge star operator for multivector fields. It may happen that the domain of definition for the rescaling function to be a proper subset of Ω. In the following we will consider the generic case when the rescaling function is defined on an open and dense subset of Ω. In order to simplify the notations, we will also denote this set by Ω.

Completely integrable systems In coordinates, the completely integrable system ẋ = X (x) canbe written as: where ν is the rescaling function. ẋ 1 = ν (C 1,...,C n 1,x 1 ) (x 1,...,x n ) ẋ 2 = ν (C 1,...,C n 1,x 2 ) (x 1,...,x n ) ẋ n = ν (C 1,...,C n 1,x n ). (x 1,...,x n ) (2)

Poisson manifolds. Hamilton-Poisson dynamical systems A Poisson structure on a manifold P is a bilinear map (Poisson bracket) {, } : C (P,R) C (P,R) C (P,R) that defines a Lie algebra structure on C (P,R) and that is a derivation in each entry. The elements in the center of this algebra are called Casimir functions. A pair consisting of a smooth manifold P together with a Poisson structure, is called a Poisson manifold and is usually denoted by (P,{, }).

Hamilton-Poisson dynamical systems The derivation property of a Poisson structure on P allows us to assign to each function H C (P,R) a vector field X H X (P) via the equality X H (F ):={F,H} for every F C (P,R). The vector field X H X (P) is called the Hamilton-Poisson vector field or the Hamiltonian vector field associated to the Hamiltonian function H. The triple (P,{, },H) is called Hamilton-Poisson dynamical system.

Hamilton-Poisson dynamical systems The derivation property of the Poisson bracket also implies that for any two functions f,g C (P,R), the value of the bracket {f,g}(z) at an arbitrary point z P depends on f only through df (z) which allows us to define a contravariant antisymmetric two-tensor Π Λ 2 (P) by Π(z)(α z,β z )={f,g}(z), where df (z)=α z T z P and dg(z)=β z T z P. Thistensoris called Poisson tensor or Poisson structure on P.

Hamilton-Poisson dynamical systems The vector bundle map Π : T P TP naturally associated to Π is defined by Π(z)(α z,β z )= α z,π (β z ). Note that for any H C (P,R) the Hamiltonian vector field X H verifies that X H =Π (dh).

Hamilton-Poisson dynamical systems If one denote by n the dimension of the manifold P, the expression of the Poisson tensor relative to a local coordinates system (x 1,...,x n ) is given by the bivector field Π= Π kl (x 1,...,x n ) xk xl, 1 k<l n where Π kl (x 1,...,x n ):={x k,x l }. If there is no danger of confusion, the skew-symmetric matrix Π:=[Π kl (x 1,...,x n )] 1 k<l n is also called the Poisson structure.

Hamilton-Poisson dynamical systems Using the local expression of the map Π one get the local expression of the Hamiltonian vector field associated to a function H C (P,R) begivenby ( ) X H = 1 i n Π ij xj H 1 j n xi.

Hamilton-Poisson realization of completely integrable systems Recall that the vector field X generating the completely integrable system (1) is proportional to the vector field ( C 1 C n 1 ) where stand for the Hodge star operator for multivector fields. Consequently, the vector field X can be realized on the open set Ω R n as the Hamilton-Poisson vector field X H X(Ω) with respect to the Hamiltonian function H := C n 1 and respectively the Poisson bracket defined by: {f,g} ν;c1,...,c n 2 dx 1 dx n = νdc 1...dC n 2 df dg, where ν is a given real function (rescaling).

Hamilton-Poisson realization of completely integrable systems Equivalently, the above Poisson bracket {, } ν;c1,...,c n 2 {f,g} ν;c1,...,c n 2 = ν (C 1,...,C n 2,f,g). (x 1,...,x n ) is given by: The functions C 1,...,C n 2 form a complete set of Casimirs for the Poisson bracket {, } ν;c1,...,c n 2.

Hamilton-Poisson realization of completely integrable systems The Hamiltonian vector field X = X H is acting on a smooth real function as: X H (f )={f,h} ν;c1,...,c n 2.

Hamilton-Poisson realization of completely integrable systems Hence, the completely integrable system (1) can be locally written in Ω as a Hamilton-Poisson dynamical system of the type: ẋ 1 = {x 1,H} ν;c1,...,c n 2 ẋ 2 = {x 2,H} ν;c1,...,c n 2 ẋ n = {x n,h} ν;c1,...,c n 2.

