Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Mechanics, Analysis and Geometry: 200 Years after Lagrange
Mechanics, Analysis and Geometry: 200 Years after Lagrange
Mechanics, Analysis and Geometry: 200 Years after Lagrange
Ebook920 pages

Mechanics, Analysis and Geometry: 200 Years after Lagrange

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Providing a logically balanced and authoritative account of the different branches and problems of mathematical physics that Lagrange studied and developed, this volume presents up-to-date developments in differential goemetry, dynamical systems, the calculus of variations, and celestial and analytical mechanics.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780444597373
Mechanics, Analysis and Geometry: 200 Years after Lagrange

Related to Mechanics, Analysis and Geometry

Physics For You

View More

Reviews for Mechanics, Analysis and Geometry

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Mechanics, Analysis and Geometry - Elsevier Science

    Francaviglia

    Part I

    DYNAMICAL SYSTEMS

    Outline

    Chapter 1: PERIODIC SOLUTIONS NEAR THE LAGRANGE EQUILIBRIUM POINTS IN THE RESTRICTED THREE-BODY PROBLEM, FOR MASS RATIOS NEAR ROUTH’S CRITICAL VALUE

    Chapter 2: LOWER BOUND ON THE DIMENSION OF THE ATTRACTOR FOR THE NAVIER-STOKES EQUATIONS IN SPACE DIMENSION 3

    Chapter 3: HOMOCLINIC CHAOS FOR RAY OPTICS IN A FIBER: 200 YEARS AFTER LAGRANGE

    Chapter 4: ON THE VORTEX–WAVE SYSTEM

    PERIODIC SOLUTIONS NEAR THE LAGRANGE EQUILIBRIUM POINTS IN THE RESTRICTED THREE-BODY PROBLEM, FOR MASS RATIOS NEAR ROUTH’S CRITICAL VALUE

    Gianfausto DELL’ANTONIO,     Dipartimento di Matematica, Università di Roma, 00100 Roma; Scuola Internazionale Superiore di Studi Avanzati (SISSA/ISAS), 34014 Trieste, Italy; CNR, Gruppo Nazionale Geometria e Fisica

    Publisher Summary

    This chapter focuses on periodic solutions near the Lagrange equilibrium points in the restricted three-body problem for mass ratios near Routh’s critical value. It reviews the basic lemma on commuting vector fields. The chapter further discusses the case of Hamiltonian vector fields and uses the theory of normal forms to show the relation between periodic solutions near equilibrium and critical points of suitable function F on the level sets of H2. The Floquet multipliers with the eigenvalues of the Hessian of F are further reviewed in the chapter. The chapter also presents an analytical description of the restricted three-body problem.

    1 INTRODUCTION

    We study periodic solutions for the flow of a vector field Y in Rα near an isolated critical point xo. We assume that the linear part is at (or near) resonance (all eigenvalues are imaginary and have a common divisor). We do not require that the linear part be diagonizable at resonance. Under smoothness assumptions on Y, for every T > 0 we find, through the solution of an algebraic equation, all periodic solutions of x = Y(x) which lie sufficiently close to xo and have periodic ≤ T. We apply this result to the study of periodic solutions for the restricted three-body system for mass values close to Routh’s critical values; we determine their position and (linear) stability.

    In ¹ we have characterized the points in a neighborhood of an equilibrium which are initial data for periodic solutions of an Hamiltonian system, provided the Hamiltonian is sufficiently regular, has a quadratic part H2 given by

    and the set of frequencies νi is at or near resonance.

    The (linearized) stability of such periodic solutions was also determined. A basic tool in 1 was a lemma on commuting vector fields. In this paper was extended the basic lemma to the non-semisimple case (i.e. the case when H2 cannot be written in diagonal form) and use this result to extend to the general case the characterization of small periodic orbits and the analysis of their stability.

    As an application, we determine the periodic orbits near the Lagrange equilibrium points for the restricted three-body problem when the mass ratio is close to Routh’s critical value. We also determine the stability of such orbits ².

