spin-orbit interaction and chaos in celestial mechanics

108
Spin-Orbit Interaction and Chaos in Celestial Mechanics Manuel Maria Murteira Barreira da Cruz Thesis to obtain the Master of Science Degree in Engineering Physics Supervisor(s): Prof. Doutor Rui Manuel Agostinho Dilão Examination Committee Chairperson: Prof a ¯ . Doutora Maria Joana Patrício Gonçalves de Sá Supervisor: Prof. Doutor Rui Manuel Agostinho Dilão Member of the Committee: Prof. Doutor Paulo Jorge Soares Gil May 2017

Upload: others

Post on 16-Oct-2021

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Spin-Orbit Interaction and Chaos in Celestial Mechanics

Manuel Maria Murteira Barreira da Cruz

Thesis to obtain the Master of Science Degree in

Engineering Physics

Supervisor(s): Prof. Doutor Rui Manuel Agostinho Dilão

Examination Committee

Chairperson: Profa¯ . Doutora Maria Joana Patrício Gonçalves de SáSupervisor: Prof. Doutor Rui Manuel Agostinho DilãoMember of the Committee: Prof. Doutor Paulo Jorge Soares Gil

May 2017

Page 2: Spin-Orbit Interaction and Chaos in Celestial Mechanics

ii

Page 3: Spin-Orbit Interaction and Chaos in Celestial Mechanics

In memory of my grandfather...

iii

Page 4: Spin-Orbit Interaction and Chaos in Celestial Mechanics

iv

Page 5: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Acknowledgements

I would like to start by thanking my supervisor, Professor Rui Dilao, for his guidance, valuable advices

and support throughout this work. He has been responsible for introducing me to the world of dynamical

systems and celestial mechanics research in particular, and for broadening my horizons in physics and

mathematics in general, and I am very grateful for the opportunity of working with him.

Finally, I wish to thank my friends who accompanied me during this journey at Instituto Superior

Tecnico, and my family, especially my grandmother, my mother and my two grandaunts, for giving me

the means to complete this course, and for their continued support and encouragement. A special word

to my grandfather, who passed away shortly after I entered the course, and without whom I couldn’t have

reached this far.

v

Page 6: Spin-Orbit Interaction and Chaos in Celestial Mechanics

vi

Page 7: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Resumo

No contexto da interaccao spin-orbita em Mecanica Celeste podem ocorrer sincronizacoes ou res-

sonancias (instabilidades) entre dois movimentos periodicos.

Nesta dissertacao encara-se o problema da interaccao spin-orbita considerando um sistema con-

stituıdo por um corpo, que se assume ter a forma de um haltere, e uma massa pontual, que interagem

atraves da forca gravitacional. O modelo do haltere e util, pois este constitui o exemplo mais simples

de um corpo rıgido. A este sistema e dado o nome de Haltere Kepleriano. Sao obtidas as equacoes

do movimento do Haltere Kepleriano e estudadas as suas solucoes estacionarias, bem como a sua

estabilidade. Mostra-se que todas as solucoes estacionarias do Haltere Kepleriano sao instaveis, e que

o sistema e estruturalmente instavel no caso de massas iguais do haltere.

Depois de analisado o Haltere Kepleriano, o estudo da dinamica do haltere e incorporado num

contexto do Problema Restrito dos Tres Corpos. Derivam-se as equacoes do movimento do Problema

Restrito dos Tres Corpos com satelite em forma de haltere e analisam-se dois casos particulares — o

Problema Planar Restrito dos Tres Corpos e o Problema de Sitnikov — sendo determinadas condicoes

necessarias a sua ocorrencia. Com base nesta analise, sao encontradas as solucoes estacionarias do

sistema. Estas sao os analogos aos pontos de Lagrange do Problema Circular Restrito dos Tres Corpos

convencional.

Por ultimo, e realizada uma simulacao numerica do caso em que o haltere esta confinado a mover-

se e a rodar no plano de Lagrange, sendo encontradas evidencias de caos nalgumas das solucoes do

problema.

Palavras-chave: Haltere Kepleriano, Interaccao Spin-Orbita, Sincronizacao, Ressonancia,

Caos, Problema Restrito dos Tres Corpos

vii

Page 8: Spin-Orbit Interaction and Chaos in Celestial Mechanics

viii

Page 9: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Abstract

A synchronisation or a resonance (instability) between two periodic motions may occur in the context of

spin-orbit interaction in Celestial Mechanics.

We approach the study of the spin-orbit interaction by considering a dumbbell-shaped body, which

has the minimal features of a rigid body, moving in the gravitational field of a point mass, a system that

we call the Keplerian Dumbbell system. We derive the equations of motion of the Keplerian Dumbbell

system and analyse the steady states and their stability. We show that all the steady states of the

Keplerian Dumbbell system are Lyapunov unstable. For the case where the two masses of the dumbbell

are equal, the Keplerian Dumbbell system is structurally unstable.

Following up our study of the Keplerian Dumbbell, we incorporate the analysis of the dumbbell dy-

namics into the framework of the Restricted Three-Body Problem. We derive the equations of motion

of the Restricted Three-Body Problem with dumbbell satellite and study two special cases — the Planar

Circular Restricted Three-Body Problem and the Sitnikov Problem. Necessary conditions for these spe-

cial cases to occur are determined, and, based on that, the steady states of the system are found. These

steady states are the direct analogues of the Lagrangian points in the conventional Circular Restricted

Three-Body Problem.

Finally, we do a numerical analysis of the case in which the dumbbell is constrained to move and

rotate in the Lagrange plane. Evidences of chaotic dynamics are found in several solutions to this

problem.

Keywords: Keplerian Dumbbell, Spin-Orbit Interaction, Synchronisation, Resonance, Chaos,

Restricted Three-Body Problem

ix

Page 10: Spin-Orbit Interaction and Chaos in Celestial Mechanics

x

Page 11: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Contents

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

Resumo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv

List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

List of Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxiii

1 Introduction 1

1.1 The Spin-Orbit Effect in the Solar System . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1.1 Tidal Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.1.2 An Averaged Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.1.3 Capture into Synchronisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.1.4 Surface of Section and Chaos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Stationary Solutions of the Keplerian Dumbbell System 11

2.1 Equations of motion of the Keplerian Dumbbell system . . . . . . . . . . . . . . . . . . . . 12

2.2 Equations of motion in dimensionless form . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.3 Steady states and stability analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.3.1 Steady state 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.3.2 Steady state 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3.3 Steady state 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.3.4 Steady states 4 to 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.4 Numerical analysis of the non-stability of some steady state orbits . . . . . . . . . . . . . 24

3 Restricted Three-Body Problem with Dumbbell Satellite 27

3.1 The Restricted Three-Body Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.2 Equations of motion of the dumbbell satellite . . . . . . . . . . . . . . . . . . . . . . . . . 31

3.3 Equations of motion in the synodic reference frame, in dimensionless form . . . . . . . . 33

3.4 Steady states and other solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.4.1 PCR3BP with dumbbell satellite . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.4.2 Sitnikov Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

xi

Page 12: Spin-Orbit Interaction and Chaos in Celestial Mechanics

3.5 Numerical analysis of the case under category 4 of the PCR3BP with dumbbell satellite . 63

4 Conclusions 69

Bibliography 71

A Fixed Points and Steady States of Differential Equations 75

A.1 Differential equations as dynamical systems . . . . . . . . . . . . . . . . . . . . . . . . . . 75

A.2 Stability of the fixed points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

B Particular Solutions of the General Three-Body Problem 81

B.1 Eulerian solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

B.2 Lagrangian solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

B.3 Other periodic solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

xii

Page 13: Spin-Orbit Interaction and Chaos in Celestial Mechanics

List of Tables

1.1 Solar System data, including the rotation states of some of the major, natural satellites. . 2

xiii

Page 14: Spin-Orbit Interaction and Chaos in Celestial Mechanics

xiv

Page 15: Spin-Orbit Interaction and Chaos in Celestial Mechanics

List of Figures

1.1 Inertial (a) and rotating (b) reference frames used in the calculation of the asymptotic spin

rate of a satellite in the presence of tidal friction. . . . . . . . . . . . . . . . . . . . . . . . 4

1.2 The coordinates used in the averaging theory of spin-orbit interaction. . . . . . . . . . . . 5

1.3 A Surface of Section (SOS) for α = 0.2 and e = 0.1 and ten different trajectories in(θ, θ

)phase space. Image taken from [6]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.4 A SOS for α = 0.89 and e = 0.1, values appropriate for Hyperion. Image taken from [16]. . 9

2.1 Reference frames and coordinates used in the study of the Keplerian Dumbbell (Keplerian

Dumbbell (KD)) system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2 Steady state 1 of the KD system, represented in configuration space (Eulerian solutions,

see Appendix B). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3 Steady states 2 (a) and 3 (b) of the KD system in configuration space. . . . . . . . . . . . 19

2.4 Steady state 6 of the KD system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.5 Eigenvalues λ1 and λ2 of the reduced matrix Mss, (2.37), as a function of the steady state

orbital radius u0 of the dumbbell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.1 Motion of the primariesm1,m2 in the inertial reference frame S centred at their barycentre.

An eccentricity of 0.5, a mass ratio m2/m1 of 0.2, and a longitude of the pericentre of π/6

were used. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.2 Reference frames and coordinates used in the study of the Restricted Three-Body Prob-

lem (Restricted Three-Body Problem (R3BP)) with dumbbell satellite. . . . . . . . . . . . . 30

3.3 Inertial S and synodicR reference frames used in the analysis of the R3BP with dumbbell

satellite. The reference frame R corotates with the masses m1 and m2. . . . . . . . . . . 34

3.4 Possible configuration of the dumbbell and the primaries in the Planar Circular Restricted

Three-Body Problem (PCR3BP) with dumbbell satellite. . . . . . . . . . . . . . . . . . . . 44

3.5 Two steady states of the PCR3BP with dumbbell satellite, for η = 0.2 and ε = 0.15,

represented in the inertial S and synodic R frames at θ = 3π/5 ((a), (c)) and at θ = 5π/4

((b), (d)), for which the dumbbell lies in the Lagrange plane (ψ = π/2) and maintains a

right angle to the line that joins the primaries (θ − φ = π/2). . . . . . . . . . . . . . . . . . 47

xv

Page 16: Spin-Orbit Interaction and Chaos in Celestial Mechanics

3.6 Steady states (3.53) of the PCR3BP with dumbbell satellite represented in the inertial S

((a), (b)) and synodic R ((c), (d)) frames at θ = 5π/6, for ε = 0.15. (a), (c) The dumbbell

lies in the Lagrange plane (ψ = π/2) and maintains a right angle to the line that joins the

primaries (θ − φ = π/2); (b), (d) the dumbbell is lined up with the W axis (ψ = 0). . . . . . 49

3.7 Steady states (3.58) of the PCR3BP with dumbbell satellite, for ε = 0.15, in which the

dumbbell lies in the Lagrange plane (ψ = π/2) and is aligned along the direction of the

segment line that joins the primaries (θ = φ), represented in the inertial S ((a), (b)) and

synodic R ((c), (d)) reference frames. (a), (c) v = 0: the dumbbell and primaries are

depicted at θ = 4π/3; (b), (d) v(ε = 0.15, ψ = π/2) ≈ 0.867: the dumbbell and primaries

are depicted at θ = 7π/4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

3.8 Steady state (3.66) of the PCR3BP with dumbbell satellite, for η = 0.2 and ε = 0.15,

represented in the inertial S (a) and synodic R (b) reference frames at θ = 5π/3. The

dumbbell is aligned perpendicularly to the Lagrange plane (ψ = 0). . . . . . . . . . . . . . 54

3.9 Schematic representation of a configuration of the dumbbell and the primaries in the Sit-

nikov Problem (SP), in the inertial frame S. . . . . . . . . . . . . . . . . . . . . . . . . . . 58

3.10 Trajectory of the centre of mass of the dumbbell in inertial space (a) and time evolution

of the angular difference θ − φ (b) of a solution of (3.70) which starts close to one of the

fixed points (3.51). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

3.11 Trajectory of the centre of mass of the dumbbell in inertial space (b), its initial phase (a),

ending phase (c), and time evolution of the angular difference θ − φ (d) of a solution of

(3.70) which starts close to another of the fixed points (3.51). . . . . . . . . . . . . . . . . 65

3.12 Ejection of the dumbbell from the system (c) and initial phase of its trajectory (a) for a

solution of (3.70) which starts close to one of the fixed points (3.51). . . . . . . . . . . . . 66

3.13 Trajectory of the centre of mass of the dumbbell in inertial space (b), its initial phase (a),

ending phase (c), and time evolution of the angular difference θ − φ (d) of a solution of

(3.70) which does not start close to any fixed point. . . . . . . . . . . . . . . . . . . . . . . 67

A.1 Lyapunov stable fixed point (a). Asymptotically stable fixed point (b). . . . . . . . . . . . . 77

A.2 Graph of a Lyapunov function (a). Geometrical meaning of the Lyapunov theorem (b). . . 78

A.3 Orbits of a nonlinear dynamical system in phase space (left), and corresponding approxi-

mate linear system (right), in a neighbourhood of an hyperbolic fixed point. . . . . . . . . 79

B.1 Eulerian solutions of the General Three-Body Problem. (a) unequal masses; (b)–(c) all

masses equal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

B.2 Lagrangian solutions of the General Three-Body Problem for the case of equal masses.

(a) the bodies move in identical ellipses; (b) the bodies trace the same circular orbit. . . . 82

xvi

Page 17: Spin-Orbit Interaction and Chaos in Celestial Mechanics

List of Symbols

Greek Symbols

α (Ch. 1) Asphericity parameter.

α,β,γ (Chs. 2, 3) Unit vectors defining the principal

axes of inertia of the dumbbell.

αc Critical value of the asphericity parameter,

above which chaos is expected.

αH Asphericity parameter of Hyperion.

γ (Ch. 1) Angle between the long axis of the

satellite and the satellite-planet centre line at

pericentre.

γ0 Equilibrium value of γ.

δ (Ch. 1) Phase lag of the tidal bulge. (Ch. 2, 3)

Dimensionless parameter measuring the rela-

tive weight of the masses of the dumbbell.

∆E Energy dissipated over one cycle of tidal work-

ing.

ε Lag angle of the tidal bulge.

ε Dimensionless parameter measuring the length

of the dumbbell.

η (Ch. 3) Dimensionless parameter measuring

the relative weight of the masses of the pri-

maries.

η (Ch. 1) Satellite’s spin rate in a frame centred

on and rotating with the satellite’s mean motion.

θ (Ch. 1) Angle between the long axis of the

satellite and the major axis of the satellite’s or-

bit. (Ch. 2) Angular coordinate of the orbit of

the centre of mass of the dumbbell. (Ch. 3) An-

gular coordinate of the Keplerian orbits of the

primaries.

xvii

Page 18: Spin-Orbit Interaction and Chaos in Celestial Mechanics

θ (Ch. 1) Satellite’s spin rate. (Ch. 2) Angular ve-

locity of the centre of mass of the dumbbell.

(Ch. 3) Angular velocity of the primaries about

each other and their barycentre.

θ∗ Angular velocity for the circular orbit of the Ke-

pler problem.

λ (Ch. 2) Positive eigenvalues of the Hessian ma-

trix Ms (steady states 4–6). (Ch. 3) Semi-major

axis of the orbit of one of the primaries about

the other normalised to the radius of the circu-

lar orbit of the Kepler problem.

λ1, λ2 (Ch. 2) Eigenvalues of the reduced Hessian

matrix Mss (steady states 4–6).

µ (Ch. 2) Dimensionless parameter measuring

the relation between the masses of the dumb-

bell and of the primary. (Ch. 3) The product of

the gravitational constant G and the sum of the

masses of the primaries, m1 +m2.

ν (Ch. 3) Dimensionless parameter measuring

the relation between the masses of the dumb-

bell and the masses of the two primaries.

ξ (Ch. 2) Inverse of the steady state orbital radius

of the dumbbell, u0 (steady states 4–6). (Ch. 3)

Longitude of the pericentre.

ρ (Ch. 3) Dimensionless distance between the

primaries.

ρij (Ch. 2, 3) Dimensionless vector pointing from

mass i to mass j.

σ (Ch. 2) Normalisation frequency.

τ Dimensionless time.

ϕ Angle between the radius vector from the pri-

mary to the satellite and the line joining the

satellite to the empty focus of its elliptical orbit.

φ (Chs. 2, 3) Azimuthal angle on the sphere of

rotation of the dumbbell.

χ (Ch. 2) Canonical variable (difference between

the angles θ and φ).

xviii

Page 19: Spin-Orbit Interaction and Chaos in Celestial Mechanics

ψ (Ch. 1) Angle between the long axis of the

satellite and the satellite-planet centre line.

(Chs. 2, 3) Polar angle on the sphere of rota-

tion of the dumbbell.

ω (Ch. 2) Steady state angular velocity of the cen-

tre of mass of the dumbbell.

ω (Chs. 2, 3) Rotational angular velocity of the

dumbbell.

ω0 Libration frequency (pendulum equation).

Latin Symbols

a Semi-major axis of the Keplerian elliptical orbit.

A,B,C (Ch. 1) Satellite’s principal moments of inertia.

c (Ch. 2) Constant (steady state 1).

C Centre of mass of m1,m2,m3 (Ch. 2) or m1,m2

(Ch. 3).

C ′ Centre of mass of the dumbbell.

d day (unit of time).

D Magnitude of the tidal torque.

e Eccentricity.

e1, e2, e3 Unit vectors of the coordinate axes of the refer-

ence frames S and S ′.

E Satellite’s total energy.

E0 Peak energy stored in the body’s tidal distortion

over one cycle of tidal working.

f (Chs. 1, 3) True anomaly.

f, h,m Functions.

g (Ch. 1) Guiding centre. (Ch. 2) Energy scale of

the Keplerian Dumbbell (KD) system.

G Gravitational constant.

H (Ch. 2) Hamiltonian of the KD system. (Ch. 3)

Hamiltonian of the dumbbell.

H∗,Heff Effective Hamiltonians.

H(p, e) Power series in the eccentricity.

Iα, Iβ, Iγ Principal moments of inertia of the dumbbell.

k2 Love number of degree 2.

` Dumbbell’s length.

xix

Page 20: Spin-Orbit Interaction and Chaos in Celestial Mechanics

`2, `3, `4 (Ch. 2) `2, `3: distances of the masses m2 and

m3 to the centre of mass C ′ of the dumbbell,

respectively. (Ch. 3) `3, `4: distances of the

masses m3 and m4 to the centre of mass C ′

of the dumbbell, respectively.

L, L (Ch. 2) Total angular momentum of the KD sys-

tem.

Lz (Ch. 2) Third component of the total angular

momentum of the KD system. (Ch. 3) Angular

momentum integral of the Kepler problem.

L1+2 (Ch. 3) Total orbital angular momentum of the

primaries.

L,L (Ch. 2) Lagrangian of the KD system. (Ch. 3)

L: Lagrangian of the Four-Body Problem.

L1+2 Lagrangian of the Kepler problem.

Ldumbb.,Ldumbb. (Ch. 3) Lagrangian of the dumbbell.

mp Mass of the planet (primary).

ms Mass of the satellite.

m1,m2,m3,m4 (Ch. 2) m1: mass of the primary; m2,m3:

masses of the dumbbell. (Ch. 3) m1,m2:

masses of the primaries; m3,m4: masses of

the dumbbell.

M Mean anomaly.

Ms Hessian matrix.

Mss Reduced Hessian matrix.

n Average orbital angular velocity (mean motion).

Ns Tidal torque.

p Ratio of the satellite’s spin rate to its mean mo-

tion (rational number).

px Canonical momenta.

Q Specific dissipation function.

Qp Planet’s specific dissipation function.

Qs Specific dissipation function of the satellite.

r Instantaneous radius of the Keplerian orbit.

r0 (Ch. 2) Characteristic length of the KD system

(normalisation distance).

r∗ Radius of the circular orbit of the Kepler prob-

lem.

xx

Page 21: Spin-Orbit Interaction and Chaos in Celestial Mechanics

r, θ, z (Ch. 2) Cylindrical coordinates of the centre of

mass of the dumbbell, referred to mass m1, in

the inertial reference frame S.

rij (Ch. 2, 3) Vector pointing from mass i to mass

j.

Rs Satellite radius.

R (Ch. 3) Synodic reference frame.

S Inertial reference frame centred at the centre of

mass of m1,m2,m3 (Ch. 2) or m1,m2 (Ch. 3).

S ′ Reference frame centred at the centre of mass

of the dumbbell, with coordinate axes parallel

to those of S.

t Time.

T (Ch. 1) Orbital period. (Ch. 2) Total kinetic en-

ergy of the KD system. (Ch. 3) Kinetic energy

of the dumbbell.

Trot Rotation period of Nereid.

u0 (Ch. 2) Steady state orbital radius of the dumb-

bell.

u, v, w (Ch. 3) Dimensionless, cartesian coordinates

of the centre of mass of the dumbbell in the

synodic reference frame R.

u, θ, v (Ch. 2) Dimensionless, cylindrical coordinates

of the centre of mass of the dumbbell, referred

to mass m1, in the inertial reference frame S.

U (Ch. 2) Total potential energy of the KD system.

(Ch. 3) Potential energy of the dumbbell.

U, V,W (Ch. 3) Dimensionless, cartesian coordinates

of the centre of mass of the dumbbell in the in-

ertial reference frame S.

V (Ch. 3) Effective potential.

x, y, z Cartesian coordinates in the reference frames

S,S ′ (Ch. 2) and in the synodic reference frame

R (Ch. 3).

X,Y, Z Cartesian coordinates in the reference frames

S,S ′ (Ch. 3).

y year (unit of time).