Hamilton-Poisson realization of completely integrable systems Equivalently the above system can be written as: ẋ 1 = ν (C 1,...,C n 2,x 1,H) (x 1,...,x n ) ẋ 2 = ν (C 1,...,C n 2,x 2,H) (x 1,...,x n ) ẋ n = ν (C 1,...,C n 2,x n,h). (x 1,...,x n ) (3)

Divergence free vector field associated with the vector filed X Next result gives a method to construct a divergence free vector field out of the vector field X. The divergence operator we will use in this approach is the divergence associated with the standard Lebesgue measure on R n, namely L X (dx 1 dx n )=(divx )dx 1 dx n. Theorem The vector field X := 1 ν X is a divergence free vector field on Ω \ Z (ν), wherez (ν)={(x 1,...,x n ) Ω ν(x 1,...,x n )=0}.

Generic Hypothesis Assume that there exists a C rescaling function ν such that the Lebesgue measure of the set { O := x Ω 0 div(x )(x) (1/ν,C } 1,...,C n 2,H) (x)=0 (x 1,...,x n ) in Ω 0 is zero.

Main Theorem In the above hypothesis, the change of variables ((x 1,...,x n ),t) ((u 1,...,u n ),s) givenby u 1 =1/ν(x 1,...,x n ) u 2 = C 1 (x 1,...,x n )/ν(x 1,...,x n ) u n 1 = C n 2 (x 1,...,x n )/ν(x 1,...,x n ) u n = H(x 1,...,x n )/ν(x 1,...,x n ) ds = div(x )dt, in the open and dense subset Ω 00 := Ω 0 \ O of Ω 0, transforms the completely integrable system (3) restricted to Ω 00 into the linear differential system u 1 = u 1, u 2 = u 2,..., u n = u n, where prime stand for the derivative with respect to the new time s.

Local normal form of a completely integrable system Consequently we obtained that the core of the complete integrability is the integrability of the simplest linear differential equation u = u. Another consequence is that the interesting dynamical behavior of a completely integrable system is generated generically by a set of zero Lebesgue measure.

Local normal form of a completely integrable system If the rescaling function ν =: ν cst. is a constant function, then the Lebesgue measure of the set O =Ω 0 = Ω is nonzero in Ω 0,and hence the assumptions of the main theorem do not hold. In this case, we search for a new C rescaling function μ defined on an open and dense subset Ω 0 of Ω, such that the vector field μ X satisfies the assumptions of the main theorem.

Local normal form of a completely integrable system The function μ satisfies: div(μ X )= μ,x + μ div(x )= μ,x, since in the case of a constant function ν = ν cst. we have that 0=div(1/ν cst. X )=(1/ν cst. ) div(x ), and hence div(x )=0. Consequently, we have to search for a rescaling function μ such that div(μ X )= μ,x it is not identically zero in Ω 0.

Local normal form of a completely integrable system The second condition that μ has to satisfy is that the function it is not identically zero in Ω 0. (1/μ,C 1,...,C n 2,H) (x 1,...,x n )

Local normal form of a completely integrable system The transformation between the differential system (1) and the differential system generated by the vector field μ X is done by using the new time transformation dt = μ(x)dt. More exactly, the differential system (1) generated by X,namely: dx dt = X (x), is transformed via the new time transformation dt = μ(x)dt into the system: dx = μ(x) X (x). (4) dt The differential system (4) can also be realized as a Hamilton-Poisson dynamical system with respect to the Poisson bracket {, } μ νcst. ;C 1,...,C n 2 and respectively the same Hamiltonian H as for the Hamilton-Poisson realization (3) of the vector field X.

Complete integrability versus symmetry One of the most important results in the literature that study the implication symmetry integrability in the context of a general dynamical system is a classical result due to Lie which says that if one have n linearly independent vector fields X 1,...,X n on R n that generates a solvable Lie algebra under commutation: [X 1,X j ]=c 1 1,j X 1,[X 2,X j ]=c 1 1,j X 1 + c 2 2,j X 2,..., [X n,x j ]=c 1 1,j X 1 + c 2 2,j X 2 + + c n n,j X n,forj {1,...,n}, wherec k i,j are the structural constants of the Lie algebra, then the differential equation ẋ = X 1 (x) is integrable by quadratures.