    This paper is organized as follows:

    In Section 2 we extend to the non-semisimple case the proof of the basic lemma on commuting vector fields.

    In Section 3 we specialize to the case of Hamiltonian vector fields, and use the theory of normal forms to give the relation between periodic solutions near equilibrium and critical points of suitable function F on the level sets of H2. We also connect the Floquet multipliers with the eigenvalues of the Hessian of F.

    In Section 4 we apply these results to the restricted three-body problem.

    2 A LEMMA ON COMMUTING VECTOR FIELDS

    Consider in Rd the vector field

    (2.1)

    where G is superlinear in z at the origin, of class C¹ in z for |z|

    (2.2)

    The matrix A is antisymmetric with eigenvalues ±iυ1, …±iυd which satisfy

    (2.3)

    The matrix B in nilpotent, i.e. Bq = 0 for some integer q.

    Let Πo be the orthogonal projection onto KerB, the kernel of B. One can find b1>0 such that

    (2.4)

    We set

    (2.5)

    and denote by τ the largest positive number for which

    (2.6)

    We define N by

    (2.6) 1

    It follows from (2.3) that every solution φo(t,z) of the linear system

    (2.7)

    is periodic, and τ is a period (possibly not the minimal one).

    Let a and T be arbitrary but fixed positive numbers, with T≥τ. For ε≥0, denote by φε(t,z*) the solution of

    (2.8)

    For ε sufficiently small we want to characterize those z* for which φε(t,z*) is contained in the b all Ba = (z:|z|≤a) and is periodic with minimal period ≤T; we denote by Mε the subset of Rd for which these properties hold.

    Notice the following: if for given ε,z* there exists λ such that

    (2.9)

    then this relation holds also at φ0(t,z*),∀t since the fields Az and Yε(z) commute. Therefore

    (2.10)

    so that in particular φε(t,z*) is periodic with minimal period

    We now prove (Lemma 2.1) that if ε is sufficently small, z* belongs to Mε precisely if (2.10) holds.

    It is an easy consequence of regular perturbation theory that one can find positive constants ε0 and c1 (depending on T, a, on the frequencies νi and on the Lipshitz constant of G), such that, if ε≤ εo and z*∈ Mε, there exists an integer K(ε,z*) for which

    (2.11)

    If z*∈ Mε one has, form the integrated version of (2.8),

    (2.12)

    where

    for some new constant c2.

    Since by assumption ϕε(Tε(z*),z*) = z*, projecting (2.12) with |-Π0 and using (2.4) one derives, for z*ε Mε and ε≤ εo

    (2.13)

    and then, using the commutativity of B and A

    (2.14)

    Since G(z.ε) is superlinear at the origin, there is no loss of generality in assuming that |z*|≥a/2; indeed, scaling a→z/2 gives a new vector field of the form (2.1) with a smaller value for ε.

    It follows then that (2.13) holds (possibly with a larger constant c4) in anε-neighborhood of φo(t,z*) which contains the full orbit of φε(t,z*), i.e.

    One can now prove

    Lemma 2.1

    Let Yε(z) be given as in (2.1) and let T, a be fixed. Then one can find ε1>0 with the property that, if ε<ε1 and |z*|≤a, the function φε(t,z*) is periodic with minimal period Tε(z*)≤T precisely if (2.9) holds for some constant λ. If this is the case, then φε(t,z*) coincides with φo(t,z*) up to a time scale. /

    Proof

    We have already remarked that (2.9) is a sufficient condition, and that it implies the last statement in the lemma. We prove now that if ε is sufficiently small, (2.9) is also necessary.

    For fixed z*, ε define λ(ε,z*) as follows:

    (2.15)

    where K(ε,z*) is defined in (2.11)

    Then, in view of (2.11), |λ(ε,z*)|≤c5 for all z*∈ Mε, ε<ε0.

    Define ψε(t,z*)by

    (2.16)

    The ψε is periodic in t of period NTε(z*) since N(1+ελ)Tε(z*) = NKτ = Kτ and exp{-KτA} = I for every integer K.