Superscripts

T Transpose.

xxi

Page 22: Spin-Orbit Interaction and Chaos in Celestial Mechanics

xxii

Page 23: Spin-Orbit Interaction and Chaos in Celestial Mechanics

List of Acronyms

CR3BP Circular Restricted Three-Body Problem

KD Keplerian Dumbbell

PCR3BP Planar Circular Restricted Three-Body Problem

R3BP Restricted Three-Body Problem

SOS Surface of Section

SP Sitnikov Problem

xxiii

Page 24: Spin-Orbit Interaction and Chaos in Celestial Mechanics

xxiv

Page 25: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Chapter 1

Introduction

Nix is a small satellite that orbits the Pluto-Charon system. It was discovered in 2005 and confirmed in

2006, by the Hubble Space Telescope, together with the other small moon Hydra [1]. Analysis of the

variations in the light reflected by these moons shows unpredictable changes in their brightness. The

spin axis of both moons tumbles chaotically, likely due to the time-varying, asymmetric gravitational field

created by Pluto and Charon, as these orbit their common barycentre, and this effect is only enhanced

by the non-spherical shape of the moons. Showalter and Hamilton [2] have made dynamical simulations

of the rotational period and orientation of Nix versus time which support this hypothesis. Their results

demonstrate that Nix’s orientation is fundamentally unpredictable and that, most of the time, its instanta-

neous rotational period is completely unrelated to its orbital period, although, for brief periods, the small

moon’s long axis tends to oscillate about the direction towards the barycentre of the Pluto-Charon binary

system [2].

The main goal of this work is to characterise the spin-orbit coupling mechanism responsible for

the observed chaotic behaviour of Nix and to describe the synchronisation and the transition to chaos

of Nix’s rotation, in the framework of the Restricted Three-Body Problem (R3BP), Szebehely [3]. We

shall start, though, by analysing the simpler problem which we call the Keplerian Dumbbell (KD), which

concerns the spin-orbit interaction of a dumbbell satellite moving around a primary planet in an elliptic

Keplerian orbit. In the next section, we recall and review some of the main results and procedures of the

theory of spin-orbit interaction.

1.1 The Spin-Orbit Effect in the Solar System

The so-called spin-orbit interaction, or coupling, in Celestial Mechanics refers to the dependence of the

rotational period of a satellite on its orbital period around a primary planet or star, and it manifests itself

in the form of a synchronisation of two periodic motions, [4, 5]. There are commensurabilities between

various types of frequencies or periods; for instance, the orbit-orbit coupling involves the orbital periods

of two or more bodies. However, our main focus is on the synchronisation and/or resonance between

the rotational and the orbital periods of a single body. The most obvious example of these is the Moon,

1

Page 26: Spin-Orbit Interaction and Chaos in Celestial Mechanics

whose orbital period is exactly equal to its rotational period, which leads to the well-known fact that it

keeps the same face towards the Earth at all times: we say that the Moon is in a 1:1 or synchronous

state. Another example of a synchronous 1:1 rotation of particular interest is that of Charon, the major

moon of the dwarf planet Pluto. In this case, not only are the rotational and orbital periods of Charon

equal, they are also equal to the period of rotation of Pluto, which results in the satellite being seen

always in the same position in the sky from the planet.1 In fact, though, most of the major natural

satellites in the solar system are rotating in a 1:1 synchronous state, as shown in Table 1.1. The notable

exceptions to this spin-orbit state among the satellites are Hyperion and Nereid; the rotation period of

the former is chaotic [6, 7], while the rotation period of the latter seems to be near a 750:1 state [8].

Nereid is a mystery, and its rotation appears to be rather irregular [9]. Mercury, on the other hand, was

found to be in a 3:2 synchronous spin-orbit state as it revolves around the Sun, thereby completing three

rotations on its axis while making two revolutions around the Sun [10].

Planet Satellite e T (d) Rotation StateMercury 0.206 87.97 3:2Earth Moon 0.054900 27.321661 1:1Mars Phobos 0.0151 0.318910 1:1

Deimos 0.00033 1.262441 1:1Jupiter Io 0.0041 1.769138 1:1

Europa 0.0101 3.551810 1:1Ganymede 0.0015 7.154553 1:1Callisto 0.007 16.689018 1:1

Saturn Epimetheus 0.009 0.694590 1:1Enceladus 0.0045 1.370218 1:1Titan 0.0292 15.945421 1:1Hyperion 0.1042 21.276609 chaotic

Uranus Miranda 0.0027 1.413 1:1Ariel 0.0034 2.520 1:1Umbriel 0.0050 4.144 1:1Oberon 0.0008 13.463 1:1

Neptune Proteus 0.000 1.122315 1:1Triton 0.0004 -5.876854 1:1Nereid 0.7512 360.13619 Trot = 0.48 d†

Pluto Charon 0.0076 6.387223 1:1

Table 1.1: Orbital period (T ) and synchronisation state of some of the major, natural satellites in thesolar system. We also show the eccentricity (e) of the orbit of the satellites. Data taken from [4]. † Thevalue of the rotation period of Nereid Trot has been taken from [8].

1.1.1 Tidal Theory

A body raises a tide on another because the gravitational force varies across bodies, and a tidal bulge is

created in a position directly under the perturber since no body is truly rigid. As is known from the theory

of the tides, tidal oscillations are dissipative processes, because there is friction between and inside the

various layers of a planet or satellite, and so, energy is lost and a phase lag δ is generated in the tidal

response of the planet or satellite. When the rotational angular velocity or spin rate of a satellite differs

from its instantaneous orbital angular velocity or the rate of change of its true anomaly2f , a tidal torque

is created which acts to either brake or spin up the satellite, according to whether its tidal bulge is carried1This is called a state of mutual “tidal-locking” [4].2In celestial mechanics, the true anomaly f is an angular coordinate that determines the position of a body moving along a

Keplerian ellipse. It is defined as the angle between the position of the body and the direction to the pericentre, as seen from thefocus of the ellipse which is filled by the primary body [4].

2

Page 27: Spin-Orbit Interaction and Chaos in Celestial Mechanics

ahead of (positive phase shift) or lags behind (negative phase shift) the direction towards the perturber,

respectively [11]. As Murray and Dermott [4] pointed out, in order that a zero mean tidal torque acts

over one orbital period and equilibrium is reached, the satellite has to spin faster than its average orbital

angular velocity or mean motion n, which suggests that the 1:1 spin-orbit state is not stable. This was

first demonstrated by Peale and Gold [12], which showed that certain models of tidal friction could make

an axially symmetric planet in an eccentric orbit approach an asymptotic spin rate which is faster than its

mean motion, the value of which depends on the variation of the planet’s specific dissipation function3

Qp with the frequency and amplitude of the tides. Consequently, other torques must be present if we

are to have synchronous rotation. These torques are related to permanent deviations from sphericity, or

quadrupole moments, which require the bodies to be partially solid in order to be able to sustain such

deformations [13]. Goldreich [14] calculated the asymptotic spin rate for a satellite in a low eccentricity

orbit in the absence of permanent bulges. In terms of a tidal drag model due to MacDonald [11] that

takes a constant (frequency-independent) lag angle ε = δ/2 for the tidal bulge, the tidal torque he used

was ([4])

Ns = −D(ar

)6sgn (η − ϕ) , (1.1)

where

D =3

2

k2Qs

n4

GR5s (1.2)

is a positive constant, a is the semi-major axis of the orbit of the satellite, r is the instantaneous radius

of this orbit, G is the gravitational constant, G = 6.67408 × 10−11 m3 kg−1 s−2 in SI units, Rs, Qs and

k2 are the radius, specific dissipation function and Love number4 of the satellite, respectively, the time

derivative η is the spin rate of the satellite in a frame that is centred on the satellite and rotates with

its mean motion n (Figure 1.1b), ϕ is the rate of change of the angle between the radius vector from

the primary to the satellite and the line joining the satellite to the empty focus of its elliptical orbit (see

Figures 1.1a, 1.1b), and sgn(·) is the sign function.

The result obtained was η = 192 e

2n [14, 4]. For the Moon, this would amount to about 3% faster

rotation than the synchronous rate, and we would see both sides of the natural satellite over a period of

2.6 y [4].

1.1.2 An Averaged Equation of Motion

In the following we will always assume, unless otherwise stated, that the axis of rotation or spin axis of

the satellite, modelled as an ellipsoid, coincides with its axis of largest principal moment of inertia (the

shorter axis of the satellite) and is normal to the orbital plane of its elliptical path.5 We’ll also neglect any

secular perturbations6, and so this ellipse and spin orientation may be regarded as fixed. Under these

3The specific dissipation function Q of a body is defined by Q = 2πE0/∆E, where ∆E is the energy dissipated and E0 is thepeak energy stored in the body’s tidal distortion over one cycle of tidal working [11].

4The Love number of degree 2 k2 is a dimensionless parameter describing the ratio of the gravitational potential of a tidally-deformed body at a point to the perturbing potential at that point. It measures the susceptibility of a body’s shape to change inresponse to a tidal potential and is zero for a rigid body [15].

5Tidal torques tend to drive the obliquity, the angle between the spin axis and the orbit normal, to zero, thus it is natural to studythis problem in which the spin axis of the satellite is normal to the orbital plane [16].

6A secular perturbation is a non periodic perturbation over the time frame of interest.

3

Page 28: Spin-Orbit Interaction and Chaos in Celestial Mechanics

empty focus planet

ε

f

satellite

n+ η

(a)

satellite

η

g

r

planet

ϕε

(b)

Figure 1.1: (a) The elliptical path of a satellite, as seen in a reference frame centred on the primary(planet). The dashed line indicates the axis of the tidal elongation raised on the satellite by the planet,which is carried ahead of the line that joins the two bodies by the angle ε, in this position of the satellite’sorbit. The arrow marks the direction from the centre of mass of the satellite to the empty focus of its orbit.In this reference frame, the satellite’s spin rate is denoted by n + η, n being its mean motion. (b) Thepath of the planet in a reference frame centred on and rotating with the satellite’s mean motion. In thisrotating reference frame, and for small values of the eccentricity e, the planet moves about its guidingcentre, g, on an ellipse with semi-major and semi-minor axes in the ratio 2:1 (see Sections 2.6 and 4.10of [4] on the guiding centre approximation). The satellite’s spin rate equals η in this frame.

conditions, the (full) equation for the rotational motion of the satellite, whose permanent bulge is acted

upon by a surrounding planet, is ([4, 5, 13])

Cθ − 3

2(B −A)

Gmp

r3sin 2ψ = 0, (1.3)

where A < B < C are the three principal moments of inertia of the satellite, θ is the angle between the

long axis of the satellite (the axis of the ellipsoid with the smallest principal moment of inertia, A) and

the semi-major axis of the satellite’s orbit, ψ is the angle the long axis makes with the line that joins the

centres of the two bodies (see Figure 1.2), r is the radius of the orbit, and mp is the mass of the planet.

As r and ψ vary with the true anomaly f , which is a nonlinear function of time for eccentric orbits, the

4

Page 29: Spin-Orbit Interaction and Chaos in Celestial Mechanics

elliptical path

pericentreplanet

mp

r

θ

ψ

f

satellite

Figure 1.2: The coordinates used in the averaging theory of spin-orbit interaction. The long axis of thesatellite, which is modelled as an ellipsoid, makes an angle ψ with the line that connects the planet,seen as a point mass, to the satellite’s centre of mass, and an angle θ with an axis that coincides withthe major axis of the elliptical orbit of the satellite, which is assumed to be fixed in inertial space. Thesatellite rotates in the orbital plane.

equation is nonintegrable. However, near those cases of interest to us, namely, when the spin rate θ is a

rational multiple of the mean motion, an approximate equation may be derived and the motion analysed

in terms of a slowly varying variable γ = θ−pM , where M is the mean anomaly and p a rational number.

One starts by expanding the full equation in a Fourier series in terms of e and M and then averages all

the terms over one orbital period holding γ fixed to obtain ([4])

γ = −sgn(H(p, e))1

2ω20 sin 2γ . (1.4)

This is just the well-known pendulum equation, where

ω0 = n

(3

(B −AC

)|H(p, e)|

) 12

is the libration frequency, and the H(p, e) are power series in the eccentricity, which depend on the

synchronisation state p and are tabulated in Table I of [13]. Moreover, the leading term in the series is

of order e2|p−1| [4, 13].

In this analysis it is implicitly assumed that each synchronisation may be investigated independently

from all others, in that these average out to zero. The angle γ corresponds physically to the deviation of

the long axis of the satellite from the line joining the centres of the planet and satellite at pericentre.

We should emphasise that the only terms which can contribute to the averaged equation of motion

(1.4) are those for which p is an integer multiple of 1/2. This can be seen by inspecting the arguments of

the cosines in the full, series-expanded equation [4]. Therefore, in the context of the underlying theory,

spin-orbit coupling can be expected to arise only when θ is an integer multiple of half the mean motion

5

Page 30: Spin-Orbit Interaction and Chaos in Celestial Mechanics

(p = −1,− 1

2 ,+12 ,+1,+ 3

2 ,etc...).

If we assume that the rotational period of the satellite was much shorter in the beginning of time and

that its spin has been tidally braked, then a mean tidal torque term should be added to the right hand

side of the pendulum equation (1.4) in order to bring the spin closer to a synchronous state. Of course,

it should be at the same time sufficiently close to one of these states for the equation to be a good

approximation to the motion. We then have

γ = −sgn(H(p, e))1

2ω20 sin 2γ +

|〈Ns〉|C

. (1.5)

The magnitude of the tidal torque term should be lower than the amplitude of the torque on the perma-

nent bulge of the satellite; in other words, |〈Ns〉|C < 12ω

20 , so that γ changes sign and the satellite will

librate about γ = 0(θ = pn

), remaining trapped in a synchronous state. This is called the strength

criterion [4]. In this case, γ will librate about a value γ0 for which γ = 0, given by

γ0 =1

2arcsin

(2 |〈Ns〉|

sgn(H(p, e))ω20C

). (1.6)

If the tidal torque is much smaller than the torque on the quadrupole moment, then the value of γ0 is

determined by the sign of H(p, e): if H(p, e) > 0, γ0 ≈ 0 or π, and the long axis of the satellite will point

in the direction of the planet-satellite line at pericentre; otherwise, γ0 ≈ π2 or 3π

2 , and it will be the shorter

axis that will point in the direction of the planet-satellite line at pericentre.

A sufficient stability condition for the pth synchronisation may be derived by substituting equations (1.1)

and (1.2) for the tidal torque due to MacDonald [11] into the strength criterion. A straightforward calcu-

lation shows that (B −A)/C must exceed a threshold value given by ([4])

(B −AC

)t

=5

2

k2Qs

(Rsa

)3mp

ms

1

|H(p, e)|, (1.7)

where ms is the mass of the satellite. It has been shown in [4, 13] that all the spin-orbit states with

1.0 ≤ p ≤ 2.5 for the Moon, for which (B−A)/C ≈ 2.28× 10−4 ([17]), as well as those with 1.0 ≤ p ≤ 3.0

for Mercury, where it is assumed that (B − A)/C for Mercury is of the same order, are stable. Thus, we

now understand the stability of the present spin-orbit states for both the Moon and Mercury. In the next

section we make a brief review of what is believed to be the mechanics behind capture into a spin-orbit

synchronisation.

1.1.3 Capture into Synchronisation

The effective mechanism of capture into synchronisation remains relatively unknown. However, it is

widely accepted that one has to invoke dissipative torques in order to explain the state of a satellite or

planet [18]. One way of accomplishing this is by assuming that the initial spin period of the body was

short and introducing tidal torques, which act to brake this spin. The body will then pass through several

spin-orbit states, and may be captured into any one of them if the value of the spin rate is below a critical

6

Page 31: Spin-Orbit Interaction and Chaos in Celestial Mechanics

value when the angle γ in equation (1.5) enters its first libration, in analogy with a pendulum [19]. In

general, though, one cannot know a priori the initial conditions of the body and, consequently, if this

requirement will be satisfied. So, a suitable probability distribution over the initial angular velocity should

be introduced in order to calculate capture probabilities. Goldreich and Peale [13, 19], for instance,

define these capture probabilities as the ratio of the range of the energy integral of (1.5)7,

E =1

2Cγ2 − 3

4(B −A)n2H(p, e) cos 2γ, (1.8)

for which capture results, to the whole range of E for which γ becomes an angle of libration. However,

they also argue that the mean tidal torque, 〈Ns〉, has to somehow vary with γ for capture to occur,

because if it is constant, the body will simply pass through synchronisation and continue to despin

[13, 19].

Estimating capture probabilities is then just a matter of incorporating a model of tidal dissipation that

accounts for the variation of 〈Ns〉 with γ into the theory. This has been done in [13, 19] for two models in

which the tidal potential is expanded in a Fourier time series and each component of the tide is given a

phase lag [20]. In the first model, which assumes that the phase lags depend on the tidal frequencies,

a very low value of about 7% was calculated for the probability of capture of Mercury into the 3:2 state,

for e = 0.206, and this probability was found to vary with the value of (B−A)/C and the eccentricity. On

the other hand, the second model, in which only the signs, but not the magnitudes of the lags, depend

on frequency, gave a much higher value of about 70% for the corresponding probability, and this was

found to be determined by the synchronisation state p and the eccentricity alone. This model has some

shortcomings though, as it leads to unrealistic constant torques and doesn’t account for the damping of

the librations of the body [4].

Goldreich and Peale [13] have also computed capture probabilities for Mercury using the tidal torque

in equations (1.1) and (1.2), but this torque is actually unrealistic and is only studied for its simplicity.

First, in keeping the geometric lag angle ε constant, it makes the time lag of the tide vary with a body’s

position in its orbit for a nonvanishing eccentricity [19], and second, it permits for abrupt switches of the

torque and is inherently contradictive as it hinges on a tacit assumption that the time lag is constant [21].

An alternative mechanism for Mercury’s capture into the 3:2 state has been proposed in [22], where it

is argued that for any eccentricity, the spin rate of a body is naturally driven towards an equilibrium value

which depends on its current eccentricity. Then, since the eccentricity varies due to the chaotic evolution

of Mercury’s orbit, the spin rate can be raised and lowered, thus making it possible for a synchronisation

to be repeatedly crossed, and increasing Mercury’s probability of becoming trapped. Correia and Laskar

[22] have computed a probability of capture into the 3:2 state for Mercury of 55.4%.

1.1.4 Surface of Section and Chaos

The spin-orbit problem, as defined in Section 1.1.2, may be analysed by producing a Surface of Sec-

tion (SOS) of the motion (Poincare [23], Henon and Heiles [24]), which allows us to test the validity of

7The energy integral is just the result of the integration with respect to time of equation (1.5).

7

Page 32: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Figure 1.3: A SOS for α = 0.2 and e = 0.1 and ten different trajectories in(θ, θ

)phase space. Values of

θ and θ are plotted at successive pericentre passages, and units have been chosen such that the meanmotion is one. The image is taken from [6].

the averaging theory. This is obtained by looking at the system stroboscopically every time the satellite

passes through pericentre and plotting the corresponding values of θ against θ that result from a numer-

ical integration of the full equation of motion, equation (1.3). The variable γ (taken modulo π) is simply

θ at pericentre passage, so librations in γ are equivalent to librations in θ on the SOS [6]. This is why

the choice of the pericentre is a convenient one. A SOS for α =√

3(B −A)/C = 0.2 and e = 0.1 is

shown in Figure 1.3, which illustrates the clear separation between chaotic behaviour and quasiperiodic

motion characteristic of the phase space of Hamiltonian systems with two or more degrees of freedom

[16]. While close to the stable equilibrium points p = +1, p = + 12 and p = + 3

2 trajectories follow a closed

path, successive points in non-synchronous rotation trajectories trace a smooth curve which covers all

values of θ, and motion at the separatrices is chaotic. According to a discussion in [4], there is a very

good agreement between the trajectories shown in Figure 1.3 and the analytic solutions that result from

the averaging method. This is not always the case though, as the half-widths of the libration regions,

measured by the libration frequency ω0, grow with the asphericity parameter α, and, at some point,

the synchronous states will begin to overlap. Chirikov’s resonance overlap criterion then states that as

soon as the sum of two unperturbed half-widths matches the separation of the stable equilibrium points,

chaos will develop [25]. An estimate of the critical value of α above which chaos is expected, made with

Chirikov’s criterion applied to the p = +1 and p = + 32 states, is given by

αc =1

2 +√

14e, (1.9)

and this prediction is actually in excellent agreement with the numerical results [6].

In Figure 1.4 a SOS for values of α and e appropriate for Hyperion is shown, and it is readily apparent

that much of the simplicity of the previous picture is gone. A large chaotic sea now surrounds all the

8

Page 33: Spin-Orbit Interaction and Chaos in Celestial Mechanics

stable islands from p = + 12 to p = +2, and the averaging method is no longer applicable, since each

synchronous state can no longer be taken in isolation. This is primarily a consequence of the irregular

shape of Hyperion, which grants the moon of Saturn a value of α much larger than the critical one,

αH = 0.89, but also of its high forced eccentricity, e ≈ 0.1, due to an orbit-orbit synchronisation with Titan

[6]. Furthermore, the p = + 32 island has disappeared altogether, and a new, second-order8, p = + 9

4

island is now visible in the top centre of the chaotic sea. There is also a displacement associated with

a forced libration with the same period as the orbital period and amplitude equal to the displacement of

the island centres from the theoretical p values [6].

Figure 1.4: A SOS for α = 0.89 and e = 0.1, values appropriate for Hyperion. The image is taken from[16].