Complete integrability versus symmetry Moreover, by solving these quadratures one obtain n 1 functionally independent first integrals of the system ẋ = X 1 (x), and consequently one get his complete integrability. This result was recently improved by Kozlov, by proving that in fact each of the differential equation ẋ = X j (x) is integrable by quadratures. Recall also that the above mentioned solvable Lie algebra integrates to a solvable n-dimensional Lie group that acts freely on R n and permutes the trajectories of the differential system, and hence is a symmetry group of the set of all trajectories of the system. Note that by permutation we mean that the group action maps trajectory to trajectory.

Complete integrability versus symmetry The purpose of this section is to study the converse implication: integrability symmetry. The main result states that for a completely integrable differential system ẋ = X (x), one can associate a vector field X, defined on an open and dense set, such that to each Y ker ( ad X ) it corresponds a Lie symmetry of the vector field X ; more exactly, there exists a real scalar function μ = μ(y ) and a vector field Y = Y (Y ) such that [X,Y ]=μ X. Since the existence of a pair consisting of a vector field Y and a scalar real function μ such that [X,Y ]=μ X imply the existence of a one-parameter Lie group which permutes the trajectories of the differential system ẋ = X (x), one obtains that on an open and dense set, integrability symmetry.

Complete integrability versus symmetry Let us start this section by recalling a classical result concerning Lie symmetries of dynamical systems. In order to do that, consider a n-dimensional dynamical system, ẋ = X (x),generatedbya vector field X = X 1 (x 1,...,x n ) x1 + + X n (x 1,...,x n ) xn,anda one-parameter Lie group of transformations, G,givenby x 1 (x 1,...,x n ;ε)=x 1 + εξ 1 (x 1,...,x n )+O(ε 2 ) x n(x 1,...,x n ;ε)=x n + εξ n (x 1,...,x n )+O(ε 2 ), acting on an open subset U R n with associated infinitesimal generator, Y = ξ 1 (x 1,...,x n ) x1 + + ξ n (x 1,...,x n ) xn X(U).

Complete integrability versus symmetry In the above hypothesis, a classical result states that G is a symmetry group for the dynamical system ẋ = X (x) (i.e., his action maps trajectory to trajectory), if Y is a Lie symmetry of X, in the sense that [X,Y ]=μ X, for some scalar function μ = μ(x 1,...,x n ).

Complete integrability versus symmetry Theorem In the hypothesis of main theorem, to each vector field Y ker ( ad X ),where X = u1 u1 + + u n un, the vector field Y =Φ Y is a Lie symmetry of the vector field X that generates the dynamical system (1), whereφ Y is the pull-back of the vector field Y through the local diffeomorphism Φ defined be the change of variables (x 1,...,x n ) (u 1,...,u n )=Φ(x 1,...,x n ) which brings the completely integrable system generated by the vector filed X to the linear differential system generated by the vector field X.

Three dimensional Lotka-Volterra system The 3D Lotka-Volterra system we is given by the following differential system: dx 1 dt = x 1(x 2 + x 3 ) dx 2 dt = x 2( x 1 + x 3 ) dx 3 dt = x 3( x 1 x 2 ). If one denote: (5) X =[x 1 (x 2 + x 3 )] x1 +[x 2 ( x 1 + x 3 )] x2 +[x 3 ( x 1 x 2 )] x3, then div(x )=2(x 3 x 1 ).

Three dimensional Lotka-Volterra system The system (5) admits a Hamilton-Poisson realization of the type (3), where: ν(x 1,x 2,x 3 )= x2 1 x2 3 x 1 + x 2 + x 3, C(x 1,x 2,x 3 )= x 2(x 1 + x 2 + x 3 ) x 1 x 3, H(x 1,x 2,x 3 )=x 1 + x 2 + x 3.

Three dimensional Lotka-Volterra system The sets introduced in the main theorem in the case of the Lotka-Volterra system are given by: Ω={(x 1,x 2,x 3 ) R 3 x 1 x 3 0}, Ω 0 = {(x 1,x 2,x 3 ) R 3 x 1 x 3 0; x 1 + x 2 + x 3 0}, O = {(x 1,x 2,x 3 ) R 3 x 1 = x 3 ; x 1 x 3 0; x 1 + x 2 + x 3 0}, Ω 00 = {(x 1,x 2,x 3 ) R 3 x 1 x 3 ; x 1 x 3 0;x 1 + x 2 + x 3 0}.