    Differentiating (2.15) and using the fact that Bz and G(z,ε) commute with Az, one concludes that ψε(t,z*) is the solution of

    (2.17)

    where

    (2.18)

    Notice that in view of (2.14) the vector field Θ(z,ε) is uniformly bounded for z*∈ Mε and ε<ε0. It is also uniformly bounded on the solutions of (2.17) since A commutes with πo.

    Without loss of generality we can assume that z* is chosen so that

    Therefore

    (2.19)

    From (2.14) and the fact that A and B commute it follows then that (2.13) holds (with the larger constant c4) in an ε-neighborhood of z* which contains the full orbit of ψε(t,z*).

    Since (I-Πo)Bz=Bz, one has for all t

    (2.20)

    The function ψε is periodic in t of period NTε(z*) and therefore

    (2.21)

    We set L=NTε(z*); iterating (2.21) q-1 times, collecting all terms proprotional to (LB)S for some s and using (2.19), Bq = 0 and that G is Lipshitz-continuous, one obtains

    ! is invertible and therefore

    so that, if

    (2.22)

    one must have Θ=0, i.e. (2.9).

    This concludes the proof of Lemma 2.1. /

    Remark 2.2

    Condition (2.9) can be written in the form

    so that (2.9) is equivalent to the condition that z be a critical point for the projection of ε²AG(z,ε)+ABz onto TEc. Denote by Γ this vector field.

    The flow t→exp{A} provides a natural S¹ action on Γ and on the (compact) manifold Ec. A lower bound for the number of critical points of Γ on Ec (and therefore on the number of periodic solutions of (2.8) on Ec with period smaller than T) can be obtained by the use of Equivariant Morse Theory or of Cohomology Index methods ³. /

    Remark 2.3

    Consider the case in which Yε is a hamiltonian vector field. Denoting by J the standard symplectic map, one will have

    (2.23)

    where

    Condition (2.9) takes the form

    i.e. z is a critical point of ε²H"(z)+H′(z) on the surface Ec ={z:E2(z)=c). As before, there is a natural S¹ action, and one can use Equivariant Morse theory. In particular, if Ec is convex, there are at least d circles of critical points, and therefore at least d periodic solutions on Ec. It follows easily from the last part of Lemma 2.1 that these solutions form continuous families, parametrized by the energy.

    We determine now the stability of the periodic solutions found through Lemma 2.1. We follow closely the analysis given in ¹.

    Lemma 2.4

    The Floquet multipliers {ρi, i+1…d} for the periodic solution described in Lemma 2.1 satisfy

    (2.24)

    where μi are the eigenvalues of the matrix

    and N is defined in (2.6) 1; in particular, if T0(z*) = τ, then N=1 in (2.24). /

    Proof

    Let p(t) be a periodic solution with period Tε as described in Lemma 2.1:

    The Floquet multipliers are the eigenvalues of the map

    (2.25)

    where σ(t,ζ) is the solution of

    (2.26)

    Define χ(t,ζ) by

    (2.27)

    By construction one has

    and therefore

    (2.28)

    On the other hand

    (2.29)

    since G(z) and Bz commute with Az.

    The conclusions of Lemma 2.3 follow now immediately from (2.28) and (2.29). /

    A special important case is the one in which the vector field is Hamiltonian, so that through a symplectic transformation it can be written [4] in the form (2.23).

    One has then

    (2.30)

    Remark that the vectors ∇H2 and JdH2 are constant under the flow (2.30) and define χ as follows:

    Definition

    χ is the restriction of DJ[dH′ + ε²dH"] to the othogonal complement in R²d of the subspace generated by ∇H2 and JdH2.

    One has then:

    Corollary 2.5

    If the vector field has the form (2.20) and p(t) is a periodic solution of (2.4) for ε sufficiently small (as described in Lemma 2.1), then p(t) is elliptic precisely if all the eigenvalues of χ are purely imaginary, is hyperbolic if all the eigenvalues of χ have a non-zero real part and is unstable if at least one of the eigenvalues has positive real part.