Analysis by Wisdom [6, 16, 26] of the stability of the spin axis of Hyperion at the libration regions

and in the chaotic sea has shown that the p = +1 state as well as the chaotic zone are mainly attitude

unstable, which means that the slightest deviation of the spin axis from the orbit normal while the satellite

is in one of these zones grows exponentially on a short timescale, and the satellite tumbles chaotically

in three-dimensional space. Ground-based observations have actually confirmed this behaviour of Hy-

perion’s axis [7]. In fact, though, the half-width of the chaotic separatrix also grows exponentially with α

(while depending only linearly on the eccentricity), and hence, even the small, irregularly shaped satel-

lites in the solar system rotating in the p = +1 state have significant chaotic zones associated with the

separatrix. It has also been shown by Wisdom [16, 26] that this chaotic zone around the 1:1 state is

attitude unstable, so he has argued that all satellites with irregular shapes presently in this spin-orbit

state may have spent a considerable9 amount of time tumbling chaotically before the spin finally left this

zone.

We end this section by noting that capture into synchronisation will only occur as described in Sec-

tion 1.1.3 if the reduction in energy over one librational cycle due to tidal torques is significantly greater

8In addition to primary synchronisations, those studied in 1.1.2 under the averaging theory, the phase space may containsecondary, or even more complicated synchronisations, which only appear under a second-order treatment of spin-orbit dynamicsor in a SOS.

9of the order of the tidal despinning time.

9

Page 34: Spin-Orbit Interaction and Chaos in Celestial Mechanics

than the width of the chaotic separatrix. Otherwise, the capture process becomes essentially a random

process [6].

Although the tidal dissipation hypothesis may be invoked to justify capture into synchronisation, this

hypothesis seems to be weak. Not only the models of tidal dissipation rely on poorly determined pa-

rameters which are related to the internal structure of the bodies involved in the spin-orbit interaction,

tides on small, solid bodies like some moons or asteroids are much weaker than those that act on larger

bodies in hydrostatic equilibrium [5]. In Chapter 2, we follow a different approach. We examine a simpler

and minimal model for the spin-orbit interaction, which we hope describes the different types of spin-orbit

synchronisation effects and resonances — the Keplerian Dumbbell (KD) system. This model describes

the satellite as two point masses joined by a rigid massless rod, resembling the shape of a dumbbell,

hence the name. We study in detail the dynamics of the KD system, by deriving the exact equations of

motion and then analysing the steady state orbits and their stability.

Chapter 3 deals with the extension of the KD model into the framework of the R3BP, where Nix (the

satellite) is modelled by a dumbbell which is subjected to the gravitational fields of Pluto and Charon

(the primaries). The exact equations of motion for this problem are derived, and two special cases are

studied: the Planar Circular Restricted Three-Body Problem (PCR3BP) with dumbbell satellite, in which

the dumbbell’s centre of mass is constrained to move in the plane of the orbits of the primaries, and

the so-called Sitnikov Problem (SP), where it is constrained to move along an axis that is orthogonal to

this plane and that passes through the barycentre of the primaries [27]. Necessary conditions for these

motions to occur are determined, and the steady states of the system and other types of solutions are

found. We also obtain sufficient conditions for the existence of a Jacobi-like invariant ([28]) in the circular

problem. The chapter concludes with a numerical analysis of a subcase of the PCR3BP with dumbbell

satellite, for which the spin axis of the dumbbell is perpendicular to the orbital plane of the primaries.

In the last chapter, we summarise our results and draw some conclusions.

10

Page 35: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Chapter 2

Stationary Solutions of the Keplerian

Dumbbell System

The Keplerian Dumbbell (KD) system is a three-body problem where two of the bodies are connected

by a rigid massless rod — the dumbbell. The KD system is the simplest possible system exhibiting the

minimal features of a rigid body, namely, it has more than one point mass, it is extended in space, and

the distance between masses is conserved during motion. Besides, such a system may prove useful in

describing the dynamics of elongated asteroids with a prolate spheroidal shape like 4769 Castalia [29],

433 Eros [30] or Nix [1], or small satellites [31].

The simplest system showing spin-orbit coupling and eventually synchronisation and resonance be-

tween the translational and rotational motion is the KD system. Despite the apparent simplicity of the

KD system, an exact analysis of its full dynamics, steady state orbits, eventual synchronisation and res-

onance is lacking. Recently, it has been found numerically that the KD system, restricted to a planar

elliptic orbit, shows chaotic dynamics and is nonintegrable [32, 33]. For the planar cases and under

approximate Hamiltonian equations of motion, the existence of steady state orbits has been analysed

by Celletti and Sidorenko [34]. The stability of the steady state orbits starting from an approximate La-

grangian approach has been partially analysed by Elipe et al. [35]. In all these approaches the equations

of motion are approximate. For an extensive review of previous works on the KD system, we refer to

[34].

In this chapter, we derive the exact equations of motion of the KD system, following a Lagrangian

approach. The equations of motion are not restricted to the plane containing the three masses and are

written in the inertial reference frame of the three-body system. We then analyse the steady or stationary

orbits of the KD system. For the case where the two masses of the dumbbell are equal and the motion

of the centre of mass of the dumbbell is planar, the steady states are Lyapunov unstable (Appendix A).

As these steady states are not preserved if the masses of the dumbbell are different, this shows that

this system is structurally unstable. Some of these structurally unstable systems are Eulerian solutions

of the general three-body problem (see Appendix B). For the case where the dumbbell is aligned with

the direction connecting its centre of mass to the centre of mass of the KD system, there are unstable,

11

Page 36: Spin-Orbit Interaction and Chaos in Celestial Mechanics

planar steady state trajectories. This result is exact, suggesting that synchronised spin-orbit effect in

celestial mechanics is an intrinsic non-linear phenomenon.

This chapter is organised as follows: in the next section we derive the Lagrangian and Hamiltonian

equations of motion of the KD system. In Section 2.2, we rescale these equations to a dimensionless

form. The new, dimensionless form simplifies the numerical analysis of the equations of motion. In

Section 2.3, we analyse all the steady states of the KD system and their stability. Finally, in Section 2.4

we complement the stability analysis with numerical simulations.

2.1 Equations of motion of the Keplerian Dumbbell system

The Keplerian Dumbbell (KD) system consists of a primary point mass m1 and a dumbbell, interacting

through their mutual gravitational force. The dumbbell is formed by two point masses m2 and m3,

connected by a rigid massless rod of length `. To describe the motion of the KD system, we consider

the inertial reference frame S = (Cxyz) centred at the centre of mass of the three masses (Figure 2.1).

Figure 2.1: Reference frames of the KD problem. In the reference frame S of the centre of mass of thethree-body system, the centre of mass C ′ of the dumbbell has cylindrical coordinates (r, θ, z) referred tomass m1. In S ′, the orientation of the dumbbell is specified by the angles ψ (polar) and φ (azimuthal).The orientations of the coordinate axes of S and S ′ are the same.

The dumbbell is allowed to rotate in the three-dimensional ambient space, and the configuration

manifold of each of the dumbbell masses is a sphere S2, centred at the centre of massC ′ of the dumbbell.

To describe the attitude dynamics of the dumbbell relative to the reference frame S ′ = (C ′xyz), we

consider the azimuthal and the polar spherical angles φ and ψ, respectively. The distances of the masses

m2 and m3 to the centre of mass C ′ of the dumbbell are `2 = `m3/(m2 +m3) and `3 = `m2/(m2 +m3).

The unit vectors of the coordinate axes (x, y, z) are {e1, e2, e3}. We denote by γ the unit vector directed

along the dumbbell towards mass m2. In spherical coordinates, γ = cosφ sinψ e1 + sinφ sinψ e2 +

12

Page 37: Spin-Orbit Interaction and Chaos in Celestial Mechanics

cosψ e3. The projection of the rod on the (x, y)-horizontal plane of S ′ is p(φ) = cosφ e1 + sinφ e2,

then we define a new unit vector α = p(φ+ π/2) = − sinφ e1 + cosφ e2 (Figure 2.1). As γ • α = 0, α is

perpendicular to γ. On the other hand, as α•e3 = 0, the two vectors {γ, e3} define a plane perpendicular

to α and, therefore, the angular velocity of the dumbbell around the instantaneous direction of rotation

α is ψ. Let β = γ ∧ α = − cosψ cosφ e1 − cosψ sinφ e2 + sinψ e3 be a third unit vector. Then the unit

vectors {α,β,γ} are mutually perpendicular and define the principal axes of inertia of the dumbbell.

In the reference frame {C ′αβγ}, the inertia tensor of the dumbbell is a diagonal matrix, whose

diagonal elements are

Iα = Iβ = m2`22 +m3`

23 =

m2m3

m2 +m3`2 and Iγ = 0. (2.1)

To calculate the equations of motion of the KD system we follow a Lagrangian perspective. Let r1 and

rC′ be the position vectors of the mass m1 and of the centre of mass of the dumbbell in the reference

frame S (Figure 2.1). As the kinetic energy of the dumbbell is the sum of the kinetic energy of its centre

of mass, assuming that the total mass m2 +m3 is concentrated at C ′, plus the kinetic energy of rotation,

the total kinetic energy of the KD system is

T =1

2m1 ‖r1‖2 +

1

2(m2 +m3) ‖ ˙rC′‖2 +

1

2ωT

Iω, (2.2)

where↔

I is the tensor of inertia of the dumbbell in the reference frame (C ′αβγ), and ωT, the instanta-

neous rotational angular velocity of the dumbbell evaluated in the same frame, is

ωT = ψα+ φ e3 = ψα+ φ sinψ β + φ cosψ γ. (2.3)

By construction, e3 = sinψ β + cosψ γ. Introducing (2.1) and (2.3) into (2.2), we finally conclude that

T =1

2m1 ‖r1‖2 +

1

2(m2 +m3) ‖ ˙rC′‖2 +

1

2

m2m3

m2 +m3`2(ψ2 + φ2 sin2 ψ

). (2.4)

The potential energy of the KD system, in the inertial reference frame S, is

U = −Gm1

(m2

‖r12‖+

m3

‖r13‖

), (2.5)

where ‖r12‖ and ‖r13‖ are the distances between mass m1 and each of the other two masses, m2 and

m3, respectively. Hence, in the inertial reference frame S, by (2.4) and (2.5), the Lagrangian of the KD

system is

L = T − U =1

2m1 ‖r1‖2 +

1

2(m2 +m3) ‖ ˙rC′‖2

+1

2

m2m3

m2 +m3`2(ψ2 + φ2 sin2 ψ

)+Gm1

(m2

‖r12‖+

m3

‖r13‖

).

(2.6)

We can now rewrite the Lagrangian (2.6) as a function of the relative distances between masses.

13

Page 38: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Defining rC′ = r1 + r1C′ and, as in the center of mass reference frame S, m1r1 + (m2 +m3) rC′ = 0,

we have r1 = − m2 +m3

m1 +m2 +m3r1C′

rC′ =m1

m1 +m2 +m3r1C′ .

(2.7)

Therefore, the Lagrangian (2.6) rewrites as

L =1

2

m1 (m2 +m3)

m1 +m2 +m3‖ ˙r1C′‖2 +

1

2

m2m3

m2 +m3`2(ψ2 + φ2 sin2 ψ

)+Gm1

(m2

‖r12‖+

m3

‖r13‖

),

(2.8)

where ‖r1C′‖ is the distance between the mass m1 and the centre of mass of the dumbbell. Introducing

cylindrical coordinates (r, θ, z), r1C′ = r cos θ e1 + r sin θ e2 + z e3, the Lagrangian (2.8) becomes

L =1

2

m1 (m2 +m3)

m1 +m2 +m3

(r2 + r2θ2 + z2

)+

1

2

m2m3

m2 +m3`2(ψ2 + φ2 sin2 ψ

)+Gm1

(m2

‖r12‖+

m3

‖r13‖

),

(2.9)

where the distances between masses are

r12 = r1C′ + rC′2 =

r cos θ + m3

m2+m3` sinψ cosφ

r sin θ + m3

m2+m3` sinψ sinφ

z + m3

m2+m3` cosψ

,

r13 = r1C′ + rC′3 =

r cos θ − m2

m2+m3` sinψ cosφ

r sin θ − m2

m2+m3` sinψ sinφ

z − m2

m2+m3` cosψ

,

rC′2 = `2γ,

rC′3 = −`3γ.

(2.10)

Due to the large number of parameters in the Lagrangian (2.9), we rescale this Lagrangian to dimen-

sionless variables.

2.2 Equations of motion in dimensionless form

We introduce the scaling constants and new variables through the relations

r = r0u, z = r0v, ` = r0ε, t =1

στ, (2.11)

where (u, v) are the new dimensionless variables, r0 is assumed to be a characteristic length of the KD

system, ε is a rescaled small parameter, σ is a normalisation unit of frequency, and τ is a dimensionless

14

Page 39: Spin-Orbit Interaction and Chaos in Celestial Mechanics

parameter measuring time in units of 1/σ. In the rescaled variables, the Lagrangian (2.9) transforms into

L =Gm1 (m2 +m3)

r0

[1

2

σ2r03

G (m1 +m2 +m3)

(u′

2+ v′

2+ u2θ′

2)

+1

2

(1− δ)δGm1

σ2r03ε2(ψ′

2+ φ′

2sin2 ψ

)]+Gm1 (m2 +m3)

r0

1− δ‖ρ12‖

+Gm1 (m2 +m3)

r0

δ

‖ρ13‖,

where‖ρ12‖2 = u2 + v2 + δ2ε2 + 2 δ ε (u sinψ cos(θ − φ) + v cosψ)

‖ρ13‖2 = u2 + v2 + (1− δ)2 ε2 − 2 (1− δ) ε (u sinψ cos(θ − φ) + v cosψ)

(2.12)

and δ = m3/(m2 +m3). The prime symbol (′) denotes derivation with respect to dimensionless time

τ . The constant g = Gm1 (m2 +m3) /r0 has dimensions of an energy and represents the energy scale

of the problem. We take as our new Lagrangian, which we call L, the dimensionless quantity which

multiplies this prefactor. This is just the Lagrangian written in units of g. We further make a choice of the

units of length and time such that σ2r03 = G (m1 +m2 +m3), and we finally get the new dimensionless

Lagrangian

L =1

2

(u′

2+ v′

2+ u2θ′

2)

+1

2

(1− δ)δµ

ε2(ψ′

2+ φ′

2sin2 ψ

)+

1− δ‖ρ12‖

‖ρ13‖, (2.13)

where µ = m1/(m1 +m2 +m3), and the dimensionless distances between the masses are ‖ρ12‖ and

‖ρ13‖, as defined in (2.12).

Therefore, from the dimensionless Lagrangian (2.13), the equations of motion in the (u, θ, v, φ, ψ)

coordinates are

u′′ − uθ′2 = −(1− δ)u+ δ ε sinψ cos (θ − φ)

‖ρ12‖3+ δ

(1− δ) ε sinψ cos (θ − φ)− u‖ρ13‖3

u2θ′′ + 2uu′θ′ = u(1− δ)δ ε sinψ sin (θ − φ)

(1

‖ρ12‖3− 1

‖ρ13‖3

)v′′ = −(1− δ)v + δ ε cosψ

‖ρ12‖3− δ v − (1− δ) ε cosψ

‖ρ13‖3

ε(φ′′ sin2 ψ + 2φ′ ψ′ sinψ cosψ

)= −µu sinψ sin (θ − φ)

(1

‖ρ12‖3− 1

‖ρ13‖3

)

ε(ψ′′ − φ′2 sinψ cosψ

)= −µ (u cosψ cos (θ − φ)− v sinψ)

(1

‖ρ12‖3− 1

‖ρ13‖3

).

(2.14)

The dimensionless equations of motion (2.14) depend on the three parameters δ, ε and µ. The

parameter δ ∈ (0, 1) measures the relative weight of the masses of the dumbbell. For a symmetric

dumbbell with m2 = m3, δ = 1/2. The parameter ε measures the length of the rod of the dumbbell in

units of r0, the approximate radius of the trajectory of the dumbbell. For a dumbbell satellite or asteroid,

ε is close to zero. If ε = 0, we have a Kepler problem. The mass parameter 0 < µ ≤ 1, measures the

relation between the masses of the primary and of the dumbbell. For a small dumbbell satellite, µ ' 1.

The left hand side of the second equation in (2.14) can be written in the formd(u2θ′

)dτ

, and u2θ′ is

15

Page 40: Spin-Orbit Interaction and Chaos in Celestial Mechanics

proportional to the angular momentum of the centre of mass of the dumbbell with respect to the centre

of mass of the three-body system.

The dimensionless equations (2.14) can be derived from an effective Hamiltonian, with the standard

techniques of Hamiltonian dynamics, [36]. In fact, from the Lagrangian (2.13), the conjugate momenta

are

pu = u′, pθ = u2θ′, pv = v′, pφ =(1− δ)δ

µε2φ′ sin2 ψ, pψ =

(1− δ)δµ

ε2ψ′, (2.15)

and the dimensionless Hamiltonian associated with the Lagrangian (2.13) is

H =1

2

(pu

2 +pθ

2

u2+ pv

2

)+ µ

pφ2

2(1− δ)δε2 sin2 ψ+ µ

pψ2

2(1− δ)δε2

− 1− δ‖ρ12‖

− δ

‖ρ13‖. (2.16)

The five second order equations of motion (2.14) have the Hamiltonian (2.16) and the angular mo-

mentum as conservation laws. The angular momentum of the KD system is

L =m1(m2 +m3)

m1 +m2 +m3r1C′ ∧ ˙r1C′ +

m2m3

m2 +m3`2(ψα+ φ sinψβ

)=

m1(m2 +m3)

m1 +m2 +m3

((rz sin θ − z(r sin θ + rθ cos θ))e1

+((zr − rz) cos θ − rzθ sin θ)e2 + r2θe3

)+

m2m3

m2 +m3`2(ψα+ φ sinψβ

),

(2.17)

where, in the first line, the first term on the right hand side of (2.17) is the angular momentum of the

centre of mass of the dumbbell, relative to the centre of mass of the KD system, and the second term

is the angular momentum of the dumbbell relative to its own centre of mass. Introducing the rescaling

variables (2.11) into (2.17), the angular momentum written in units of g/σ (units of angular momentum)

isL = (uv′ sin θ − v(u′ sin θ + uθ′ cos θ))e1

+((vu′ − uv′) cos θ − uvθ′ sin θ)e2 + u2θ′e3

+δ(1− δ)ε2

µ(ψ′α+ φ′ sinψβ) .

(2.18)

As L is a constant of motion, by (2.18) it follows that the dumbbell has planar motion only if

u2θ′ +δ(1− δ)ε2

µφ′ sin2 ψ = pθ + pφ = Lz = constant. (2.19)

This planar conservation law, derived from (2.18), could also have been deduced from (2.14), by adding

the second and fourth equations. In general, the dumbbell has an intrinsic three-dimensional motion in

configuration space.

16

Page 41: Spin-Orbit Interaction and Chaos in Celestial Mechanics

2.3 Steady states and stability analysis

The fixed points of the system of equations (2.14) are the steady states of the KD system. Here, we call

steady states to periodic trajectories of the dumbbell. To calculate the coordinates of the fixed points of

the system of equations (2.14), we impose that some of the components of the vector field defined by

(2.14) are zero (see Section A.1 of Appendix A).

2.3.1 Steady state 1

With the choice u = 0 in the first equation in (2.14), the centre of mass of the KD system coincides

with the position of the primary mass m1. Therefore, to fulfill a fixed point condition we must also have

v = 0, together with the two speed conditions u′ = 0 and v′ = 0. In this case, by (2.12), ‖ρ12‖ = δε and

‖ρ13‖ = (1 − δ)ε, and the right hand side of the first equation in (2.14) is zero only if δ = 1/2, implying

that m2 = m3 and ‖ρ12‖ = ‖ρ13‖. The second equation in (2.14) is identically zero, the orbital angular

momentum of the dumbbell is also zero, and the dumbbell masses rotate with constant angular speed

θ′ = ω = ±(2√

2/ε)3/2, as we shall prove below in (2.24). Therefore, in the configuration space, the KD

system is constrained to move along a circumference of arbitrary radius, with centre at the primary body

(Figure 2.2). If m2 6= m3, this steady state does not exist.

We now analyse the possible trajectories of the dumbbell around the primary body m1. Under the

conditions just derived, by (2.14), the attitude of the dumbbell around the primary is described by the

equations d

dt

(φ′ sin2 ψ

)= 0

ψ′′ − φ′2 sinψ cosψ = 0.(2.20)

The first equation in (2.20) is a conservation law, implying that φ′ sin2 ψ = c, where c is constant. If

c = 0, then φ(t) is constant, for every t ∈ R, and the second equation in (2.20) reduces to ψ′′ = 0. In

this case, the dumbbell system rotates around the primary, with fixed azimuthal angle φ =constant. In

Figure 2.2a), we depict a solution trajectory of the KD system for φ(t) = 0, for every t ∈ R.

If c 6= 0, the second equation in (2.20) reduces to

ψ′′ − c2 cosψ

sin3 ψ= 0. (2.21)

In the two-dimensional (ψ,ψ′) phase space, equation (2.21), with ψ ∈ (0, π), has a unique fixed point with

coordinates (ψ = π/2, ψ′ = 0). By (2.16) and (2.15), equation (2.21) is derivable from the Hamiltonian

H∗(ψ,ψ′) = ψ′2/2 + c2/(2 sin2 ψ). For the fixed point (ψ = π/2, ψ′ = 0), H∗(π/2, 0) = c2, φ′ = c, and the

dumbbell has a three-body Eulerian type solution (Figure 2.2b)), [37].

It is easily shown that the fixed point (ψ = π/2, ψ′ = 0) of equation (2.21) is stable of the centre

type. For any initial condition away from the fixed point in the (ψ,ψ′) phase space, the dumbbell librates

around (ψ = π/2, ψ′ = 0). Due to the first equation in (2.20), φ(t) is also periodic, with the same period

as ψ(t), and the trajectory in the configuration space is also periodic (Figure 2.2c)).