Three dimensional Lotka-Volterra system Then by the main theorem, the change of variables ((x 1,x 2,x 3 ),t) ((u 1,u 2,u 3 ),s) defined by: u 1 = x 1 + x 2 + x 3 x 2 1 x2 3 u 2 = x 2(x 1 + x 2 + x 3 ) 2 x 3 1 x3 3 u 3 = (x 1 + x 2 + x 3 ) 2 x 2 1 x2 3 ds =2(x 1 x 3 )dt, for (x 1,x 2,x 3 ) Ω 00 transforms the Lotka-Volterra system (5) into du 1 the linear differential system: ds = u 1, du 2 ds = u 2, du 3 ds = u 3.

3D Euler s equations of rigid body dynamics Recall that Euler s equations from the free rigid body dynamics are given by the differential system: dx 1 dt = I 2 I 3 x 2 x 3 I 2 I 3 dx 2 dt = I 3 I 1 x 1 x 3 (6) I 1 I 3 dx 3 dt = I 1 I 2 x 1 x 2, I 1 I 2 where the nonzero real numbers I 1,I 2,I 3 are the components of the inertia tensor. In the following we consider the case when I 2 I 3.

3D Euler s equations of rigid body dynamics If one denote: X =( I 2 I 3 I 2 I 3 x 2 x 3 ) x1 +( I 3 I 1 I 1 I 3 x 1 x 3 ) x2 +( I 1 I 2 I 1 I 2 x 1 x 2 ) x3, then div(x ) = 0, and hence the assumptions of the main theorem do not hold.

3D Euler s equations of rigid body dynamics The system (6) admits a Hamilton-Poisson realization of type (3), where: ν(x 1,x 2,x 3 )=:ν cst. (x 1,x 2,x 3 )= 1, C(x 1,x 2,x 3 )= 1 2 (x2 1 + x 2 2 + x 2 3 ), H(x 1,x 2,x 3 )= 1 2 (x2 1 I 1 + x2 2 I 2 + x2 3 I 3 ).

3D Euler s equations of rigid body dynamics In order to correct the vector field X such that one can apply the main theorem, we choose a rescaling function μ(x 1,x 2,x 3 )=x 1, and the associated new time transformation dt = μ(x 1,x 2,x 3 )dt defined on the open and dense subset of R 3 given by {(x 1,x 2,x 3 ) R 3 x 1 0}.

3D Euler s equations of rigid body dynamics This new time transformation, transforms the system (6) into the differential system: dx 1 dt = I 2 I 3 x 1 x 2 x 3 I 2 I 3 dx 2 dt = I 3 I 1 x1 2 I 1 I x 3 (7) 3 dx 3 dt = I 1 I 2 x1 2 I 1 I x 2, 2

3D Euler s equations of rigid body dynamics Recall that the vector field which generates the differential system (7) is given by: μ X =( I 2 I 3 I 2 I 3 x 1 x 2 x 3 ) x1 +( I 3 I 1 I 1 I 3 x 2 1 x 3 ) x2 +( I 1 I 2 I 1 I 2 x 2 1 x 2 ) x3. One note that the divergence of the vector field μ X is given by: div(μ X )= I 2 I 3 I 2 I 3 x 2 x 3. For I 2 I 3 we have that div(μ X ) it is not identically zero.

3D Euler s equations of rigid body dynamics The system (7) admits a Hamilton-Poisson realization of the type (3), where: ν(x 1,x 2,x 3 )=ν cst. (x 1,x 2,x 3 ) μ(x 1,x 2,x 3 )= x 1, C(x 1,x 2,x 3 )= 1 2 (x2 1 + x 2 2 + x 2 3 ), H(x 1,x 2,x 3 )= 1 2 (x2 1 I 1 + x2 2 I 2 + x2 3 I 3 ).

3D Euler s equations of rigid body dynamics The sets introduced in the main theorem in the case of the system (7) are given by: Ω=R 3, Ω 0 = {(x 1,x 2,x 3 ) R 3 x 1 0}, O = {(x 1,x 2,x 3 ) R 3 x 1 0; x 2 x 3 =0}, Ω 00 = {(x 1,x 2,x 3 ) R 3 x 1 x 2 x 3 0}.

3D Euler s equations of rigid body dynamics Then by the main theorem, the change of variables ((x 1,x 2,x 3 ),t ) ((u 1,u 2,u 3 ),s) defined by: u 1 = 1 x 1 u 2 = x2 1 + x2 2 + x2 3 2x 1 u 3 = x 1 x2 2 x2 3 2I 1 2x 1 I 2 2x 1 I 3 ds = I 3 I 2 x 2 x 3 dt, I 2 I 3 for (x 1,x 2,x 3 ) Ω 00 transforms the system (7) into the linear du 1 differential system: ds = u 1, du 2 ds = u 2, du 3 ds = u 3.