    Remark 2.6

    It follows from (2.9) that λ(ε,z*) = ελo(ε,z*), where |λo(ε,z*)| is uniformly bounded for 0≤ ε≤ εo, |z*|≤ a. Therefore (2.9) is equivalent to

    (2.9)’

    One can then take ε² as parameter, and write Yε(z) as

    (2.1)’

    We shall do so in the following sections, and write λ for λo.

    3 APPLICATION TO HAMILTONIAN SYSTEMS

    We use the results of Section 2 to study periodic solutions near equilibrium for Hamiltonian systems, under the condition that the frequencies of the quadratic part are at or near resonance.

    It is proven in 4 (see also 5) that in a neighborhood of an isolated equlibrium if the Hamiltonian is of class Cs, s≥3, and the eigenvalues of the linear part of the Hamiltonian vcector field are ±iυ1,…±iυd (not counting multiplicities) one can choose canonical coordinates {zk, k= 1…d}, zk = {qk, Pk}, such that the equilibrium is the origin and in the z’s the Hamiltonian has the form

    (3.1)

    where B is nilpotent, A = diag{νk}, H* is of class Cs in a neighborhood of the origin, and infinitesimal at the origin of order m≥3.

    If the frequencies {νi} are near resonance, they can be written

    (3.2)

    where δ is a (small) parameter. The matrix B in (3.17) can be chosen so that it commutes with A0, where

    (3.3)

    Since we are interested in the behaviour near the origin, we introduce the relevant scale through the canonical transformation

    (3.4)

    where ε is another (small) parameter.

    In the following we shall often set δ=εα; the constant α will depend on H* and will be chosen so that the detuning will be of size comparable to that of the leading non-linear terms in the normal form.

    In the new variables Hamilton’s equation becomes

    (3.5)

    with

    (3.6)

    By construction, Q(z,ε) is of class Cs in z, uniformly for ε sufficiently small.

    We use now the theory of normal forms, and choose a new set of canonical variables in such a way that the leading term in the non-linear part of (3.5) commutes with A0z. This requires in general that Q be sufficiently smooth; the precise degree of differentiability depends on the {nk} in (3.3) and on the form of Q. For the sake of simplicity we shall assume that Q(z,ε) is of class C∞ in z, uniformly for ε sufficiently small; in the course of the proof it will be clear which is the minimum degree of differentiability required.

    We shall make use of the following Lemma ⁶, which we present in a form best suited for our analysis.

    Lemma 3.1.

    Let Hε(z) have the form

    (3.7)

    where K(z) = 1/2(z,A0z), Aμ(z) = 1/2(z,Aμz), B(z) = 1/2(z,Bz), the matrices A0, Aμ, B are given in (3.3), (3.6) and H′(z,ε) is of class C∞ in z, uniformly for ε sufficiently small, and infinitesimal in z of order at least three. For every integer m≥3 one can then find ε1>0 and a symplectic transformation Φε,m, asymptotic to the identity when ε→0, and such that if ε<ε1

    (3.8)

    Rm is of class C∞ in z and is uniformly bounded for |z|≤1 and ε<ε1. The Pn,ε are polynomials in z of order n+2 and satisfy {Pn,ε, K}P.B =0. Some of the Pn,ε could be identically zero. The function Hε,m(z) obtained from (3.8) by deleting the term Rm(z) is called normal form of to order m (relative to K(z)). /

    Notice that the hamiltonian vector field JdHε,m commutes by construction with JdK. We can therefore apply directly Lemma 2.1 and conclude that for ε sufficiently small the periodic solutions Ψε(t,z) of

    (3.9)

    with minimal period smaller than a fixed number T0 are in one-to-one correspondence with the orbits (under the flow of JdK) of the solutions of

    (3.10)

    We study now, still for ε sufficiently small, the periodic solutions φε(t,z) of

    (3.11)

    with minimal period smaller that T0.

    Since we are interested in small values of ε we want to make use of the Inverse Function Theorem. We make therefore an assumption on the Jacobian of the map z→Xm,ε,μ where X is defined in (3.10). This assumption is satisfied, for m large enough (depending on the frequencies in K(z)), if Σn=1…mPn,ε does not belong to a subspace of codimension at least one in the vector space of the polinomials of order ≤m+2 which are normal with respect to K.