The solutions of the KD system depicted in Figure 2.2 are also Eulerian type solutions of the general

17

Page 42: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Figure 2.2: Steady solutions in the three-dimensional configuration space of the dumbbell with equalmasses (m2 = m3). The positions of the dumbbell masses m2 and m3 at time τ = 40 are shown. In a),c = 0 and, in the configuration space, the KD rotates around the primary body. This trajectory has beencalculated for the azimuthal angle φ(t) = 0. In b) and c), c = 1. In b), the dumbbell has been obtainedwith the initial condition (ψ = π/2, ψ′ = 0), and φ′ 6= 0 is a constant. In c), the dumbbell librates aroundthe fixed point (ψ = π/2, ψ′ = 0) in the (ψ,ψ′) phase space, which corresponds to a closed trajectory inthe configuration space. This orbit has been calculated with the initial conditions φ(0) = 0, ψ(0) = π/6and φ′(0) = 0. All the solutions depicted are also Eulerian solutions of the general three-body problem(see Appendix B).

three-body problem, provided m2 = m3 (Appendix B). These solutions can be obtained by numerical

integration of the three-body equations with an appropriate choice of initial conditions, [37, 38].

The periodic orbits of the KD system in Figure 2.2 occur for the following conditions:

u = 0; v = 0; u′ = 0; v′ = 0; m2 = m3. (2.22)

This fixed point, or the Eulerian trajectories in Figure 2.2, are Lyapunov unstable (Section A.2 of Ap-

pendix A). In fact, linearising the first equation in (2.14) around (2.22), with θ′ = ω and ψ = π/2 (Fig-

ure 2.2b), we obtain

u′′ = (ω2 + 16/ε3)u,

showing that the dumbbell Eulerian trajectories are unstable for perturbations in the radial coordinate u.

On the other hand, as these steady trajectories only exist for m2 = m3, any infinitesimal variation on

the value of the masses destroys the periodic orbits. This implies that the dynamical system (2.14), with

m2 = m3, is structurally unstable, [39].

2.3.2 Steady state 2

Assuming that the masses of the dumbbell are equal, m2 = m3, we have δ = 1/2. Then the system of

equations (2.14) has the equilibrium solution

u = u0 > 0; v = 0; ψ = 0, π;

u′ = v′ = ψ′ = 0; m2 = m3,(2.23)

according to which the dumbbell is always oriented vertically. As the centre of mass of the dumbbell

describes a circular orbit of radius u0 with v = 0 in configuration space, by the second equation in (2.14),

18

Page 43: Spin-Orbit Interaction and Chaos in Celestial Mechanics

the angular momentum of the centre of mass of the dumbbell Lz = u2θ′ is conserved, and the distances

‖ρ12‖ and ‖ρ13‖ are equal. Equating u′′ to zero in the first equation in (2.14), and by (2.12), we obtain

θ′2

:= ω2 =1(

u02 + ε2

4

)3/2 . (2.24)

Conditions (2.23) and (2.24) define a continuous family of periodic solutions of the KD system, pa-

rameterised by u0, with angular velocity θ′, which can be positive or negative. This relationship is the

equivalent of the third law of Kepler for KD systems. As this fixed point only exists for m2 = m3, any

infinitesimal perturbation of the values of the masses destroys the steady state 2 and the dynamical

system (2.14) is structurally unstable.

To show that the steady state 2 is unstable, we linearise the first equation in (2.14) around (2.23),

obtaining u′′ = (u − u0)3u20/(u20 + ε2/4)5/2, and u0 is calculated from (2.24). This system is clearly

unstable for small perturbations around u = u0.

In Figure 2.3a), we show the orbit of steady state 2 in configuration space.

Figure 2.3: Circular orbits (steady states 2 and 3) in the three-dimensional configuration space of theKD system, with m2 = m3. These are the dumbbell isosceles steady state configurations. In both cases,these trajectories are unstable. The positions of the dumbbell masses m2 and m3 at time τ = 29 areshown. In a), we show a circular orbit for the steady state 2, where ψ(t) = 0 or ψ(t) = π, for every t ≥ 0.The angular velocity of the dumbbell is given by (2.24). In b), we show the stationary circular orbit forthe steady state 3, where ψ(t) = π/2 and θ(t) − φ(t) = ±π/2, for every t ≥ 0. The rotational angularvelocity of the dumbbell has the same sign as the angular velocity of translation.

2.3.3 Steady state 3

The third steady state is similar to the previous one, but occurs for a different orientation of the dumbbell:

u = u0 > 0; v = 0; θ − φ = ±π/2; ψ = π/2;

u′ = v′ = ψ′ = 0; θ′ = φ′; m2 = m3.(2.25)

19

Page 44: Spin-Orbit Interaction and Chaos in Celestial Mechanics

For this family of solutions of equations (2.14), the centre of mass of the dumbbell describes a circular

orbit of radius u0 at v = 0 in configuration space, and the radius and the angular velocity are also related

through (2.24).

In Figure 2.3b), we show the orbit in configuration space of this circular steady state. As m2 = m3,

the angular momentum is also conserved. From the point of view of an observer at the origin of the

coordinate system S, m2 and m3 maintain the same relative orientation, and the rotation of the dumbbell

in the local reference frame S ′ is locked with the translation. This corresponds to spin-orbit coupling in a

1 : 1 synchronisation. However, as m2 = m3, any infinitesimal perturbation on the masses destroys the

periodic point as in the previous steady states. As before, linearising the first equation in (2.14) around

(2.25), we obtain u′′ = (u− u0)3u20/(u20 + ε2/4)5/2, and, therefore, this steady state is also unstable.

We can summarise the stability properties of the steady states 1, 2 and 3 in the theorem:

Theorem 1. The steady states (2.22), (2.23) and (2.25) of the KD system, described by equations (2.14)

and depicted in Figures 2.2 and 2.3, are unstable (to small variations on the initial conditions). Moreover,

if m2 = m3, the KD equations of motion (2.14) are structurally unstable.

2.3.4 Steady states 4 to 6

The system of equations (2.14) has three more steady states:

u = u0 > 0; v = 0; θ − φ = 0, π; ψ = π/2;

u′ = v′ = ψ′ = 0; θ′ = φ′.(2.26)

These steady states are numbered according to the conditions:

steady state 4: m2 < m3, φ = θ

steady state 4b: m2 > m3, φ = θ + π

steady state 5: m2 > m3, φ = θ

steady state 5b: m2 < m3, φ = θ + π

steady state 6: m2 = m3, φ = θ or φ = θ + π.

(2.27)

Interchanging m2 and m3, the pairs of steady states 4 and 4b, and 5 and 5b, correspond to the same

geometric arrangement of the dumbbell in configuration space.

From the first equation in (2.14), u0 is a solution of the equation

f(u) = uθ′2 − (1− δ) u± δ ε

‖ρ12‖3+ δ±(1− δ) ε− u‖ρ13‖3

= 0, (2.28)

where‖ρ12‖2 = (u± δ ε)2 ,

‖ρ13‖2 = [u∓ (1− δ) ε]2 ,(2.29)

and the upper signs correspond to the case θ − φ = 0, while the lower signs correspond to θ − φ = π.

20

Page 45: Spin-Orbit Interaction and Chaos in Celestial Mechanics

As these two choices can be converted into one another by interchanging m2 and m3, in the following, it

is enough to consider the upper signs in (2.28) and (2.29). So, under these conditions, the u coordinate

of each steady state is a solution of

f(u) = uθ′2 − (1− δ) 1

(u+ δε)2− δ 1

(u− (1− δ)ε)2= 0, (2.30)

provided u0 − (1 − δ)ε > 0, which implies u0 > (1 − δ)ε. For these steady states, the analogous of

Kepler’s third law is given by the dependence of u0 on θ′, u0 ≡ u0(θ′, δ, ε), obtained by solving equation

(2.30). Therefore, for any choice of θ′ = ω, positive or negative, there is a unique circular trajectory of the

centre of mass of the dumbbell with radius u0, relative to the centre of mass of the KD system, obtained

from (2.30).

That there is indeed a unique solution of equation (2.28) is easily shown. In fact, as limu→+∞ f(u) =

+∞ and f(0) = −(1−3δ(1−δ))/(δ2ε2(1−δ)2) < 0, there exists at least one solution of f(u) = 0, provided

θ′ 6= 0. The unicity follows from the monotonicity of f.

As φ(t) = θ(t), the dumbbell rotates in the direction of the translational motion, and the two masses

m2 and m3 are aligned with the direction defined by the origins of the reference frames S and S ′,

Figure 2.1. The families of equilibrium solutions (2.26)–(2.27) correspond to the 1 : 1 synchronisation of

the translational and rotational motions.

In Figure 2.4, we show a sequence of dumbbell positions along a circular orbit in configuration space

S.

Figure 2.4: Circular orbit in the three-dimensional configuration space of the KD system, with m2 = m3,corresponding to steady state 6. The positions of the dumbbell masses are calculated at times τ =0, 1.4, 3.0 and 4.5. The orbit of the dumbbell is circular around the centre of mass of the KD system. Theperiod of rotation of the dumbbell around its centre of mass is the same as the period of translation ofthe dumbbell, corresponding to a 1 : 1 synchronisation.

21

Page 46: Spin-Orbit Interaction and Chaos in Celestial Mechanics

To analyse the stability of the steady states (2.27), we first consider the planar case. With the

conditions ψ = π/2, pψ = 0, v = 0 and pv = 0 in (2.16), the equations of motion in Hamiltonian form are

u′ = pu

p′u =pθu3− (1− δ) u+ δε cos(θ − φ)

(u2 + δ2ε2 + 2δεu cos(θ − φ))3/2

−δ u− (1− δ)ε cos(θ − φ)

(u2 + (1− δ)2ε2 − 2(1− δ)εu cos(θ − φ))3/2

θ′ =pθu2

p′θ =u(1− δ)δε sin(θ − φ)

(u2 + δ2ε2 + 2δεu cos(θ − φ))3/2

− u(1− δ)δε sin(θ − φ)

(u2 + (1− δ)2ε2 − 2(1− δ)εu cos(θ − φ))3/2

φ′ =µ

(1− δ)δε2pφ

p′φ = − u(1− δ)δε sin(θ − φ)

(u2 + δ2ε2 + 2δεu cos(θ − φ))3/2

+u(1− δ)δε sin(θ − φ)

(u2 + (1− δ)2ε2 − 2(1− δ)εu cos(θ − φ))3/2.

(2.31)

As the total angular momentum is conserved, by (2.31) or (2.19), we can eliminate p′φ from the above

equations, obtaining

u′ = pu

p′u =pθu3− (1− δ) u+ δε cos(θ − φ)

(u2 + δ2ε2 + 2δεu cos(θ − φ))3/2

−δ u− (1− δ)ε cos(θ − φ)

(u2 + (1− δ)2ε2 − 2(1− δ)εu cos(θ − φ))3/2

θ′ =pθu2

p′θ =u(1− δ)δε sin(θ − φ)

(u2 + δ2ε2 + 2δεu cos(θ − φ))3/2

− u(1− δ)δε sin(θ − φ)

(u2 + (1− δ)2ε2 − 2(1− δ)εu cos(θ − φ))3/2

φ′ =µ

(1− δ)δε2(Lz − pθ),

(2.32)

where we have introduced the conservation law

pθ + pφ = Lz, (2.33)

and Lz is a constant. Introducing the new variable χ = θ − φ and making the choice pχ = pθ, equations

22

Page 47: Spin-Orbit Interaction and Chaos in Celestial Mechanics

(2.32) simplify to

u′ = pu

p′u =pχu3− (1− δ) u+ δε cosχ

(u2 + δ2ε2 + 2δεu cosχ)3/2

−δ u− (1− δ)ε cosχ

(u2 + (1− δ)2ε2 − 2(1− δ)εu cosχ)3/2

χ′ =pχu2− µ

(1− δ)δε2(Lz − pχ)

p′χ =u(1− δ)δε sinχ

(u2 + δ2ε2 + 2δεu cosχ)3/2

− u(1− δ)δε sinχ

(u2 + (1− δ)2ε2 − 2(1− δ)εu cosχ)3/2.

(2.34)

Equation (2.31) is Hamiltonian, with effective Hamiltonian function

Heff =1

2

(pu

2 +pχ

2

u2− 2

µ

(1− δ)δε2Lzpχ +

µ

(1− δ)δε2pχ

2

)− 1− δ‖ρ12‖

− δ

‖ρ13‖.

(2.35)

To analyse the stability of the family of fixed points 4 to 6, we use the fact that, if the Hamiltonian

has a local maximum or minimum at the fixed point, then that fixed point is Lyapunov stable, [40]. So,

we calculate the Hessian matrix of Hamiltonian (2.35), evaluated at the fixed points under analysis. If

the Hessian matrix of Heff is positive or negative definite, then the fixed points are Lyapunov stable. On

the contrary, if the eigenvalues have different signs and at least one of them is zero, then we are in the

presence of an unstable fixed point.

In the Hamiltonian (2.35), the coordinates of the fixed points (2.29) are u = u0, χ = 0, pu = 0 and

pχ = u20ω. Then, the Hessian matrix (2.35) calculated at these fixed points is

Ms =

1 0 0 0

0 1u20

+ µ(1−6Lzu0ω)(1−δ)δε2 − 2ω

u20

0

0 − 2ωu20

3ω2

u20− 2(A(1− δ) +Bδ) 0

0 0 0 (B −A)u0(1− δ)δε

, (2.36)

where

A =1

(u0 + δε)3, B =

1

|(u0 − (1− δ)ε)|3

and (B − A) > 0. As Ms is a symmetric matrix, all its eigenvalues are real. The matrix Ms has two

positive eigenvalues, λ = 1 and λ = (B −A)u0(1− δ)δε. We now prove the following theorem:

Theorem 2. For a sufficiently large radius u0 of the dumbbell circular trajectory, the steady states (2.26)–

(2.27) of the KD system, described by equations (2.14) and depicted in Figure 2.4, are Lyapunov unsta-

ble. Moreover, if m2 = m3, the KD equations of motion (2.14) are structurally unstable.

Proof. The proof is based on the analysis of the eigenvalues of the Hessian matrix Ms in (2.36) of the

Hamiltonian (2.35), calculated at the steady states. Due to the particular form of the matrixMs, it has two

positive eigenvalues λ = 1 and λ = (B −A)u0(1− δ)δε. The other two eigenvalues are the eigenvalues

23

Page 48: Spin-Orbit Interaction and Chaos in Celestial Mechanics

of the reduced matrix

Mss =

1u20

+ µ(1−6Lzu0ω)(1−δ)δε2 − 2ω

u20

− 2ωu20

3ω2

u20− 2(A(1− δ) +Bδ)

. (2.37)

Introducing the new variable ξ = 1/u0 into (2.37) and developing this matrix in Taylor series around

ξ = 0, the reduced matrix Mss becomes, up to fourth order in ξ,

Mss =

ξ2 + µ(1−δ)δε2 0

0 −2δξ2

+O(ξ5). (2.38)

Therefore, for ξ sufficiently small (or equivalently, u0 sufficiently large), Mss has one positive and one

negative eigenvalue. This property is also valid for the Hessian matrix (2.36). So, this implies that the

Hamiltonian function (2.35) is at least a 1−saddle near the fixed point, [41], and, therefore, the steady

states (2.26)–(2.27) of the KD system are unstable. The structural instability of the equations of motion

(2.14), for equal masses of the dumbbell m2 = m3, has been shown above.

In the next section, the stability properties of the steady states (2.26)–(2.27) of the KD system are

analysed numerically.

2.4 Numerical analysis of the non-stability of some steady state

orbits

In the previous section, we have shown that the steady state orbits of the KD system are Lyapunov

unstable. This has been shown rigorously for steady states (2.22), (2.23) and (2.25), and partially for

steady states (2.26)–(2.27).

In Figure 2.5, we show the numerically calculated values of the eigenvalues of the reduced matrix

Mss, as a function of the steady state orbital radius of the dumbbell, u0, for different choices of the

parameters. These numerical simulations show that the reduced matrix (2.37) can have one positive

and one negative eigenvalue, or two negative eigenvalues. This demonstrates that the Hessian matrix

(2.36) can be locally a 1−saddle or a 2−saddle, which implies globally that the fixed points (2.26)–(2.27)

of the KD system are Lyapunov unstable.

24

Page 49: Spin-Orbit Interaction and Chaos in Celestial Mechanics

1 2 3 4 5u0

-100

100

200

300

400

500

λ1

λ2

a)

1 2 3 4 5u0

-100

100

200

300

400

500

λ1λ2

b)

Figure 2.5: Eigenvalues λ1 and λ2 of the reduced matrix Mss, (2.37), as a function of the steady stateorbital radius u0 of the dumbbell. In a), δ = 0.5, µ = 1.0 and ε = 0.1. In b), δ = 0.8, µ = 1.0 and ε = 0.3.

25

Page 50: Spin-Orbit Interaction and Chaos in Celestial Mechanics

26

Page 51: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Chapter 3

Restricted Three-Body Problem with

Dumbbell Satellite

Nix is a small natural satellite of the Pluto-Charon system, discovered in 2005 along with Hydra, another

moon of the binary system [1]. In recent years, interest has risen in these small moons, not only because

the New Horizons spacecraft flew through the double planet system and captured images of their sur-

faces, but also because it was discovered that they spin and wobble unpredictably [2]. The effect is likely

due to the constantly shifting gravitational field produced by the larger bodies, Pluto and Charon, and is

only enhanced by the prolate spheroidal shape of the moons, which are the subject of asymmetric grav-

itational torques. In this chapter we study the dynamics of the three-body system constituted by Pluto,

Charon and Nix, by using a convenient toy model for Nix — that of a dumbbell satellite. This system

may be strictly regarded as a special case of the four-body problem, for which there is a constraint on

the distance between two of the bodies. This is the most simple and minimal model of a rigid body that

will enable us to analyse the mechanism of resonance and the transition to chaos of the small moon.

Furthermore, since the mass of Nix is much smaller than that of Pluto or Charon ([42]), we treat the case

in study in the framework of the Restricted Three-Body Problem (R3BP), [3].

The problem of analysing the dynamics of a dumbbell satellite in the context of the R3BP is not new.

For instance, Vera [43] considered a rigid dumbbell satellite placed at the equilateral point L4 of the

R3BP (see [44, 3] for more details) and studied the attitude dynamics of the satellite, providing sufficient

conditions for the existence of periodic orbits via Averaging Theory (Chapter 1). However, as far as

the author is aware, the full dynamics of the dumbbell, including spin-orbit interaction, has never been

explored in the literature. This is the problem we shall tackle in this chapter.

This chapter is organised as follows: in the next section we formulate the problem in more precise

terms. In Section 3.2 we derive the exact equations of motion of the system, following a Lagrangian

approach. The equations of motion are written in the inertial reference frame of the centre of mass of

the two more massive bodies. In order to simplify the analysis, we rewrite and rescale the equations

of motion in a rotating reference frame in Section 3.3. In Section 3.4 we restrict our attention to two

special cases of the R3BP with dumbbell satellite, as we call it: the Planar Circular Restricted Three-

27

Page 52: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Body Problem (PCR3BP) with dumbbell satellite, in which the dumbbell’s centre of mass is constrained

to move in the plane of the orbits of the two massive bodies, and the so-called Sitnikov Problem (SP)

(Sitnikov [27]), where it is constrained to move along an axis that is orthogonal to this plane and that

passes through the barycentre of these bodies. Necessary conditions for these motions to occur are

determined, and the steady states of the system and other types of solutions are found. We also obtain

sufficient conditions for the existence of a Jacobi-like invariant ([28]) in the circular problem. Lastly,

Section 3.5 concludes with a numerical analysis of a subcase of the PCR3BP with dumbbell satellite, for

which the axis of rotation of the dumbbell is perpendicular to the orbital plane of the massive bodies.

3.1 The Restricted Three-Body Problem

Nix’s mass is, according to estimations from orbital integration, about five and six orders of magnitude

smaller than those of Charon and Pluto, respectively [42, 45]. Consequently, it is reasonable to assume

that its gravitational influence on the much more massive bodies is quite negligible. This is the premise

of the R3BP: one of the masses of the bodies is taken to be infinitely small, so that it does not perturb

the motion of the other two, also called the primaries [3]. The motion of the primaries becomes then

decoupled of the third body, and one investigates the motion of the latter subject to the a priori known

motions of the former. This model plays an important role in the analysis of the motions of artificial

satellites or small asteroids ([38]), and its study was started by Newton, d’Alembert, Euler, Lagrange

and Poincare more than three centuries ago [46, 44, 28, 23, 47]. For an extensive review of previous

work done on the R3BP, we refer the reader to Musielak and Quarles [38] and references therein. For a

mathematical description of the problem, we refer to Szebehely [3].

In the context of the R3BP, let us then consider two point masses m1 and m2 describing Keplerian

orbits around their common centre of mass in a plane — the Lagrange plane. This plane is conserved

throughout the motion and is normal to the direction of the angular momentum of the two masses.

Without loss of generality, let (X, Y ) be cartesian coordinates on this plane. According to the theory

of the Kepler problem, the equation governing the motion of these masses in the centre of mass or

barycentric reference frame is (see, for instance, Murray and Dermott [4] or Fitzpatrick [5])r =

Lz2

r3− µ

r2

θ(t) = θ(0) +

∫ t

0

Lzr2dt,

(3.1)

where r > 0, the relative distance between the bodies, is given as a function of the angular coordinate θ

by

r(θ − ξ) =Lz

2/µ

1 + e cos(θ − ξ), (3.2)

which is the equation of a conic section in polar coordinates, e being the eccentricity of the conic,

Lz = r2θ the angular momentum integral, associated to the relative motion of the two masses1, and

1Lz is not the actual orbital angular momentum of the Keplerian subsystem, L1+2. Rather, it relates to L1+2 through Lz =∥∥L1+2

∥∥ (m1 +m2) /(m1m2).