3d quadratic integrable systems In this section we show that any first-order autonomous three-dimensional differential equation possessing two independent quadratic constants of motion which admits a positive/negative definite linear combination, is affinely equivalent to the classical Euler s equations of the free rigid body dynamics with linear controls.

3d quadratic integrable systems as Hamilton-Poisson systems Let K be a 3 3 skew-symmetric matrix and k R 3.Letus introduce the Poisson manifold (R 3,{, } (K,k) ), where the Poisson bracket {, } (K,k) is defined by: {f,g} (K,k) := C (K,k) ( f g), for any f,g C (R 3,R), and the smooth function C (K,k) C (R 3,R) isgivenby C (K,k) (u):= 1 2 ut Ku + u T k.

3d quadratic integrable systems as Hamilton-Poisson systems The center of the Poisson algebra C (R 3,R) is generated by the Casimir invariant function C K C (R 3,R), C K (u)= 1 2 ut Ku + u T k.

3d quadratic integrable systems as Hamilton-Poisson systems Consequently, a quadratic Hamilton-Poisson system on (R 3,{, } (K,k) ), is generated by a smooth function H (A,a) C (R 3,R), given by H (A,a) (u):= 1 2 ut Au + u T a, where A Sym(3) is an arbitrary real symmetric matrix, and a R 3.

3d quadratic integrable systems as Hamilton-Poisson systems Hence, the associated Hamiltonian system is governed by the following differential equation: u =(Ku + k) (Au + a), u R 3. (8)

3d quadratic integrable systems as Hamilton-Poisson systems In the above settings, if there exists α,β R such that αa + βk is positive (or negative) definite, then the system (8) is affinely equivalent to Euler s equations of the free rigid body dynamics with linear controls, namely: u = u (Du + d), u R 3, (9) where D = diag(λ 1,λ 2,λ 3 ) is a real diagonal 3 3 matrix, and d R 3. In the case when d = 0, the equations (9) are exactly the Euler s equations of the free rigid body dynamics.

3d quadratic integrable systems as Hamilton-Poisson systems Note that in coordinates, the system (9) become: ẋ 1 =(λ 3 λ 2 )x 2 x 3 + d 3 x 2 d 2 x 3, ẋ 2 =(λ 1 λ 3 )x 1 x 3 d 3 x 1 + d 1 x 3, ẋ 3 =(λ 2 λ 1 )x 1 x 2 + d 2 x 1 d 1 x 2, where (d 1,d 2,d 3 ) are the coordinates of d.

3d quadratic integrable systems as Hamilton-Poisson systems If the matrices A and K commutes, then the system (8) is orthogonally equivalent to the dynamical system: v =(D K v + ˆk) (D A v + â), v R 3, (10) where the real diagonal 3 3 matrices D A, D K are given by D A = R T AR, D K = R T KR, wherer O(Id)=O(3,R) isa3 3 orthogonal matrix, and ˆk := det(r)r T k, â := det(r)r T a.

3d quadratic integrable systems as Hamilton-Poisson systems Using coordinates, the system (10) becomes: ẋ 1 =(K 2 A 3 K 3 A 2 )x 2 x 3 +(K 2 a 3 k 3 A 2 )x 2 +(k 2 A 3 K 3 a 2 )x 3 +k 2 a 3 k 3 a 2 ẋ 2 =(K 3 A 1 K 1 A 3 )x 1 x 3 +(k 3 A 1 K 1 a 3 )x 1 +(K 3 a 1 k 1 A 3 )x 3 +k 3 a 1 k 1 a 3 ẋ 3 =(K 1 A 2 K 2 A 1 )x 1 x 2 +(K 1 a 2 k 2 A 1 )x 1 +(k 1 A 2 K 2 a 1 )x 2 +k 1 a 2 k 2 a 1, (11) where D A = diag(a 1,A 2,A 3 ), D K = diag(k 1,K 2,K 3 ), â =(a 1,a 2,a 3 ), and ˆk =(k 1,k 2,k 3 ).

Bibliography R.M. Tudoran, The free rigid body dynamics: generalized versus classic, arxiv: 1102.2715 R.M. Tudoran, On a class of three-dimensional quadratic Hamiltonian systems, Appl. Math. Lett. 25(9)(2012), 1214 1216. R.M. Tudoran, A normal form of completely integrable systems, J. Geom. Phys. 62(5)(2012), 1167 1174.