    In this sense, condition (3.12) in the next Proposition is satisfied generically.

    Proposition 3.2

    Let the Hamiltonian Hε be as in (3.8). Define Xm,ε as in (3.10) and denote by Hess0(F) the restriction of the Hessian matrix of F to the orthogonal complement in R²d of the subspace generated by ∇K and by JdK. Denote by φ0(t,z) the periodic solution of dz/dt = JdK(z), φ0(0,z) = z. For every c1, c2, c3,>0 and every μ∈ Rd one can find ε1>0 such that the following statements hold true:

    I If ε<ε1 and z* is a solution of

    (3.12)

    which satisfies

    (3.13)

    Then one can find yε∈ R²d such that

    (b) φε(t,yε) is a periodic solution of (3.11) with period approximately equal to the minimal period of φ0(t,z*).

    (c) Δ(φε,φ0)

    The point yε is not defined uniquely by these conditions, however the periodic solution φε(t,yε) is uniquely determined by the condition Hε(yε)=Hε(z*).

    II Conversely, if ε<ε1 and yε∈ R²d is such that (3.13) is satisfied and φε(t,yε) is a periodic solution of (3.11) with minimal period Tε not larger than T0, there exists zε∈ R²d and λε∈ R such that

    a) Xm,ε,μ (z*)=0, so that Ψε(t,z*) is a periodic solution of (3.9) which differs from φ0(t,z*) only by a rescaling of time. The minimal period of this solution is approximately equal to Tε.

    b) |yε-zε|

    c) Δ(φε,φ0)<εm-S+1

    The -periodic solution Ψε(t,z*) is uniquely determined by the condition Hε(zε)=Hε(yε).

    Proof

    We shall only give a brief sketch of the proof, which is a simple application of the Inverse Function Theorem. Details can be found in ¹. To prove II, denote by yε(t,z) the solution of (3.9) starting at z. By regular perturbation theory,

    (3.14)

    The statement that φε(t,y) is a periodic solution of (3.11) of period Tε is equivalent to

    (3.15)

    Since Hε is constant along the flow of JdHε, it is sufficient to consider the restriction of (3.15) to the energy surface through y, more precisely to the orthogonal complement to ∇Hε at y. Moreover, since the Hamiltonian does not depend on time, it is sufficient to consider a Poincaré section at y, and we do this by further restricting (3.15) to the orthogonal complement to JdHε at y.

    (T′,z) is defined as in (3.15) but for the Hamiltonian Hε (and also the restriction is defined relative to JdHε and to ∇Hε). From (3.14), (3.11) one concludes through the Implicit Function Theorem that, if yε solves (3.15), one can find zε,T′ which are solution of Φε(T′,zε)=0 and have the properties described in part II of Lemma 3.2. Part II of Lemma 3.2 is then a consequence of Lemmas 2.1 and 3.1.

    The proof of part I follows along the same lines, interchanging the role of Hε and of Hε.

    The stability properties of the peridic solutions of (3.11) which are described in Lemma 3.2 follow immediately from Corollary 2.5 and from Krein’s theory of regular perturbation for the Floquet multipliers. One has in particular

    Corollary 3.3

    The solution φε(t,yε) described in Proposition 3.2 is elliptic if all the eigenvalues of D0JdHε(zε) are purely imaginary and have multiplicity one. It is unstable if at least one of these eigenvalues has non-zero real part and hyperbolic if all eigenvalues have non-zero real part. /

    4 APPLICATION TO THE RESTRICTED THREE-BODY PROBLEM

    We apply now the results of the preceding section to the study of the periodic solutions near an equilibrium for the restricted three-body problem.

    This problem can be defined as follows: two material points revolve in a circular orbit around their center of mass under the influence of their mutual gravitational attraction. We call this the primary system. A third material point - gravitationally attracted by the previous two but not influencing their motion - moves in the plane defined by the two revolving points. The restricted three-body problem is to describe the motion of the third material point.