28

Page 53: Spin-Orbit Interaction and Chaos in Celestial Mechanics

µ = G(m1 + m2), where G is the gravitational constant. Here we will focus solely on elliptical orbits of

the primaries, for which e ∈ [0, 1), since most bodies in the solar system, with the exception of comets,

have e � 1. In this case, the length Lz2/µ is related to the eccentricity via Lz2/µ = a(1 − e2), where a

is the semi-major axis of the ellipse, and we also have

r(θ − ξ) =a(1− e2)

1 + e cos(θ − ξ). (3.3)

The angle ξ, measured with respect to an arbitrary reference direction, is the angle at which the closest

approximation between the primaries occurs, also called the longitude of the pericentre.

Denoting by r1 and r2 the position vectors of masses m1 and m2, referred to their common centre

of mass, respectively, it is true that

m1r1 +m2r2 = 0. (3.4)

This means that r1 is always in the opposite direction to r2, and thus, the centre of mass is always on

the line joining the masses. It follows from this thatr1 = − m2

m1 +m2(r(θ − ξ) cos θ, r(θ − ξ) sin θ, 0) = (X1, Y1, 0)

r2 =m1

m1 +m2(r(θ − ξ) cos θ, r(θ − ξ) sin θ, 0) = (X2, Y2, 0) .

(3.5)

Therefore, in the inertial reference frame with origin at the centre of mass of the primaries, each mass

also traces a path in space given by the same conic section as the one describing their relative motion,

albeit with a different scale, and the centre of mass sits at one of the foci of both ellipses. The eccen-

tricities are the same, although the semi-major axes are not, therefore all the ellipses are similar; the

periods of their orbits, the mean motions and the instantaneous angular velocities are also equal, but

the pericentres differ by π (Figure 3.1).

Let us consider additionally a dumbbell-shaped satellite formed by two point masses m3 and m4,

connected by a rigid massless rod of length `, such that m3 +m4 � m1, m2. As pointed out, we assume

that these bodies don’t influence the motions of the other two. To study the dynamics of the system of

four masses, we take the inertial reference frame S = (CXY Z) with origin at the centre of mass of the

primaries m1 and m2 (Figure 3.2).

As in Chapter 2, the dumbbell is allowed to rotate in the three-dimensional ambient space, and the

configuration manifold of each of the dumbbell masses is a sphere S2, centred at the centre of mass

C ′ of the dumbbell. To describe the attitude dynamics of the dumbbell relative to the reference frame

S ′ = (C ′XY Z), we consider the azimuthal and the polar spherical angles φ and ψ, respectively. The

distances of the masses m3 and m4 to the centre of mass C ′ of the dumbbell are `3 = `m4/(m3 +m4)

and `4 = `m3/(m3 +m4), respectively. The unit vectors of the coordinate axes (X,Y, Z) are {e1, e2, e3}.

We denote by γ the unit vector directed along the dumbbell towards mass m3. In spherical coordinates,

γ = cosφ sinψ e1 + sinφ sinψ e2 + cosψ e3. The projection of the rod on the (X,Y )-horizontal plane of

S ′ is p(φ) = cosφ e1 + sinφ e2, then we define a new unit vector α = p(φ + π/2) = − sinφ e1 + cosφ e2

29

Page 54: Spin-Orbit Interaction and Chaos in Celestial Mechanics

m2

m1X

Y

C

Figure 3.1: The motion of the masses m1 and m2 in the inertial reference frame S centred at the centreof mass, C. An eccentricity of 0.5, a mass ratio m2/m1 of 0.2, and a longitude of the pericentre of π/6were used.

x y

Z, z

S,R

θm1

m2

X

e1 Ye2

e3

rC′C

φX

e1 Ye2

Z

e3

S ′

m3

m4

α

βγ

C′

ψ

Figure 3.2: Reference frames used in the study of the R3BP with dumbbell satellite. In the referenceframe S of the centre of mass of the primaries, the centre of mass C ′ of the dumbbell has cartesiancoordinates (X,Y, Z). In S ′, the orientation of the dumbbell is specified by the angles ψ (polar) and φ(azimuthal). The orientations of the coordinate axes of S and S ′ are the same. The reference frame R,also called synodic, corotates with the primaries, which are assumed to move in the (X,Y )-plane, andis rotated by an angle θ with respect to S. The centre of mass of the dumbbell has cartesian coordinates(x, y, z) in R.

(Figure 3.2). As γ • α = 0, α is perpendicular to γ. On the other hand, as α • e3 = 0, the two vectors

{γ, e3} define a plane perpendicular to α and, therefore, the angular velocity of the dumbbell around the

instantaneous direction of rotation α is ψ. Let β = γ ∧α = − cosψ cosφ e1 − cosψ sinφ e2 + sinψ e3 be

a third unit vector. Then the unit vectors {α,β,γ} are mutually perpendicular and define the principal

axes of inertia of the dumbbell.

In the reference frame {C ′αβγ}, the inertia tensor of the dumbbell is a diagonal matrix, whose

30

Page 55: Spin-Orbit Interaction and Chaos in Celestial Mechanics

diagonal elements are

Iα = Iβ = m3`23 +m4`

24 =

m3m4

m3 +m4`2 and Iγ = 0. (3.6)

3.2 Equations of motion of the dumbbell satellite

To determine the equations of motion of the dumbbell we once again follow a Lagrangian perspective.

Let r3, r4 and rC′ be the position vectors of masses m3, m4 and of the centre of mass of the dumbbell

in the reference frame S. Moreover, let us define vectors rij = rj − ri, which point from mass mi to

mass mj . Then, in S, the Lagrangian of the system of four masses is

L =1

2m1 ‖r1‖2 +

1

2m2 ‖r2‖2 +

1

2m3 ‖r3‖2 +

1

2m4 ‖r4‖2 +G

m1m2

‖r12‖

+Gm1m3

‖r13‖+G

m2m3

‖r23‖+G

m1m4

‖r14‖+G

m2m4

‖r24‖

+Gm3m4

‖r34‖,

(3.7)

where ‖rij‖ is the distance between masses i and j. As the distance between mass m3 and mass m4,

or the length of the dumbbell, is a constant of the motion, the last term of (3.7) is just a constant and does

not contribute towards the equations of motion. Furthermore, since, by assumption, the motion of the

primaries is independent from the other masses, their Lagrangian, L1+2 = m1 ‖r1‖2 /2 +m2 ‖r2‖2 /2 +

Gm1m2/‖r12‖, becomes decoupled from L. As such, the motion of the dumbbell under the action of

the gravitational field generated by m1 and m2, as these whirl about each other, is governed in S by the

Lagrangian

Ldumbb. =1

2m3 ‖r3‖2 +

1

2m4 ‖r4‖2

+Gm1

(m3

‖r13‖+

m4

‖r14‖

)+Gm2

(m3

‖r23‖+

m4

‖r24‖

).

(3.8)

On the other hand, the motions of m1 and m2 are to be determined from (3.1), which is derived from

L1+2, and whose solution is given parametrically by (3.3) and (3.5).

It is straightforward to show that the kinetic energy of the dumbbell T = m3 ‖r3‖2 /2 + m4 ‖r4‖2 /2

decomposes into a translational part, associated to the motion of C ′, plus a rotational part, that is,

T =1

2(m3 +m4) ‖ ˙rC′‖2 +

1

2ωT

Iω, (3.9)

where↔

I is the inertia tensor of the dumbbell in the reference frame (C ′αβγ), and ωT, the instantaneous

angular velocity of the dumbbell evaluated in the same frame, is

ωT = ψα+ φ e3 = ψα+ φ sinψ β + φ cosψ γ. (3.10)

31

Page 56: Spin-Orbit Interaction and Chaos in Celestial Mechanics

By construction, e3 = sinψ β + cosψ γ. Plugging (3.6) and (3.10) into (3.9), we conclude that

T =1

2(m3 +m4) ‖ ˙rC′‖2 +

1

2

m3m4

m3 +m4`2(ψ2 + φ2 sin2 ψ

), (3.11)

and, accordingly, the Lagrangian (3.8) of the dumbbell satellite in S rewrites as

Ldumbb. =1

2(m3 +m4)

(X2 + Y 2 + Z2

)+

1

2

m3m4

m3 +m4`2(ψ2 + φ2 sin2 ψ

)+Gm1

(m3

‖r13‖+

m4

‖r14‖

)+Gm2

(m3

‖r23‖+

m4

‖r24‖

),

(3.12)

where we have introduced cartesian coordinates (X,Y, Z) for the position of C ′ in S, rC′ = X e1 +

Y e2 + Z e3 (Figure 3.2). The distances between masses are, by (3.5),

r13 = rC′ + rC′3 − r1 =

X + m4

m3+m4` sinψ cosφ−X1

Y + m4

m3+m4` sinψ sinφ− Y1

Z + m4

m3+m4` cosψ

,

r14 = rC′ + rC′4 − r1 =

X − m3

m3+m4` sinψ cosφ−X1

Y − m3

m3+m4` sinψ sinφ− Y1

Z − m3

m3+m4` cosψ

,

r23 = rC′ + rC′3 − r2 =

X + m4

m3+m4` sinψ cosφ−X2

Y + m4

m3+m4` sinψ sinφ− Y2

Z + m4

m3+m4` cosψ

,

r24 = rC′ + rC′4 − r2 =

X − m3

m3+m4` sinψ cosφ−X2

Y − m3

m3+m4` sinψ sinφ− Y2

Z − m3

m3+m4` cosψ

,

rC′3 = `3γ,

rC′4 = −`4γ. (3.13)

Under these circumstances, we note that Lagrangian (3.12) depends explicitly on time through the an-

gular coordinate θ, which, when the orbits of the primaries are elliptical, is a non-linear function of time.2

2It is worth mentioning that the R3BP is a fundamentally different problem than the General Three-Body Problem. In therestricted problem neither total angular momentum nor energy is conserved, whereas they are in the General Three-Body Problemand, specifically, in the KD system, as we saw in the previous chapter.

32

Page 57: Spin-Orbit Interaction and Chaos in Celestial Mechanics

From this Lagrangian, the Euler-Lagrange equations for the motion of the dumbbell are

(m3 +m4) X = −Gm1

((m3

‖r13‖3+

m4

‖r14‖3

)(X −X1) +

m3m4

m3 +m4` sinψ cosφ

(1

‖r13‖3− 1

‖r14‖3

))

−Gm2

((m3

‖r23‖3+

m4

‖r24‖3

)(X −X2) +

m3m4

m3 +m4` sinψ cosφ

(1

‖r23‖3− 1

‖r24‖3

))

(m3 +m4) Y = −Gm1

((m3

‖r13‖3+

m4

‖r14‖3

)(Y − Y1) +

m3m4

m3 +m4` sinψ sinφ

(1

‖r13‖3− 1

‖r14‖3

))

−Gm2

((m3

‖r23‖3+

m4

‖r24‖3

)(Y − Y2) +

m3m4

m3 +m4` sinψ sinφ

(1

‖r23‖3− 1

‖r24‖3

))

(m3 +m4) Z = −Gm1

((m3

‖r13‖3+

m4

‖r14‖3

)Z +

m3m4

m3 +m4` cosψ

(1

‖r13‖3− 1

‖r14‖3

))

−Gm2

((m3

‖r23‖3+

m4

‖r24‖3

)Z +

m3m4

m3 +m4` cosψ

(1

‖r23‖3− 1

‖r24‖3

))

`(φ sin2 ψ + 2φ ψ sinψ cosψ

)= −Gm1 sinψ ((Y − Y1) cosφ− (X −X1) sinφ)

(1

‖r13‖3− 1

‖r14‖3

)

−Gm2 sinψ ((Y − Y2) cosφ− (X −X2) sinφ)

(1

‖r23‖3− 1

‖r24‖3

)

`(ψ − φ2 sinψ cosψ

)= −Gm1 (cosψ [(X −X1) cosφ+ (Y − Y1) sinφ]− Z sinψ)

(1

‖r13‖3− 1

‖r14‖3

)

−Gm2 (cosψ [(X −X2) cosφ+ (Y − Y2) sinφ]− Z sinψ)

(1

‖r23‖3− 1

‖r24‖3

),

(3.14)

where (X,Y, Z) are cartesian coordinates and (φ, ψ) are angular coordinates for the attitude of the

dumbbell.

By virtue of the complicated nature of the equations (3.14), in the next section we simplify the La-

grangian (3.12) and these equations, by making a change to another frame of reference.

3.3 Equations of motion in the synodic reference frame, in dimen-

sionless form

Let us consider a new reference frame R with the same origin as the inertial frame, but which is rotating

in the positive direction, and a new set of cartesian coordinates (x, y, z) associated to this frame. We

choose the direction of the x axis such that the two primary masses always lie along it (Figures 3.2,3.3).

This is commonly called the synodic reference frame [46, 3]. It is assumed that the coordinate axes of

the old and the new reference frames coincide at the origin of time. The coordinates of the centre of

mass of the dumbbell with respect to this frame are now (x, y, z).

By construction, R is rotated with respect to S by the angle θ of (3.1) and (3.5). The new cartesian

33

Page 58: Spin-Orbit Interaction and Chaos in Celestial Mechanics

m2

m1

x

y

θ

Xe1

Y

e2

Z, z

S

R

C

m3

m4

Figure 3.3: Inertial S and synodic R = (Cxyz) reference frames used in the study of the R3BP withdumbbell satellite. The reference frame R corotates with the masses m1 and m2.

coordinates are thus related to the ones in the inertial frame by the simple rotationX

Y

Z

=

cos θ − sin θ 0

sin θ cos θ 0

0 0 1

x

y

z

. (3.15)

Applying transformation (3.15) to (3.12), the Lagrangian of the dumbbell satellite is rewritten in the

synodic reference frame as

Ldumbb. =1

2(m3 +m4)

(x2 + y2 + z2 + θ2

(x2 + y2

)+ 2θ (xy − xy)

)+

1

2

m3m4

m3 +m4`2(ψ2 + φ2 sin2 ψ

)(3.16)

+ Gm1

(m3

‖r13‖+

m4

‖r14‖

)+Gm2

(m3

‖r23‖+

m4

‖r24‖

),

where, in view of (3.5) and (3.13), the squared distances between each of the masses of the dumbbell

m3 and m4 and the primaries m1 and m2 are now given by

‖r13‖2 =

(x+

m2

m1 +m2r(θ − ξ)

)2

+ y2 + z2 +

(m4

m3 +m4`

)2

+ 2m4

m3 +m4`

((x+

m2

m1 +m2

× r(θ − ξ))

sinψ cos (θ − φ)− y sinψ sin (θ − φ) + z cosψ

)

‖r14‖2 =

(x+

m2

m1 +m2r(θ − ξ)

)2

+ y2 + z2 +

(m3

m3 +m4`

)2

− 2m3

m3 +m4`

((x+

m2

m1 +m2

× r(θ − ξ))

sinψ cos (θ − φ)− y sinψ sin (θ − φ) + z cosψ

)

‖r23‖2 =

(x− m1

m1 +m2r(θ − ξ)

)2

+ y2 + z2 +

(m4

m3 +m4`

)2

+ 2m4

m3 +m4`

((x− m1

m1 +m2

34

Page 59: Spin-Orbit Interaction and Chaos in Celestial Mechanics

× r(θ − ξ))

sinψ cos (θ − φ)− y sinψ sin (θ − φ) + z cosψ

)

‖r24‖2 =

(x− m1

m1 +m2r(θ − ξ)

)2

+ y2 + z2 +

(m3

m3 +m4`

)2

− 2m3

m3 +m4`

((x− m1

m1 +m2

× r(θ − ξ))

sinψ cos (θ − φ)− y sinψ sin (θ − φ) + z cosψ

), (3.17)

and r(θ − ξ) is given by (3.3). Here θ is the common angular velocity of the primaries about each other

and their centre of mass, which is related to Lz through

θ =

(1 + e cos(θ − ξ)

1− e2

)2Lza2. (3.18)

Due to the large number of parameters in the Lagrangian (3.16), it is convenient to rescale it to

dimensionless variables. As such, let us introduce new variables through the relations

x = r∗u, y = r∗v, z = r∗w, ` = r∗ε, t =1

θ∗τ, (3.19)

where (u, v, w) are the new, dimensionless variables, ε is the rescaled length of the dumbbell, and

where we adopt r∗ = Lz2/µ, the radius of the circular orbit of the Kepler Problem for the motion of m2

about m1, and θ∗ = µ2/Lz3, the angular velocity corresponding to that orbit, as scaling constants.3 In

the rescaled variables, Lagrangian (3.16) is then transformed into

Ldumbb. = (m3 +m4)µ2

Lz2

[1

2

(u′

2+ v′

2+ w′

2+ θ′

2 (u2 + v2

)+ 2 θ′ (uv′ − u′v)

)+

1

2(1− δ)δ ε2

(ψ′

2+ φ′

2sin2 ψ

)]

+ (m3 +m4)µ2

Lz2 (1− η)

(1− δ‖ρ13‖

‖ρ14‖

)

+ (m3 +m4)µ2

Lz2 η

(1− δ‖ρ23‖

‖ρ24‖

), (3.20)

where

‖ρ13‖2 =(u+ η ρ(θ − ξ)

)2+ v2 + w2 + δ2 ε2 + 2 δ ε

((u+ η ρ(θ − ξ)

)sinψ cos (θ − φ)− v sinψ

× sin (θ − φ) + w cosψ)

‖ρ14‖2 =(u+ η ρ(θ − ξ)

)2+ v2 + w2 + (1− δ)2 ε2 − 2 (1− δ) ε

((u+ η ρ(θ − ξ)

)sinψ cos (θ − φ)

− v sinψ sin (θ − φ) + w cosψ)

‖ρ23‖2 =(u− (1− η) ρ(θ − ξ)

)2+ v2 + w2 + δ2 ε2 + 2 δ ε

((u− (1− η) ρ(θ − ξ)

)sinψ cos (θ − φ)

3The orbit r = r∗ is easily seen to be an equilibrium solution of (3.1) with vanishing eccentricity. On the other hand, the valueof θ∗ can be obtained for instance by making the substitutions a = r∗ and e = 0 in (3.18).

35

Page 60: Spin-Orbit Interaction and Chaos in Celestial Mechanics

− v sinψ sin (θ − φ) + w cosψ)

‖ρ24‖2 =(u− (1− η) ρ(θ − ξ)

)2+ v2 + w2 + (1− δ)2 ε2 − 2 (1− δ) ε

((u− (1− η) ρ(θ − ξ)

)× sinψ cos (θ − φ)− v sinψ sin (θ − φ) + w cosψ

), (3.21)

and η = m2/(m1 +m2), and δ = m4/(m3 +m4). The prime symbol (′) here denotes derivation with

respect to dimensionless time τ , and we have made the definition

ρ(θ − ξ) =λ(1− e2)

1 + e cos(θ − ξ), (3.22)

λ being the semi-major axis of the orbit of one of the primaries about the other normalised to r∗, i.e.

λ = a/r∗. In the synodic reference frame, m1 and m2 always lie along the x axis with coordinates

(−η ρ(θ − ξ), 0, 0) and(

(1− η) ρ(θ − ξ), 0, 0), respectively.

The constant (m3 +m4)µ2/Lz2 has the dimensions of an energy and represents the energy scale of

the problem. If we let the units of mass and length be chosen such that µ = 1 and r∗ = 1, respectively,

and we introduce a new parameter ν defined by ν = (m1 +m2) /(m1 +m2 +m3 +m4), it then follows

that in this system of units the four masses are

Gm1 = 1− η, Gm2 = η, Gm3 =(1− δ) (1− ν)

νand Gm4 =

δ (1− ν)

ν, (3.23)

and the energy scale is given by m3 +m4 = (1− ν) /(Gν). Likewise, the common angular velocity of the

primaries becomes θ = ρ−2(θ − ξ). In an analogous manner to what we did in Section 2.2, we take as

our new Lagrangian the dimensionless quantity that multiplies the prefactor (m3 +m4)µ2/Lz2 in (3.20),

and we call it Ldumbb.,

Ldumbb. =1

2

(u′

2+ v′

2+ w′

2+

u2 + v2

ρ4(θ − ξ)+ 2

uv′ − u′vρ2(θ − ξ)

)

+1

2(1− δ)δ ε2

(ψ′

2+ φ′

2sin2 ψ

)(3.24)

+ (1− η)

(1− δ‖ρ13‖

‖ρ14‖

)+ η

(1− δ‖ρ23‖

‖ρ24‖

).

The dimensionless distances between the various masses are now ‖ρ13‖, ‖ρ14‖, ‖ρ23‖ and ‖ρ24‖, as

defined in (3.21).