    The history of the restricted three-body problem begins with Euler ⁷; the first two bodies (material points) were the sun and the earth, the third body was the moon.

    Euler introduced a synodic (rotating) coordinate system, and found three equilibrium points for the third body in this coordinate system (these are called colinear equilibria since they correspond to configurations in which the three points are aligned): these equilibrium points are unstable.

    Lagrange ⁸ found two more solutions, called equilateral since the three bodies are at the vertices of an equilateral triangle. For some ranges of the value of the masses of the primary system these equilibria are (linearly) stable. One can then expect to find near equilibrium families of closed orbits for the third body (still in the rotating coordinate system).

    Lagrange showed in 1772 the existence of such orbits for the restricted problem formed by the Sun-Jupiter system; the discovery of their physical realization, the Trojan group of asteroids, began in 1906 with the first-seen member of this group, 588 Achilles.

    Since then, many more Trojan asteroids have been discovered and their orbits have been computed; we refer to ⁹ for a theoretical analysis and detailed computations.

    Important contributions to the study of the restricted three-body problem were made among others by C. Jacobi and G. Hill ¹⁰: periodic orbits were studies by H. Poincaré and then particularly by G. Birkoff ¹¹ (they were the motivation for his study of ring transformations).

    With increased interest in space exploations, the problem of periodic solutions near the lagrange equilibrium points became important for other primary systems, in particular the Earth-Moon system; in these cases, the third body is a man-made satellite ¹².

    We refer to ¹³ for an extensive treatment of the subject, with numerical estimates.

    The analytic aspects of the three-body problem can be found e.g. in ¹⁴. We denote by μ, 1-μ the reduced masses of the two primaries. By symmetry, it is sufficient to consider the range 0<μ≤1/2. When μ is smaller than a critical value μ1 (called Routh’s critical value) the Lagrange triangular points are stable.

    There is an infinite set μ1> μ2>…>μn>. of values of the positive parameter μ for which there is a resonance, in the sense that the ratio of the frequencies of the linearized system is an integer.

    For μ≠μk and smaller that μ1 the existence of two families of periodic solutions in a neighbourhood Nμ of the equilibrium follows from Lyapunov’s center theorem ¹⁵; the distance from equilibrium can be used as parameter.

    The radius of N decreases to zero when μ→μk. For the critical values μk the linearized system has a resonance and Lyapunov’s method provides only the existence for k≠1 of a one-parameter family of (elliptic) periodic solutions with period approximately equal to the shorter of the two periods of the linearized system. When k≥4 the existence of a second (elliptic) one-parameter family of periodic solutions, with minimal period approximately equal to the longer period of the linearized system, was proved in ¹⁶ (where one can find a very extensive treatment, both analytical and numerical, of the periodic solutions near the equilibrium point L4).

    In fact one can use the results of ¹ to study these cases and also the cases k=2,3 and in general the behaviourof these periodic solutions of μ close to μk. In particular for μ=μ2 only the family with shorted period is present (this is a l:-2 resonance since the Hamiltonian is not of definite sign at the origin) whereas for μ very close to μ3 there are two families of periodic solutions.

    It should perhaps be noted that when μ has a value for which the ratio of the frequencies of the linearized system is a non-integer rational number one can have further families of periodic solutions near the equilibrium. These families, when they exist, have minimal period approximately equal to the least common multiple of the periods of the linearized system. Their existence and their behaviour for small changes of the parameter μ can be analysed with the method described in ¹.

    We shall not discuss these cases here, and we restrict our analysis to values of the parameter which are close to μ1.

    When μ>μ1, the equilibrium point becomes linearly unstable and the periodic solutions, if they exist, cannot be too close to the equilibrium. Their existence cannot be determined directly from regular bifurcation theory, since for μ=μ1 the quadratic part of H is in irreducible Jordan form; see however ¹⁷ and ¹⁸.

    Existence of periodic solutions when μ=μ1 and their behaviour when μ is close to this critical value have been studied by several authors ¹⁹ either by numerical investigation or approximate series expansion.