From the dimensionless Lagrangian (3.24), we may thus derive the equations of motion of the dumb-

36

Page 61: Spin-Orbit Interaction and Chaos in Celestial Mechanics

bell in the synodic reference frame, in the new (u, v, w, φ, ψ) coordinates, which are

u′′ − 2v′

ρ2(θ − ξ)=

1

ρ4(θ − ξ)

(u− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

v

)− (1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)(u+ η

× ρ(θ − ξ))

+ (1− δ)δ ε sinψ cos (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))− η

((1− δ‖ρ23‖3

‖ρ24‖3

)

×(u− (1− η) ρ(θ − ξ)

)+ (1− δ)δ ε sinψ cos (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

))

v′′ + 2u′

ρ2(θ − ξ)=

1

ρ4(θ − ξ)

(v + 2e

sin (θ − ξ)1 + e cos (θ − ξ)

u

)− (1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)v

− (1− δ)δ ε sinψ sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))− η

((1− δ‖ρ23‖3

‖ρ24‖3

)v − (1− δ)δ ε

× sinψ sin (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

))

w′′ = − (1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)w + (1− δ)δ ε cosψ

(1

‖ρ13‖3− 1

‖ρ14‖3

))− η

((1− δ‖ρ23‖3

‖ρ24‖3

)w + (1− δ)δ ε cosψ

(1

‖ρ23‖3− 1

‖ρ24‖3

))

ε(φ′′ sin2 ψ + 2φ′ ψ′ sinψ cosψ

)= − (1− η)

((u+ η ρ(θ − ξ)

)sinψ sin (θ − φ) + v sinψ cos (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

)− η((u− (1− η) ρ(θ − ξ)

)sinψ sin (θ − φ) + v sinψ cos (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

)

ε(ψ′′ − φ′2 sinψ cosψ

)= − (1− η)

((u+ η ρ(θ − ξ)

)cosψ cos (θ − φ)− v cosψ sin (θ − φ)− w sinψ

(1

‖ρ13‖3− 1

‖ρ14‖3

)− η((u− (1− η) ρ(θ − ξ)

)cosψ cos (θ − φ)− v cosψ sin (θ − φ)

−w sinψ)( 1

‖ρ23‖3− 1

‖ρ24‖3

),

(3.25)

Several comments are in order concerning the dimensionless equations of motion (3.25). First, the

equations depend on the parameters ε, δ, η, and as well on the eccentricity e, the semi-major axis λ and

the longitude of the pericentre ξ of the orbit of the primary masses about each other. The parameter ε

measures, as stated above, the length of the rod of the dumbbell in units of r∗. For a small satellite like

Nix, ε will be close to zero. As in Section 2.2, the parameter δ ∈ (0, 1) measures the relative weight of

the masses of the dumbbell, while η plays the same role in regard to the primaries as δ does in regard

to m3 and m4. For a symmetric dumbbell with m3 = m4, and for m1 = m2, δ = 1/2 and η = 1/2,

respectively. On the other hand, the mass parameter ν doesn’t appear naturally in the equations. This

37

Page 62: Spin-Orbit Interaction and Chaos in Celestial Mechanics

does not mean however that ν is a redundant parameter, as it is an analogue of µ (Chapter 2), providing

a link between the masses of the primaries and of the dumbbell in the framework of the R3BP. The value

of ν will always be close to unity throughout this paper, in contrast to µ, whose value can range between

zero and unity.

Furthermore, spin-orbit coupling is manifestly present (as it has already been the case in Chapter 2)

in the equations (3.25), since the equations for the motion of the centre of mass of the dumbbell depend

on the attitude angles of the body, and, correspondingly, the equations of motion of the attitude angles

depend on the position of the centre of mass. Likewise, the azimuthal angle φ occurs only in (3.25)

coupled to the orbital angle θ of the primaries through the difference θ − φ.

Lastly, we note that the transformation to the synodic reference frame has introduced terms in the

equations proportional to θ′2u and θ′

2v, the components of the centrifugal acceleration, and θ′v′ and

θ′u′, which are the components of the Coriolis acceleration and depend on the velocity of the centre of

mass of the dumbbell in the rotating frame. There are also terms present which are proportional to θ′′v

and θ′′u. These terms are related to the angular acceleration of the reference frame. When the orbits

of the primaries are circular about their common centre of mass, they vanish, and the equations greatly

simplify. This case — the Circular Restricted Three-Body Problem (CR3BP) — can be obtained from

(3.25) by setting e = 0 and ρ(θ − ξ) = 1.

The dimensionless equations of motion (3.25) are associated to Hamilton’s equations [36]

u′ =∂H∂pu

, v′ =∂H∂pv

, w′ =∂H∂pw

pu′ = −∂H

∂u, pv

′ = −∂H∂v

, pw′ = −∂H

∂w

φ′ =∂H∂pφ

, ψ′ =∂H∂pψ

pφ′ = −∂H

∂φ, pψ

′ = −∂H∂ψ

,

(3.26)

with Hamiltonian

H =1

2

((pu +

v

ρ2(θ − ξ)

)2

+

(pv −

u

ρ2(θ − ξ)

)2

+ pw2 +

1

(1− δ)δ ε2

×(pψ

2 +pφ

2

sin2 ψ

))− 1

2

1

ρ4(θ − ξ)(u2 + v2

)(3.27)

− (1− η)

(1− δ‖ρ13‖

‖ρ14‖

)− η

(1− δ‖ρ23‖

‖ρ24‖

),

which is a function of the dimensionless variables (u, v, w, φ, ψ), the canonical momenta associated to

38

Page 63: Spin-Orbit Interaction and Chaos in Celestial Mechanics

these variables,

pu = u′ − v

ρ2(θ − ξ), pv = v′ +

u

ρ2(θ − ξ), pw = w′

pφ = (1− δ)δ ε2φ′ sin2 ψ, pψ = (1− δ)δ ε2ψ′,

(3.28)

and the time, through θ. It is worth pointing out that H is an effective Hamiltonian, as we are in a

non-inertial reference frame.

If we define a scalar function V = V(u, v, w, φ, ψ; θ) by

V = − 1

2

1

ρ4(θ − ξ)(u2 + v2

)− (1− η)

(1− δ‖ρ13‖

‖ρ14‖

)− η

(1− δ‖ρ23‖

‖ρ24‖

), (3.29)

the equations of motion of the dumbbell (3.25) can also be written in the following way

u′′ − 2v′

ρ2(θ − ξ)= −∂V

∂u− 2e

ρ4(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)v

v′′ + 2u′

ρ2(θ − ξ)= −∂V

∂v+

2e

ρ4(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)u

w′′ = −∂V∂w

(1− δ) δ ε2(φ′′ sin2 ψ + 2φ′ ψ′ sinψ cosψ

)= −∂V

∂φ

(1− δ) δ ε2(ψ′′ − φ′2 sinψ cosψ

)= −∂V

∂ψ.

(3.30)

The function V may be referred to as an effective potential. The term proportional to u2 + v2 in V is the

centrifugal potential, while the terms proportional to the inverse of the distances between the masses

constitute the gravitational potential. Let us multiply the first equation in (3.30) by u′, the second equation

by v′, the third by w′, the fourth by φ′, the fifth by ψ′, and add all of them. We get

u′u′′ + v′v′′ + w′w′′ + (1− δ) δ ε2(ψ′ψ′′ + φ′φ′′ sin2 ψ + φ′

2ψ′ sinψ cosψ

)=

−(∂V∂u

u′ +∂V∂v

v′ +∂V∂w

w′ +∂V∂φ

φ′ +∂V∂ψ

ψ′)

+2e

ρ4(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)(uv′ − u′v) , (3.31)

or,

d

[1

2

(u′

2+ v′

2+ w′

2+ (1− δ) δ ε2

(ψ′

2+ φ′

2sin2 ψ

))]+

dVdτ

=∂V∂τ

+2e

ρ4(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)(uv′ − u′v) . (3.32)

The left hand side of (3.32) is just the total time derivative of H, now viewed as a function of the di-

39

Page 64: Spin-Orbit Interaction and Chaos in Celestial Mechanics

mensionless coordinates and their velocities. The first term on the right hand side of (3.32) is equal

to

∂V∂τ

=∂V∂θθ′ =

2e

ρ6(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)(u2 + v2

)+ (1− η)

{1− δ‖ρ13‖3

(e η

ρ(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)

((u+ η ρ(θ − ξ)

)+ δ ε sinψ cos (θ − φ)

)− δ ε sinψ

ρ2(θ − ξ)

((u+ η ρ(θ − ξ)

)sin (θ − φ)

+ v cos (θ − φ)))

‖ρ14‖3

(e η

ρ(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)

((u

+ η ρ(θ − ξ))− (1− δ) ε sinψ cos (θ − φ)

)+ (1− δ) ε sinψ

ρ2(θ − ξ)

×((u+ η ρ(θ − ξ)

)sin (θ − φ) + v cos (θ − φ)

))}+ η

{1− δ‖ρ23‖3

(− e (1− η)

ρ(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)

((u− (1− η)

× ρ(θ − ξ))

+ δ ε sinψ cos (θ − φ))− δ ε sinψ

ρ2(θ − ξ)

((u− (1− η)

× ρ(θ − ξ))

sin (θ − φ) + v cos (θ − φ)))

‖ρ24‖3

(− e (1− η)

ρ(θ − ξ)sin (θ − ξ)

1 + e cos (θ − ξ)

((u− (1− η) ρ(θ − ξ)

)− (1− δ) ε sinψ cos (θ − φ)

)+ (1− δ) ε sinψ

ρ2(θ − ξ)

((u− (1− η)

× ρ(θ − ξ))

sin (θ − φ) + v cos (θ − φ)))}

. (3.33)

Hence, this implies that, contrary to the conventional CR3BP where H is a constant of the motion —

the so-called Jacobi integral ([28]) — H is not conserved in the R3BP with dumbbell satellite. Moreover,

even when we consider the CR3BP with dumbbell satellite, for which e = 0 and ρ(θ − ξ) = 1, there

cannot in general be a Jacobi integral, since there are still nonvanishing terms in ∂V/∂τ which aren’t

proportional to e:

∂V∂τ

∣∣∣∣ρ(θ−ξ)=1e=0

= (1− η) (1− δ) δ ε sinψ((u+ η

)sin (θ − φ) + v cos (θ − φ)

)

×

(1

‖ρ14‖3− 1

‖ρ13‖3

)

+ η (1− δ) δ ε sinψ((u− (1− η)

)sin (θ − φ) + v cos (θ − φ)

(1

‖ρ24‖3− 1

‖ρ23‖3

), (3.34)

where ‖ρ13‖2, ‖ρ14‖2, ‖ρ23‖2 and ‖ρ24‖2 are as given in (3.21), with ρ(θ − ξ) = 1. The following are

sufficient conditions for ∂V/∂τ to become equal to zero and, consequently, for H to be a constant of the

40

Page 65: Spin-Orbit Interaction and Chaos in Celestial Mechanics

motion in the CR3BP with dumbbell satellite:

ψ = 0, π (3.35)

‖ρ13‖ = ‖ρ14‖ and ‖ρ23‖ = ‖ρ24‖ (3.36)(u+ η

)sin (θ − φ) + v cos (θ − φ) = 0(

u− (1− η))

sin (θ − φ) + v cos (θ − φ) = 0

(3.37)

‖ρ13‖ = ‖ρ24‖ and ‖ρ14‖ = ‖ρ23‖

η = 1/2 and θ − φ = 0, π

(3.38)

‖ρ13‖ = ‖ρ23‖ and ‖ρ14‖ = ‖ρ24‖

u sin (θ − φ) = −v cos (θ − φ) .

(3.39)

From (3.35) we see that whenever the motion of the dumbbell is such that it is perpendicularly oriented to

the Lagrange plane of the primaries at all times, if such motion exists, H will be a conserved quantity for

the CR3BP. For condition (3.36) to happen, it is mandatory that sinψ cos (θ − φ) = 0 holds, independently

of δ, and, specifically for equal dumbbell masses, w cosψ = v sinψ sin (θ − φ) is also verified. Condition

(3.37) is equivalent to v = 0 and θ − φ = 0, π, and it corresponds to a hypothetical motion of the

dumbbell in which it is constrained to move and rotate in the instantaneous plane which contains both

m1 and m2 and is orthogonal to the Lagrange plane. Condition (3.38) only holds for equal masses of the

primaries and 2u = ± (1− 2δ) ε sinψ, and, finally, condition (3.39) takes place for η = 1/2 and u = 0 and

θ − φ = ±π/2, hence the dumbbell is constrained to move and rotate in the instantaneous mediating

plane of the line segment connecting the primaries. More restrictive conditions include, for instance,

all four distances, ‖ρ13‖, ‖ρ14‖, ‖ρ23‖ and ‖ρ24‖, equal. In conclusion, apart from the distance ‖r34‖,

which coincides with the length of the dumbbell and is obviously a conserved quantity, there are in

general no constants of the motion for the (C)R3BP with dumbbell satellite, because neither energy nor

angular momentum is conserved.

3.4 Steady states and other solutions

The equations of motion of the dumbbell in the synodic reference frame (3.25) are still cumbersome, so

we will focus here solely on the study of two particular cases, the Planar Circular Restricted Three-Body

Problem (PCR3BP) with dumbbell satellite and the Sitnikov Problem (SP), Sitnikov [27]. In the present

section we derive the equations of motion for these two systems and we do a systematic analysis of the

solutions of those equations, including periodic or quasi-periodic motions.

It is convenient to adopt the angular coordinate θ as independent variable instead of τ . This amounts

41

Page 66: Spin-Orbit Interaction and Chaos in Celestial Mechanics

to applying the following transformation properties4,

d

dτ=

1

ρ2d

d2

dτ2=

1

ρ4

(d2

dθ2− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

d

), (3.40)

to (3.25), which then transforms into

u− 2

(v + e

sin (θ − ξ)1 + e cos (θ − ξ)

(u− v)

)= u− ρ4

((1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)(u+ η ρ

)+ (1− δ)δ ε

× sinψ cos (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))

+ η

((1− δ‖ρ23‖3

‖ρ24‖3

)(u− (1− η) ρ

)+ (1− δ)δ ε sinψ cos (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

)))

v + 2

(u− e sin (θ − ξ)

1 + e cos (θ − ξ)(v + u)

)= v − ρ4

((1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)v − (1− δ)δ ε sinψ

× sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))

+ η

((1− δ‖ρ23‖3

‖ρ24‖3

)v − (1− δ)δ ε sinψ sin (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

)))

w − 2esin (θ − ξ)

1 + e cos (θ − ξ)w = −ρ4

((1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)w + (1− δ)δ ε cosψ

(1

‖ρ13‖3

− 1

‖ρ14‖3

))+ η

((1− δ‖ρ23‖3

‖ρ24‖3

)w + (1− δ)δ ε cosψ

(1

‖ρ23‖3− 1

‖ρ24‖3

)))

ε

((φ− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

φ

)sin2 ψ + 2 φ ψ sinψ cosψ

)= −ρ4

((1− η)

((u+ η ρ

)sinψ sin (θ − φ)

+ v sinψ cos (θ − φ))( 1

‖ρ13‖3− 1

‖ρ14‖3

)

+ η((u− (1− η) ρ

)sinψ sin (θ − φ) + v sinψ cos (θ − φ)

)( 1

‖ρ23‖3− 1

‖ρ24‖3

))

ε

(ψ − 2e

sin (θ − ξ)1 + e cos (θ − ξ)

ψ − φ2 sinψ cosψ

)= −ρ4

((1− η)

((u+ η ρ

)cosψ cos (θ − φ)− v cosψ

× sin (θ − φ)− w sinψ)( 1

‖ρ13‖3− 1

‖ρ14‖3

)

+ η((u− (1− η) ρ

)cosψ cos (θ − φ)− v cosψ sin (θ − φ)− w sinψ

)( 1

‖ρ23‖3− 1

‖ρ24‖3

)),

(3.41)

Here the variables (u, v, w, φ, ψ) are to be understood as functions of θ instead of τ . The point (·) now

4For notational convenience, from now on we will omit the dependence of ρ on (θ − ξ).

42

Page 67: Spin-Orbit Interaction and Chaos in Celestial Mechanics

symbolises the derivative with respect to θ instead of t.

3.4.1 PCR3BP with dumbbell satellite

As pointed out in Section 3.3, the systems of equations (3.25) or (3.41) have no constants of the motion

or first integrals, hence they are called nonintegrable, in the sense that one cannot obtain a general

solution to these systems of equations in closed form. We thus make a further simplification to (3.41),

by confining the motion of the centre of mass of the dumbbell to the plane of the orbits of the primaries,

while assuming that the motion of m1 and m2 is circular about their common barycentre. The dumbbell

itself is still allowed to rotate around its centre of mass in three-dimensional space. This is the PCR3BP

with dumbbell satellite. The PCR3BP is an adequate starting model for many phenomena in celestial

mechanics, including the modelling of comets interacting with Jupiter and the Sun, whose motion is

very close to Jupiter’s orbital plane, or the design of spacecraft trajectories that take into account the

gravitational field of several bodies. For a comprehensive review of the applications of this model, see

[48].

This procedure effectively reduces the dimension of the configuration space of the dumbbell by one,

becoming R2 × S2, and, accordingly, the dimension of the phase space by two. A configuration of the

dumbbell is now fully specified by u, v and the attitude angles on the sphere, φ and ψ. The configuration

space of both primaries is S1. Imposing w = w = w ≡ 0 and setting e = 0 and ρ = 1 in (3.21) and

(3.41), we arrive at the equations for the PCR3BP with dumbbell satellite in the synodic reference frame,

in dimensionless form,

u− 2v = u− (1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)(u+ η

)+ (1− δ)δε sinψ cos (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))

− η

((1− δ‖ρ23‖3

‖ρ24‖3

)(u− (1− η)

)+ (1− δ)δε sinψ cos (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

))

v + 2u = v − (1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)v − (1− δ)δε sinψ sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))

− η

((1− δ‖ρ23‖3

‖ρ24‖3

)v − (1− δ)δε sinψ sin (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

))

ε(φ sin2 ψ + 2 φ ψ sinψ cosψ

)= − (1− η)

((u+ η

)sinψ sin (θ − φ) + v sinψ cos (θ − φ)

)( 1

‖ρ13‖3

− 1

‖ρ14‖3

)− η((u− (1− η)

)sinψ sin (θ − φ) + v sinψ cos (θ − φ)

)( 1

‖ρ23‖3− 1

‖ρ24‖3

)

ε(ψ − φ2 sinψ cosψ

)= − (1− η)

((u+ η

)cosψ cos (θ − φ)− v cosψ sin (θ − φ)

)( 1

‖ρ13‖3

− 1

‖ρ14‖3

)− η((u− (1− η)

)cosψ cos (θ − φ)− v cosψ sin (θ − φ)

)( 1

‖ρ23‖3− 1

‖ρ24‖3

),

(3.42)

43

Page 68: Spin-Orbit Interaction and Chaos in Celestial Mechanics

where the squared distances between the primaries and the masses of the dumbbell are given by

‖ρ13‖2 =(u+ η

)2+ v2 + δ2 ε2 + 2 δ ε sinψ

((u+ η

)cos (θ − φ)− v sin (θ − φ)

)‖ρ14‖2 =

(u+ η

)2+ v2 + (1− δ)2 ε2 − 2 (1− δ) ε sinψ

((u+ η

)cos (θ − φ)− v sin (θ − φ)

)‖ρ23‖2 =

(u− (1− η)

)2+ v2 + δ2 ε2 + 2 δ ε sinψ

((u− (1− η)

)cos (θ − φ)− v sin (θ − φ)

)‖ρ24‖2 =

(u− (1− η)

)2+ v2 + (1− δ)2 ε2 − 2 (1− δ) ε sinψ

((u− (1− η)

)cos (θ − φ)− v sin (θ − φ)

).

(3.43)

Finally, we point out that in this model the masses m1 and m2 are at rest in the synodic reference

frame, with positions given by (−η, 0, 0) and(

(1− η) , 0, 0), respectively. Therefore, they are always at

unit distance from one another in this system. In Figure 3.4, we depict a possible configuration of the

dumbbell and the primaries in the PCR3BP with dumbbell satellite.

Figure 3.4: Schematic representation of a configuration of the dumbbell and the primaries in thePCR3BP with dumbbell satellite, in the inertial frame S, at some arbitrary instant of time. Here wehave used η = δ = 1/2, so the dumbbell has equal masses, and the primaries have equal masses too.The variables (U, V,W ) are dimensionless, cartesian coordinates in the inertial frame. The dumbbellis aligned perpendicularly to the Lagrange plane of the primaries, while the latter share the same orbitaround their common barycentre.

We study now the equations (3.42). Specifically, we want to apply the theory described in this section

to the Pluto-Charon-Nix system. Since it has been shown by numerical integration that the motion of Nix

is nearly in the plane of the orbits of Pluto and Charon about their common barycentre ([45]), and since

the eccentricities of these orbits are almost negligible ([42]), if we assume that the shape of Nix, like so

many asteroids in the solar system, is that of a dumbbell ([49]), then the PCR3BP with dumbbell satellite

defined by equations (3.42) and (3.43) is a suitable candidate for studying such a system. The system

of equations (3.42) is subjected to a restriction which follows from the vanishing of the left-hand side of

the equation in w in (3.41),

(1− η)ε cosψ

(1

‖ρ13‖3− 1

‖ρ14‖3

)= −η ε cosψ

(1

‖ρ23‖3− 1

‖ρ24‖3

). (3.44)

44

Page 69: Spin-Orbit Interaction and Chaos in Celestial Mechanics

The condition (3.44) is a necessary condition for the motion of the centre of mass of the dumbbell to

be confined to the Lagrange plane of the primaries. It is thus mandatory that all solutions to (3.42)

and (3.43) also verify (3.44), otherwise the corresponding motion cannot occur. In the following we do a

systematic analysis of (3.44) in conjunction with (3.42) and (3.43), in order to derive the possible motions

for the dumbbell in the context of the PCR3BP.

All the conditions that verify equation (3.44) fall into one of the following four main categories:

category 1: ‖ρ13‖ = ‖ρ14‖ and ‖ρ23‖ = ‖ρ24‖ ( 6= ‖ρ13‖ , ‖ρ14‖) ;

category 2: ‖ρ13‖ = ‖ρ24‖ and ‖ρ14‖ = ‖ρ23‖ ( 6= ‖ρ13‖ , ‖ρ24‖)

and η =1

2; (3.45)

category 3: ‖ρ13‖ = ‖ρ14‖ = ‖ρ23‖ = ‖ρ24‖ (all 4 distances equal) ;

category 4: ‖ρ13‖ 6= ‖ρ14‖ 6= ‖ρ23‖ 6= ‖ρ24‖ (all 4 distances different) .

Category 1

The case ε = 0, for which the dumbbell reduces to a single point mass, obviously falls under this category

too in general. We don’t pursue this, since it has already been extensively researched in the literature

(see, for instance, Szebehely [3]), and we are solely interested in the study of the dumbbell dynamics.

There are two cases that fall under this category for ε > 0. By inspection of (3.43), we see that we

can make ‖ρ13‖ = ‖ρ14‖ and simultaneously ‖ρ23‖ = ‖ρ24‖ by setting δ = 1/2 (m3 = m4) and either

ψ = 0, π or

(u+ η) cos (θ − φ) = v sin (θ − φ) and(u− (1− η)

)cos (θ − φ) = v sin (θ − φ) , (3.46)

which implies (u+η) cos (θ − φ) =(u−(1− η)

)cos (θ − φ). This equation has the solutions θ−φ = ±π/2,

which, on substitution in (3.46), give v ≡ 0. For the case

δ =1

2and ψ = 0, π ,

the equations governing the motion of the dumbbell become, from (3.42),

u− 2v = u− (1− η)u+ η

‖ρ13‖3− η u− (1− η)

‖ρ23‖3

v + 2u =

(1− 1− η‖ρ13‖3

− η

‖ρ23‖3

)v

ψ = 0 ,

(3.47)

where ‖ρ13‖ = ‖ρ14‖ =

√(u+ η

)2+ v2 + ε2/4 and ‖ρ23‖ = ‖ρ24‖ =

√(u− (1− η)

)2+ v2 + ε2/4 .

Therefore, this type of motion has the dumbbell always pointing in the orthogonal direction to the La-

45

Page 70: Spin-Orbit Interaction and Chaos in Celestial Mechanics

grange plane, with ψ = const., and only exists for equal masses of the dumbbell. The case

δ =1

2and v ≡ 0 and θ − φ = ±π/2

reduces the equations of motion (3.42) to

u− (1− η)u+ η

‖ρ13‖3− η u− (1− η)

‖ρ23‖3= 0

u = 0

ψ sinψ cosψ = 0

ψ − sinψ cosψ = 0 ,

(3.48)

with ‖ρ13‖ = ‖ρ14‖ =

√(u+ η

)2+ ε2/4 and ‖ρ23‖ = ‖ρ24‖ =√(

u− (1− η))2

+ ε2/4 , since, from θ − φ = ±π/2, φ ≡ 1 and φ ≡ 0. The first two equations of

(3.48) tell us that u = u(θ0) is a constant (θ0 is the initial value of the independent variable θ) and that it

is a root of the function h : R→ R defined by

h(u; η, ε) = u− (1− η)u+ η

‖ρ13‖3− η u− (1− η)

‖ρ23‖3, (3.49)

where the distances are as defined immediately above. That h has a root is easily seen. In fact, the

value of h at u = 0, given by

h(u = 0; η, ε) = 8 (1− η) η

1(ε2 + 4 (1− η)

2) 3

2

− 1(ε2 + 4η2

) 32

, (3.50)

is negative for 0 < η < 1/2, zero for η = 1/2 and positive for 1/2 < η < 1, whatever the value of ε > 0 (in

our applications ε will always be much smaller than unity). This, combined with the continuity of h for all

u ∈ R and the fact that limu→−∞ h(u; η, ε) = −∞ and limu→+∞ h(u; η, ε) = +∞ , ensures that there is

always at least one positive root of h for the interval 0 < η < 1/2 or one negative root for 1/2 < η < 1.

On the other hand, when η = 1/2, u = 0 is a solution. Indeed, one may verify via graphical analysis

of the function h that the equation h(u; η, ε) = 0 can have 1, 2, 3, 4, or even 5 solutions, according

to the values of the parameters η and ε. Additionally, from the last two equations in (3.48), it follows

immediately that ψ = const. for all θ ∈ R and that it can only take the values 0, π or π/2. As such, the

present case corresponds to equilibrium solutions of the system of equations (3.42). These solutions

are defined by the following conditions:

u = u(η, ε); v = 0; θ − φ = ±π/2; ψ = 0, π, π/2;

u = v = ψ = 0; φ = 1; δ = 1/2.(3.51)

46

Page 71: Spin-Orbit Interaction and Chaos in Celestial Mechanics

They are the direct analogues of the Lagrangian points L1, L2 and L3 in the (conventional) CR3BP

[44, 3]. For every pair of values (η, ε), we may have several possible motions, depending on the values

taken by u, ψ or θ−φ, but only a subset of them is truly distinguishable, since the masses of the dumbbell

are equal, m3 = m4. All of the solutions (3.51) are characterized by a circular motion of radius u of the

centre of mass of the dumbbell about the common barycentre of the primaries in inertial space. These

are thus steady states of (3.42). We note that u is a function of the parameters η and ε, given implicitly

by h(u; η, ε) = 0. It is easy to verify via graphical analysis of h that u can take on values smaller or

greater than the radii of the orbits of m1 and m2, hence the dumbbell can be found either in-between the

primaries or outside that region. Since θ − φ = ±π/2, the dumbbell rotates in the same direction of the

translational motion of the primaries, always at right angles to the line segment connecting m1 to m2.

There is thus a 1 : 1 synchronisation of the translational motion of the dumbbell with its rotational motion

and the translational motion of the primaries. When ψ = π/2, the dumbbell is constrained to rotate in

the Lagrange plane, and when ψ = 0, π, it is aligned orthogonally to this plane (this is a steady solution

of (3.47)). Figure 3.5 illustrates two of the steady states (3.51) in the inertial S (Figures 3.5a, 3.5b) and

synodic R (Figures 3.5c, 3.5d) reference frames, for ψ = π/2 and θ − φ = π/2.

(a) (b)

(c) (d)

Figure 3.5: Two steady states of the PCR3BP with dumbbell satellite represented in the inertial S andsynodic R frames at θ = 3π/5 ((a), (c)) and at θ = 5π/4 ((b), (d)), for which the dumbbell lies in theLagrange plane (ψ = π/2) and maintains a right angle to the line that joins the primaries (θ − φ = π/2).The length of the dumbbell is ε = 0.15, and we have used η = 0.2 and δ = 0.5 across all figures.The dashed lines show the paths of the primaries in inertial space, whereas the solid line illustratesthe trajectory of the dumbbell’s centre of mass. (a) The centre of mass of the dumbbell describes acircular trajectory in the region between the primaries (u ≈ 0.44); (b) the centre of mass of the dumbbelldescribes a circular trajectory in the region outside the primaries (u ≈ −1.08); (c) same as (a), depictedin the synodic reference frame R; (d) same as (b), depicted in R.

47

Page 72: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Category 2

We set u = 0 and δ = 1/2 in addition to η = 1/2, in order to satisfy (3.44) with ‖ρ13‖ = ‖ρ24‖ and

‖ρ14‖ = ‖ρ23‖. Plugging these values into (3.43) we retrieve the condition ε v sinψ sin (θ − φ) = 0. This

equation is satisfied for ψ = 0, π or v = 0 or θ − φ = 0, π, provided ε is nonvanishing, which is the case

under study. The case ψ = 0, π and u = 0 and η = δ = 1/2 is seen, by inspection of (3.43), to fall under

category 3 of all four distances equal. It is a hypothetical motion associated to (3.47), and we will come

to its analysis ahead. For now, let us focus on the other two cases, which both fall under the category in

study.

The first case is

η = δ =1

2and u = v = 0 .

This solution of (3.42) has the centre of mass of the dumbbell placed at the centre of mass of the

primaries m1 and m2, and it exists only for equal masses of the primaries and equal masses of the

dumbbell. The rotation of the dumbbell in configuration space is governed by the equations

ε(φ sin2 ψ + 2 φ ψ sinψ cosψ

)= −1

2sinψ sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

)

ε(ψ − φ2 sinψ cosψ

)= −1

2cosψ cos (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

),

(3.52)

where

‖ρ13‖2 = ‖ρ24‖2 = 1/4(1 + ε2 + 2 ε sinψ cos (θ − φ)

)‖ρ14‖2 = ‖ρ23‖2 = 1/4

(1 + ε2 − 2 ε sinψ cos (θ − φ)

).

The equations (3.52) and (3.42) admit the following equilibrium solutions or steady states

u = v = 0; θ − φ = ±π/2; ψ = π/2;

u = v = ψ = 0; φ = 1; η = δ = 1/2;

u = v = 0; ψ = 0, π;

u = v = ψ = 0; η = δ = 1/2 ,

(3.53)

for which all four distances are equal to√

1 + ε2/2, and, thus, fall under the third category of solutions.

In the steady state for which ψ = π/2, the dumbbell is allowed to rotate only in the Lagrange plane and,

in view of the constant difference θ − φ = ±π/2, it rotates in inertial space in the same direction of the

translational motion of the primaries, maintaining at all times a right angle to the line that connects them.

This means that there is a 1 : 1 synchronisation between the rotation of the dumbbell and the translation

of the primaries. The dumbbell always shows the same face to each one of the primaries. For the

other steady state, corresponding to ψ = 0, π, the dumbbell maintains an orthogonal configuration with

respect to the Lagrange plane at the barycentre of m1 and m2. Nothing can be said about the rotation

48

Page 73: Spin-Orbit Interaction and Chaos in Celestial Mechanics

in φ. In Figure 3.6 we illustrate the two kinds of steady states.

(a) (b)

(c) (d)

Figure 3.6: Steady states (3.53) of the PCR3BP with dumbbell satellite represented in the inertial S ((a),(b)) and synodic R ((c), (d)) frames at θ = 5π/6, for ε = 0.15. (a), (c) The dumbbell lies in the Lagrangeplane (ψ = π/2) and maintains a right angle to the line that joins the primaries (θ − φ = π/2) (alsorepresented in (a) is the trajectory in inertial space of the masses of the dumbbell); (b), (d) the dumbbellis lined up with the W axis (ψ = 0).

The second case is

η = δ =1

2and u = 0 and θ − φ = 0, π .

Substituting these conditions into (3.42) and (3.43), it leads to the equations

v = 0

v

(1− 1

2

(1

‖ρ13‖3+

1

‖ρ14‖3

))= 0

ψ sinψ cosψ = 0

ε(ψ − sinψ cosψ

)= ∓1

2cosψ

(1

‖ρ13‖3− 1

‖ρ14‖3

),

(3.54)

where

‖ρ13‖2 = ‖ρ24‖2 = 1/4(1 + 4v2 + ε2 ± 2 ε sinψ

)49

Page 74: Spin-Orbit Interaction and Chaos in Celestial Mechanics

‖ρ14‖2 = ‖ρ23‖2 = 1/4(1 + 4v2 + ε2 ∓ 2 ε sinψ

).

Here the upper sign conforms to θ − φ = 0, while the lower sign corresponds to θ − φ = π. The first and

third equations in (3.54) imply that v and ψ are constant for all θ ∈ R, respectively. This in turn leads

to the conclusion that this case corresponds to steady state solutions of (3.42) for η = δ = 1/2. The

second equation in (3.54) has the trivial solution v = 0. Furthermore, it shows that, in general, v is a root

of the function m : R→ R defined by

m(v; ε, ψ) =1

(1 + 4v2 + ε2 + 2 ε sinψ)32

+1

(1 + 4v2 + ε2 − 2 ε sinψ)32

− 1

4, (3.55)

where ψ may also be seen as a parameter, since we have already determined that it is constant. The

function m is the same for both cases, θ − φ = 0 and θ − φ = π, and it implicitly defines the solution

v = v(ε, ψ), if it exists, as a function of ε and the polar angle ψ. We show that m has exactly two

symmetric roots. To do that, let us first collect some of its properties:

i m(−v; ε, ψ) = m(v; ε, ψ) i.e. m is an even function of v;

ii limv→±∞m(v; ε, ψ) = −1/4;

iii m reaches its maximum value for v = 0. In case ε = 1 and ψ = π/2, v = 0 is a singular point of m.

The first of these properties demonstrates that we only have to analyse the behaviour of m in the interval

[0,+∞). If the absolute maximum of m at v = 0 is positive, then, since m is a continuous, monotonically

decreasing function in that interval, there will be a unique root in [0,+∞). Considering that m is even in

v, there will be another, symmetric root in (−∞, 0]. All that is left doing then is to show that, in fact, the

maximum at v = 0 is positive i.e.

m(v = 0; ε, ψ) =1

(1 + ε2 + 2 ε sinψ)32

+1

(1 + ε2 − 2 ε sinψ)32

− 1

4> 0 . (3.56)

The value of the maximum at v = 0, m(v = 0; ε, ψ), is a function of ψ and ε. The partial derivative of this

function with respect to ψ is

∂ψm(v = 0; ε, ψ) = 3ε cosψ

(1

(1 + ε2 − 2 ε sinψ)52

− 1

(1 + ε2 + 2 ε sinψ)52

), (3.57)

whose roots are ψ = 0, π and ψ = π/2. These are the critical points of m(v = 0; ε, ψ). By continuing this

process further and calculating the second derivative of this function with respect to ψ, one may easily

verify that m(v = 0; ε, ψ) has minimums for ψ = 0, π and a maximum for ψ = π/2. This means that the

maximum of m at v = 0 takes its lowest value when either ψ = 0 or ψ = π. In other words, ψ = 0, π is

a worst-case scenario. We thus calculate m(v = 0; ε; ψ = 0, π) and derive the condition that ensures

the maximum is still positive in the worst-case scenario. It turns out that m(v = 0; ε; ψ = 0, π) =

2/(1 + ε2

)3/2 − 1/4, and this is positive if and only if |ε| <√

3. Since in our applications ε is always

smaller than unity, we have demonstrated the following result: there are always two more equilibrium

50

Page 75: Spin-Orbit Interaction and Chaos in Celestial Mechanics

solutions for the present case with symmetric v coordinates, besides the one for which v = 0. This is

independent of the value of ψ at the fixed point, and is valid in both cases θ − φ = 0 and θ − φ = π.

One can readily verify that ψ = 0, π and ψ = π/2 satisfy the last equation in (3.54). Therefore, (3.54)

and (3.42) admit the following equilibrium solutions or steady states

u = 0; v = 0, ± v(ε, ψ = π/2); θ − φ = 0, π; ψ = π/2;

u = v = ψ = 0; φ = 1; η = δ = 1/2;

u = 0; v = 0, ± v(ε, ψ = 0, π); θ − φ = 0, π; ψ = 0, π;

u = v = ψ = 0; φ = 1; η = δ = 1/2 .

(3.58)

Here ± v(ε, ψ) denotes the two symmetric roots of m. These are the direct analogues of the Lagrangian

equilateral points L4 and L5 in the conventional CR3BP [44, 3]. For the last of these two steady states,5

since ψ = 0, π, the dumbbell is aligned orthogonally to the Lagrange plane, and all the four distances

are in fact equal to√

1 + 4v2 + ε2/2, so it will fall under category 3. The motion of the centre of mass of

the dumbbell in the inertial reference frame is different according to whether v is zero or not. If v = 0,

the dumbbell is at rest at the centre of mass of the primaries, in correspondence with the last of the

steady states (3.53). Otherwise, since v = const. in the synodic reference frame, the centre of mass of

the dumbbell describes a circular trajectory of radius v(ε) ≡ v(ε, ψ = 0, π) in inertial space around this

point, much like one of the cases in (3.51). If that is the case, there will be a 1 : 1 synchronisation be-

tween the translational motions of the dumbbell and the primaries. In passing we note that the equation

m(v; ε, ψ) = 0 is actually solvable when ψ = 0, π, because all the distances are equal then. The result

is v(ε, ψ = 0, π) =√

3− ε2/2, and this is the radius of the orbit. From the topological point of view, both

solutions with v = + v(ε, ψ) and v = − v(ε, ψ) are equivalent in the inertial space.

For the steady states with ψ = π/2, the dumbbell is confined to rotate in the Lagrange plane of the

primaries, and, in view of θ − φ being equal to 0 or π, it points at all times in the direction of the line

segment connecting the primaries. The motion of the centre of mass of the dumbbell in the inertial

reference frame is exactly as stated above for the case ψ = 0, π. There are 1 : 1 synchronisations for

these solutions too, in accordance with the fact that the masses are all at rest in the synodic reference

frame. If v = 0, the synchronisation happens between the rotational motion of the dumbbell and the

translational motions of m1 and m2. If v = ± v(ε, ψ = π/2), the translational motion of the dumbbell

additionally synchronises with the other two periodic motions. Figure 3.7 depicts some of the steady

states (3.58).

5There are actually 18 steady states in (3.58), although not all of them are distinguishable, because η = δ = 1/2. “Two steadystates” refers only to the manner in which they are grouped, according to the category to which they belong. In the case understudy, some belong to category 2, others to 3.

51

Page 76: Spin-Orbit Interaction and Chaos in Celestial Mechanics

(a) (b)

(c) (d)

Figure 3.7: Steady states (3.58) of the PCR3BP with dumbbell satellite, in which the dumbbell lies in theLagrange plane (ψ = π/2) and is aligned along the direction of the segment line that joins the primaries(θ = φ), represented in the inertial S ((a), (b)) and synodic R ((c), (d)) reference frames. We have usedε = 0.15. (a), (c) v = 0: the dumbbell and primaries are depicted at θ = 4π/3 (also represented in (a) isthe trajectory of the masses of the dumbbell in the inertial space); (b), (d) v(ε = 0.15, ψ = π/2) ≈ 0.867:the dumbbell and primaries are depicted at θ = 7π/4. This steady state is an extension of the equilateralLagrangian points of the CR3BP ([44, 3]) to the case where the satellite has a dumbbell shape.

Category 3

We require that all four distances, ‖ρ13‖, ‖ρ14‖, ‖ρ23‖ and ‖ρ24‖, are equal. Equating ‖ρ13‖2 to ‖ρ23‖2,

and, separately, ‖ρ14‖2 to ‖ρ24‖2 in (3.43), one gets

(u+ η

)2=(u− (1− η)

)2 − 2δ ε sinψ cos (θ − φ) (3.59)(u+ η

)2=(u− (1− η)

)2+ 2 (1− δ) ε sinψ cos (θ − φ) . (3.60)

Subtracting (3.60) from (3.59), it follows that

−2δ ε sinψ cos (θ − φ) = 2 (1− δ) ε sinψ cos (θ − φ) ,

or sinψ cos (θ − φ) = 0 for positive ε. Substituting back into equations (3.59)– (3.60), the next condition

holds (u+ η

)2=(u− (1− η)

)2, (3.61)

with solution u+η = 1/2. This demonstrates that for all four distances to be equal, it is necessary that the

centre of mass of the dumbbell moves along the perpendicular bisector of the line segment connecting

the primaries. The coordinate u of C ′ in the synodic reference frame is then fixed by the value of η.

52

Page 77: Spin-Orbit Interaction and Chaos in Celestial Mechanics

On the other hand, equating ‖ρ13‖2 to ‖ρ14‖2 and ‖ρ24‖2 in (3.43), one obtains the condition

δ2ε2 − 2δ ε v sinψ sin (θ − φ) = (1− δ)2 ε2 + 2 (1− δ) ε v sinψ sin (θ − φ) ,

or, equivalently,

ε2 (2δ − 1) = 2ε v sinψ sin (θ − φ) , (3.62)

where we have already substituted the values for both u + η and sinψ

× cos (θ − φ). The equation (3.62) shows that the quantity v sinψ sin (θ − φ) is a constant of the mo-

tion when all four distances are equal, whose value is ε (δ − 1/2).

The cases under this category thus satisfy the following properties:

sinψ cos (θ − φ) = 0 and u =

1

2− η

ε2 (2δ − 1) = 2ε v sinψ sin (θ − φ) .

We have two possibilities: either ψ = 0, π or θ− φ = ±π/2. For ψ = 0, π, v sinψ sin (θ − φ) is identically

zero and therefore δ = 1/2, corresponding to the case

ψ = 0, π and u =1

2− η and δ =

1

2. (3.63)

The equations of motion for (3.63) are, by (3.42),u− 2v = u

(1− 1

‖ρ13‖3

)

v = v

(1− 1

‖ρ13‖3

),

(3.64)

where ‖ρ13‖ = ‖ρ14‖ = ‖ρ23‖ = ‖ρ24‖ =√

1 + 4v2 + ε2/2 from (3.43), and, of course, ψ = 0. They

govern the motion of the centre of mass of the dumbbell along the perpendicular bisector of the line that

connects m1 to m2. The equations (3.64) and (3.42) take the following equilibrium solutions or steady

statesu = v = 0; ψ = 0, π;

u = v = ψ = 0; η = δ = 1/2 ,(3.65)

for equal masses of the primaries, and

u =1

2− η; v = ± v(ε, ψ = 0, π); ψ = 0, π;

u = v = ψ = 0; δ = 1/2 ,(3.66)

for any value of η (and ε <√

3). We have already encountered the steady states (3.65) in (3.53) and

(3.58). While we knew from equation (3.58) that the fixed points (3.66) exist for η = 1/2 and u = 0,

equation (3.66) tells us that they also exist more generally for different masses of the primaries and

u = 1/2− η. Here, as before, v(ε, ψ = 0, π) =√

3− ε2/2. Figure 3.8 depicts one of these steady states,

53

Page 78: Spin-Orbit Interaction and Chaos in Celestial Mechanics

for η = 0.2 and ε = 0.15. To conclude, we note that H is a constant of the motion for (3.64), because

ψ = 0, π for the case (3.63), in accordance with (3.34)–(3.35).

(a)

(b)

Figure 3.8: Steady state (3.66) of the PCR3BP with dumbbell satellite, for η = 0.2 and ε = 0.15, repre-sented in the inertial S (a) and synodic R (b) reference frames at θ = 5π/3. The dumbbell is alignedperpendicularly to the Lagrange plane (ψ = 0). This is another steady state of the “equilateral type”, thistime for different masses of the primaries. This steady state only occurs for orthogonal configurations ofthe dumbbell satellite with respect to the Lagrange plane.

For θ − φ = ±π/2, we have the second case

θ − φ = ±π/2 and u =1

2− η and ε2 (2δ − 1) = ± 2ε v sinψ , (3.67)

where, as usual, the upper sign denotes θ−φ = π/2, while the lower sign corresponds to θ−φ = −π/2.