    A mathematical analysis of the problem has been carried out in ² (using a perturbation Lemma, adapted from²⁰ which is proved using the continuation method of Poincaré) and in ¹⁸.

    In this section we shall apply the results of Sections 2 and 3 to prove that there are precisely two families of periodic solutions with small period which vary continuously when the parameter μ is varied in a full neighborhood of the critical value μ1. We shall also prove that these solutions are elliptic.

    We give now an analytic description of the restricted three-body problem.

    In the moving reference frame we choose cartesian coordinates x,y in such a way that the coordinates of the two primary bodies are (-μ1,o) and (μ2,o), respectively, with μi=mi(m1+m2)−1. In this coordinate system, the equations of motions are

    (4.1)

    where f′=df/dt, Fx=∂F/∂x and

    (4.2)

    Units have been chosen in which both the distance between the two primaries and the angular velocity of the circular motion in the inertial system are equal to one. In these units, the Lagrange equilibrium points are

    (4.3)

    It is easy to see that the system (4.1) is Hamiltonian, with

    where Q1=x, Q2=y, and r1,r2 are given in (4.2).

    At the equilibrium L4 the coordinates P,Q take the following values:

    Remark 4.1

    It can be verified that the Pi’s are the components of the velocity of the third body in the inertial frame along the two cartesian axes we have chosen (and which rotate in the inertial frame with constant angular velocity =1)./

    We write μ for μ2; due to the symmetry of the problem we shall discuss only motion near L4 and restrict ourselves to the range 0<μ≤1/2. The characteristic equation for the linearized system is

    (4.4)

    which has the four solutions

    The critical points μ1, μ2… for the parameter μ are by definition those values for which λ+,+/λ+,- = k, k=1,2,…

    One has μk<μk if k>h and the equilibrium in (linearly) stable for 0<μ≤μ1, where μ1 (Routh’s critical value) is given by

    (4.5)

    For μ1 <μ≤1/2 the equilibrium is unstable.

    Near equilibrium we shall use canonical coordinates qi,pi which are obtained from Qi, Pi first translating so that the equilibrium corresponds to the origin and then scaling by a factor ε; the second transformation is also canonical, and the new Hamiltonian differs from the previous one by a factor ε−2.

    The new Hamiltonian H has then the form ²

    where N is C∞ uniformly in ε, and infinitesimal at the origin of order three.

    For μ=μ1 the linearized vector field has a 1:-1 resonance and is in irreducible Jordan form. One has, as in (2.23)

    (4.6)

    with

    and

    Notice that K generates rotations (of the same angle) in the two planes q1=q2=0, p1=p2=0.

    Using the theory of normal forms one can then introduce a new set of canonical variables, still denote by qi,pi, so that Hμ has the form

    (4.7)

    where Aμ is given by

    (4.8)

    with

    (4.9)

    so that a>0 corresponds to μ<μ1.

    In (4.7) N1 is a polynomial of order two in the four quadratic invariants Fk

    (there are no odd-degree invariants).

    .

    To derive (4.7) one can use the fact ¹⁸ that, for μ=μ1, the Hamiltonian can be written in the form (4.7) with α=0. If B is the quadratic part of Hμ-Hμ1, one can solve for G the equation

    One can then use F3+αG as generating function for a symplectic transformation; in the new variables the quadratic part of Hμ has the form given in (4.7). To obtain (4.7) one performs two more symplectic transformations, with generating functions F3+εG1, F3+ε²G2 where Gi, i = 1,2 are suitable homogeneous polynomial of order i+2.

    Notice that the invariants Fi satisfy the algebraic relations

    so that any additional term proportional to αF3 in (4.7) can be cancelled, up to terms of order α², through a linear symplectic transformation with generator F3+cαF2.

    One can also verify that

    (4.10)

    It follows then from (2.14) that there is a constant c such that in a neighborhood of the periodic solutions near equilibrium the following inequality is satisfied:

    (4.11)

    One can therefore consider instead of (4.8) the Hamiltonian

    (4.12)

    provided one considers only the periodic solutions which satisfy

    Enjoying the preview?
    Page 1 of 1