The equations of motion (3.42) transform, for (3.67), into

u− 2v = u

(1− 1

‖ρ13‖3

)

v = v

(1− 1

‖ρ13‖3

)ψ sinψ cosψ = 0

ψ − sinψ cosψ = 0 ,

(3.68)

54

Page 79: Spin-Orbit Interaction and Chaos in Celestial Mechanics

with ‖ρ13‖ = ‖ρ14‖ = ‖ρ23‖ = ‖ρ24‖ =

√1 + 4v2 + ε2 − (2v sinψ)

2/2 =

√1 + ε2 + (2v cosψ)

2/2. On

the other hand, the condition ε2 (2δ − 1) = ± 2ε v sinψ implies that the product v sinψ is a constant of

the motion for (3.68). We can get the corresponding value of δ from that condition, which is

δ =1

2± v sinψ

ε. (3.69)

This shows that the value of the constant is fixed by δ and ε. Moreover, since δ ∈ (0, 1), the value of

v sinψ has to lie within the interval from −ε/2 to ε/2. Thus, the equation in (3.68) governing the motion

of the centre of mass of the dumbbell in v becomes completely decoupled from the equations in ψ.

Additionally, we see from the last two equations in (3.68) that ψ is necessarily a constant, so that, by

the constancy of v sinψ, v is also constant. From this we conclude that the solution of (3.67)–(3.68)

is a steady state of (3.42). We can have either ψ = 0, π, which matches the previous case (3.63), or

ψ = π/2. Setting u = v = v ≡ 0 and ψ = π/2 in (3.68), one gets that u = v = 0. By (3.67) and (3.69),

this implies that η = δ = 1/2, hence the steady state associated to (3.68) is

u = v = 0; θ − φ = ±π/2; ψ = π/2;

u = v = ψ = 0; φ = 1; η = δ = 1/2 ,

which was encountered in (3.53) under category 2.

Category 4

The equation (3.44) restricting the motion of the dumbbell in the PCR3BP may also be satisfied by simply

imposing ψ = π/2 in (3.42). This leads to the following system of equations:

u− 2v = u− (1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)(u+ η

)+ (1− δ)δ ε cos (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))

− η

((1− δ‖ρ23‖3

‖ρ24‖3

)(u− (1− η)

)+ (1− δ)δ ε cos (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

))

v + 2u = v − (1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)v − (1− δ)δ ε sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

))

− η

((1− δ‖ρ23‖3

‖ρ24‖3

)v − (1− δ)δ ε sin (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

))

ε φ = − (1− η)((u+ η

)sin (θ − φ) + v cos (θ − φ)

)( 1

‖ρ13‖3− 1

‖ρ14‖3

)

− η((u− (1− η)

)sin (θ − φ) + v cos (θ − φ)

)( 1

‖ρ23‖3− 1

‖ρ24‖3

),

(3.70)

55

Page 80: Spin-Orbit Interaction and Chaos in Celestial Mechanics

and ψ = 0, where

‖ρ13‖2 =(u+ η

)2+ v2 + δ2 ε2 + 2 δ ε

((u+ η

)cos (θ − φ)− v sin (θ − φ)

)‖ρ14‖2 =

(u+ η

)2+ v2 + (1− δ)2 ε2 − 2 (1− δ) ε

((u+ η

)cos (θ − φ)− v sin (θ − φ)

)‖ρ23‖2 =

(u− (1− η)

)2+ v2 + δ2 ε2 + 2 δ ε

((u− (1− η)

)cos (θ − φ)− v sin (θ − φ)

)‖ρ24‖2 =

(u− (1− η)

)2+ v2 + (1− δ)2 ε2 − 2 (1− δ) ε

((u− (1− η)

)cos (θ − φ)− v sin (θ − φ)

).

(3.71)

The equations (3.70) study the planar oscillations of the dumbbell satellite in the Lagrange plane of

the primaries, in a context of spin-orbit interaction. We have already encountered some solutions with

ψ = π/2 in this section, corresponding to synchronisations 1 : 1. Those are the steady states associated

to (3.70), and, as we have seen, they are characterized by having two or more of the four distances

equal. The general solutions of (3.70) need not have any two distances equal. Moreover, all the equi-

librium solutions derived so far exist only for η and/or δ equal to 1/2. This shows that those motions are

structurally unstable (Appendix A), for any small perturbation in the values of the masses destroys the

corresponding trajectories in the phase space of the system. On the other hand, the general motions

governed by equations (3.70) do not have any restrictions on the values which η and δ can take between

0 and 1. These solutions fall under category 4. We present several numerically obtained solutions to the

system of equations (3.70) in Section 3.5.

3.4.2 Sitnikov Problem

A very interesting dynamical system is known as the Sitnikov Problem (SP). The SP is a special case

of the R3BP and it has a very simple formulation: a massless body (known as the satellite) is confined

to move along a straight line that is perpendicular to the Lagrange plane of the Keplerian orbits of two

primaries with equal masses and passes through their barycentre. Despite its simple formulation, it has

however a very rich phase space structure when the primaries move in eccentric orbits, where all kinds of

motions, namely periodic orbits, quasi-periodic orbits and chaotic motion, may be found. It is thus often

cited as a model case for the appearance of chaos. The SP is named after Russian mathematician

Kirill Sitnikov, who first showed the existence of oscillatory motions in this case of the R3BP that are

unbounded (Sitnikov [27]).

Several extensions of the SP have been studied over the years (see, for instance, Dvorak and Sui

Sun [50]). For a review of previous works on the SP, we refer to [51]. Here we will extend the problem

by assuming that the satellite has the shape of a dumbbell. The centre of mass of the dumbbell will

thus be restricted to move in a straight line perpendicular to the plane formed by the Keplerian orbits

of the primaries and passing through their common centre of mass, while the dumbbell is still allowed

to rotate in three-dimensional space. This effectively reduces the number of degrees of freedom in the

configuration space of the system of equations (3.41) by two, which becomes R× S2. A configuration of

56

Page 81: Spin-Orbit Interaction and Chaos in Celestial Mechanics

the dumbbell is now fully specified by w and the attitude angles on the sphere, φ and ψ. The equations

for the SP become

w − 2esin (θ − ξ)

1 + e cos (θ − ξ)w = −ρ4

((1− η)

((1− δ‖ρ13‖3

‖ρ14‖3

)w + (1− δ)δ ε cosψ

(1

‖ρ13‖3

− 1

‖ρ14‖3

))+ η

((1− δ‖ρ23‖3

‖ρ24‖3

)w + (1− δ)δ ε cosψ

(1

‖ρ23‖3− 1

‖ρ24‖3

)))

ε

((φ− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

φ

)sin2 ψ + 2 φ ψ sinψ cosψ

)= −ρ5 η (1− η) sinψ sin (θ − φ)

×

(1

‖ρ13‖3− 1

‖ρ23‖3+

1

‖ρ24‖3− 1

‖ρ14‖3

)

ε

(ψ − 2e

sin (θ − ξ)1 + e cos (θ − ξ)

ψ − φ2 sinψ cosψ

)= −ρ4

((1− η)

(η ρ cosψ cos (θ − φ)− w sinψ

(1

‖ρ13‖3− 1

‖ρ14‖3

)− η(

(1− η) ρ cosψ cos (θ − φ) + w sinψ)( 1

‖ρ23‖3− 1

‖ρ24‖3

)),

(3.72)

and the squared distances between the primaries and the masses of the dumbbell are given by

‖ρ13‖2 = η2ρ2 + w2 + δ2 ε2 + 2 δ ε(η ρ sinψ cos (θ − φ) + w cosψ

)‖ρ14‖2 = η2ρ2 + w2 + (1− δ)2 ε2 − 2 (1− δ) ε

(η ρ sinψ cos (θ − φ) + w cosψ

)‖ρ23‖2 = (1− η)

2ρ2 + w2 + δ2 ε2 + 2 δ ε

(− (1− η) ρ sinψ cos (θ − φ) + w cosψ

)‖ρ24‖2 = (1− η)

2ρ2 + w2 + (1− δ)2 ε2 − 2 (1− δ) ε

(− (1− η) ρ sinψ cos (θ − φ) + w cosψ

). (3.73)

Figure 3.9 depicts a theoretical configuration of the dumbbell and the primaries in the SP, in the reference

frame S.

Similarly to the case of equations (3.42) for the PCR3BP with dumbbell satellite, the system of equa-

tions (3.72), together with (3.73), is restricted by two additional conditions which follow from the vanishing

of the left hand sides of the equations in u and v in (3.41),

(1− η) η ρ

((1− δ)

(1

‖ρ13‖3− 1

‖ρ23‖3

)+ δ

(1

‖ρ14‖3− 1

‖ρ24‖3

))= − (1− δ) δ ε sinψ cos (θ − φ)

×

(1

‖ρ23‖3− 1

‖ρ24‖3

)+ (1− η)

(1

‖ρ13‖3− 1

‖ρ14‖3

)),

(1− η) ε sinψ sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

)= −η ε sinψ sin (θ − φ)

(1

‖ρ23‖3− 1

‖ρ24‖3

).

(3.74)

Equations (3.74) are necessary conditions for the motion of the centre of mass of the dumbbell to

be confined to the coordinate axis that is orthogonal to the Lagrange plane of the primaries. As in

57

Page 82: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Figure 3.9: Schematic representation of a configuration of the dumbbell and the primaries in the SP, inthe inertial frame S, at θ = (9/10)π. The parameters of the orbits of the primaries around their commoncentre of mass are as follows: λ = 3.5, ξ = 0 and e = 0.5. The length of the dumbbell is ε = 0.4. We haveused η = δ = 1/2, so the dumbbell has equal masses, and the primaries have equal masses too. Thedumbbell is aligned parallel to the Lagrange plane (ψ = π/2) and orthogonal to the line segment thatjoins the primaries (θ − φ = −π/2), while its centre of mass moves along the dashed line that coincideswith the W axis.

Section 3.4.1, we look for solutions to the SP, equations (3.72)–(3.73), consistent with these conditions,

specifically equilibrium solutions or steady states. The Hamiltonian H is not in general conserved in the

SP, because the orbits of the primaries have a nonzero eccentricity.

All the conditions that satisfy (3.74) fall into one of the following three main categories:

category 1: ‖ρ13‖ = ‖ρ24‖ and ‖ρ14‖ = ‖ρ23‖ (6= ‖ρ13‖ , ‖ρ24‖)

and η = δ =1

2;

category 2: ‖ρ13‖ = ‖ρ14‖ = ‖ρ23‖ = ‖ρ24‖ (all 4 distances equal) ;

category 3: ‖ρ13‖ = ‖ρ23‖ and ‖ρ14‖ = ‖ρ24‖ (6= ‖ρ13‖ , ‖ρ23‖) . (3.75)

Category 1

We equate ‖ρ13‖2 to ‖ρ24‖2 or ‖ρ14‖2 to ‖ρ23‖2 in (3.73), with η = δ = 1/2, to obtain

ε

(1

2ρ sinψ cos (θ − φ) + w cosψ

)= −ε

(−1

2ρ sinψ cos (θ − φ) + w cosψ

),

which has the solution w cosψ = 0. Three cases fall under this category: the first of these is

η = δ =1

2and w = 0 , (3.76)

58

Page 83: Spin-Orbit Interaction and Chaos in Celestial Mechanics

for which the equations of motion (3.72) becomeε

((φ− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

φ

)sin2 ψ + 2φ ψ sinψ cosψ

)= −1

2ρ5 sinψ sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

)

ε

(ψ − 2e

sin (θ − ξ)1 + e cos (θ − ξ)

ψ − φ2 sinψ cosψ

)= −1

2ρ5 cosψ cos (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

),

(3.77)

with ‖ρ13‖ = ‖ρ24‖ =√ρ2 + ε2 + 2 ε ρ sinψ cos (θ − φ)/2 and ‖ρ14‖ =

‖ρ23‖ =√ρ2 + ε2 − 2 ε ρ sinψ cos (θ − φ)/2. The two equations (3.77) describe the rotation of a dumb-

bell, whose centre of mass lies fixed at the barycentre of the primaries.

The second case is

η = δ =1

2and ψ = π/2 , (3.78)

and it reduces (3.72) to

w − 2e

sin (θ − ξ)1 + e cos (θ − ξ)

w = −1

2ρ4

(1

‖ρ13‖3+

1

‖ρ14‖3

)w

ε

(φ− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

φ

)= −1

2ρ5 sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

),

(3.79)

wherein ‖ρ13‖ = ‖ρ24‖ =√ρ2 + 4w2 + ε2 + 2 ε ρ cos (θ − φ)/2 and ‖ρ14‖ = ‖ρ23‖ =√

ρ2 + 4w2 + ε2 − 2 ε ρ cos (θ − φ)/2. These equations govern the motion and rotation of a dumbbell

which remains parallel to the Lagrange plane, and whose centre of mass moves along the coordinate

axis orthogonal to this plane.

Finally, it may happen that both w = 0 and ψ = π/2. This coincides with the case

η = δ =1

2and w = 0 and ψ = π/2 , (3.80)

whose equation of motion is, by (3.72),

ε

(φ− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

φ

)= −1

2ρ5 sin (θ − φ)

(1

‖ρ13‖3− 1

‖ρ14‖3

), (3.81)

with ‖ρ13‖ = ‖ρ24‖ =√ρ2 + ε2 + 2 ε ρ cos (θ − φ)/2 and ‖ρ14‖ = ‖ρ23‖ =

√ρ2 + ε2 − 2 ε ρ cos (θ − φ)/2.

Equation (3.81) characterises the planar oscillations of a dumbbell that is restricted to rotate in the La-

grange plane, and whose centre of mass lies at the barycentre of the primaries.

The equations (3.77), (3.79) and (3.81), under cases (3.76), (3.78) and (3.80), respectively, admit

the following four steady states

w = 0; θ − φ = 0, π, ±π/2; ψ = π/2;

w = ψ = 0; φ = 1; η = δ = 1/2 .(3.82)

59

Page 84: Spin-Orbit Interaction and Chaos in Celestial Mechanics

In all four steady states, the dumbbell lies in the Lagrange plane, with its centre of mass located at the

barycentre of the primaries. For θ − φ = 0, π, it rotates in inertial space always coinciding with the line

joining the primaries, while for θ − φ = ±π/2, it maintains at all times a right angle to this line. As the

fact that all the masses are at rest relative to one another in the synodic reference frame suggests, all

steady states are characterised by a 1 : 1 synchronisation between the period of rotation of the dumbbell

and the translational period of the primaries. The φ equation in (3.77), (3.79) and (3.81) transforms for

steady states (3.82) into

ε

(φ− 2e

sin (θ − ξ)1 + e cos (θ − ξ)

φ

)= 0 ,

which implies that they exist only for circular orbits of the primaries (e = 0) or at the times of pericentre

passage (θ = ξ) or apocentre passage (θ = ξ + π) of m1 and m2.6 These steady states coincide with

the ones depicted in Figures 3.6a and 3.7a for the PCR3BP.

Since all the motions described so far under this category, and particularly the steady states (3.82),

depend upon η and δ being equal to 1/2, all of them are structurally unstable, as the slightest perturbation

in the masses disrupts such solutions. The steady states (3.82) for which θ− φ = ±π/2 actually belong

to the next category, by virtue of all four distances being equal to√ρ2 + ε2/2.

Category 2

We require that all four distances, ‖ρ13‖, ‖ρ14‖, ‖ρ23‖ and ‖ρ24‖, are equal. The case

ε = 0 and η =1

2, (3.83)

for which the dumbbell reduces to a single point mass and the primaries have equal masses, clearly falls

under this category. By (3.72)–(3.73), the equation of motion for this case is

w − 2esin (θ − ξ)

1 + e cos (θ − ξ)w = −ρ4 w

(ρ2/4 + w2)3/2

. (3.84)

Noting that the left-hand side of (3.84) equals ρ4d2w

dτ2by (3.40), we immediately recognise that this is in

fact the equation for the conventional SP in dimensionless coordinates. The SP is a structurally unstable

problem, since it requires that the primaries be equally massive bodies.

There are two cases that fall under this category for ε > 0. By the same reasoning that lead to

equation (3.61) in Section 3.4.1 or simply by symmetry arguments alone, one can easily conclude that

η = 1/2 and sinψ cos (θ − φ) = 0 are necessary conditions so that all four distances are equal. Further-

more, equating ‖ρ13‖2 to ‖ρ14‖2 and ‖ρ24‖2 in (3.73), one obtains the condition

δ2ε2 + 2 δ εw cosψ = (1− δ)2 ε2 − 2 (1− δ) εw cosψ ,

6Naturally, if e 6= 0 and the steady states (3.82) exist only for θ = ξ or θ = ξ + π, then these will be static, instead of steady,states.

60

Page 85: Spin-Orbit Interaction and Chaos in Celestial Mechanics

or, equivalently,

ε2 (1− 2δ) = 2εw cosψ , (3.85)

where we have substituted the values for both η and sinψ cos (θ − φ). Equation (3.85) ascertains that the

quantity w cosψ is a constant of the motion when all four distances are equal, whose value is ε (1/2− δ).

The other cases under this category thus satisfy the following properties:

sinψ cos (θ − φ) = 0 and η =

1

2,

2εw cosψ = ε2 (1− 2δ) .

The first case is

ψ = 0, π and η =1

2and ± 2εw = ε2 (1− 2δ) , (3.86)

wherein the upper and lower signs correspond to the cases ψ = 0 and ψ = π, respectively. This case

gives an equilibrium solution. Indeed, from the first equation in (3.72) and from (3.73), one gets

w

(1− δ‖ρ13‖3

‖ρ14‖3

)= 0 ,

where ‖ρ13‖ = ‖ρ14‖ =√ρ2 + ε2/2. This has the unique solution w = 0, which, by (3.86), implies

δ = 1/2. Therefore we obtain the following steady state

w = 0; ψ = 0, π;

w = ψ = 0; η = δ = 1/2 ,(3.87)

for which the dumbbell is aligned along the coordinate axis that is orthogonal to the Lagrange plane, with

its centre of mass located at the barycentre of the primaries. Nothing can be said about synchronisation,

because ψ = 0, π are singular points of the coordinate system used. Steady state (3.87) coincides with

the one depicted in Figure 3.6b for the PCR3BP when e = 0. As this steady state only occurs for

η = δ = 1/2, we conclude that it is also structurally unstable.

The last case is

θ − φ = ±π/2 and η =1

2and 2εw cosψ = ε2 (1− 2δ) , (3.88)

whose equations of motion are, by (3.72)–(3.73),

w − 2esin (θ − ξ)

1 + e cos (θ − ξ)w = −ρ4 w

‖ρ13‖3

ε

(ψ sinψ cosψ − e sin (θ − ξ)

1 + e cos (θ − ξ)sin2 ψ

)= 0

ε

(ψ − 2e

sin (θ − ξ)1 + e cos (θ − ξ)

ψ − sinψ cosψ

)= 0 .

(3.89)

61

Page 86: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Here ‖ρ13‖ =

√ρ2 + 4w2 + ε2 − (2w cosψ)

2/2 =

√ρ2 + ε2 + (2w sinψ)

2/2. As a result of the fact that

w cosψ is a constant of the motion for (3.88)–(3.89), the w equation in (3.89) becomes decoupled from

the equations for ψ. Hence, after solving the equation for w, one may readily obtain the solution for ψ

from the additional condition w cosψ = const. The value of this constant is fixed by δ and ε. Moreover,

since δ ∈ (0, 1), the value of w cosψ has to lie within the interval from −ε/2 to ε/2. We can get the

corresponding value of δ from (3.85), which is

δ =1

2− w cosψ

ε. (3.90)

The steady states (3.82) and (3.87) found above, with θ − φ = ±π/2, are fixed points of (3.89).

Category 3

As we have seen, it is necessary that sinψ cos (θ − φ) = 0 so that ‖ρ13‖ = ‖ρ23‖ and ‖ρ14‖ = ‖ρ24‖. By

inspection of (3.73), this in turn implies that η = 1/2, or, the masses of the primaries have to be equal.

On the other hand, the equality of the distances in category 3 is not sufficient to satisfy the second

equation of (3.74). The only way to satisfy this equation and the condition sinψ cos (θ − φ) = 0 is to have

ψ = 0, π. Accordingly, the only case under category 3 is the following:

ψ = 0, π and η =1

2, (3.91)

for which the equations of motion (3.72) become

w − 2esin (θ − ξ)

1 + e cos (θ − ξ)w = −ρ4

((1− δ‖ρ13‖3

‖ρ14‖3

)w

± (1− δ) δ ε

(1

‖ρ13‖3− 1

‖ρ14‖3

)), (3.92)

with ‖ρ13‖ = ‖ρ23‖ =√ρ2/4 + w2 + δ2 ε2 ± 2 δ εw and ‖ρ14‖ = ‖ρ24‖ =√

ρ2/4 + w2 + (1− δ)2 ε2 ∓ 2 (1− δ) εw . As before, the upper and lower signs correspond to the cases

ψ = 0 and ψ = π, respectively.

Equation (3.92) governs the motion of the centre of mass of a dumbbell lined up with theW axis along

that same axis. As this motion occurs only for equal masses of the primaries, it is structurally unstable.

The steady states (3.87) are fixed points of (3.92) which occur for equal masses of the dumbbell.

We conclude from this analysis that the SP with dumbbell satellite is a structurally unstable problem,

as all the possible motions and steady states require that the mass parameter η be equal to 1/2, that is,

they only exist for equal masses of the primaries.

62

Page 87: Spin-Orbit Interaction and Chaos in Celestial Mechanics

3.5 Numerical analysis of the case under category 4 of the PCR3BP

with dumbbell satellite

We give some examples of orbits for the system of equations (3.70) together with (3.71), which studies

the planar oscillations of a rigid dumbbell satellite in the Lagrange plane of the primaries in a context of

spin-orbit interaction. These may be regarded as being only a small subset of all the possible orbits of

(3.70). We used parameters appropriate for the Pluto-Charon-Nix system, namely ε = 10−3 ([52, 42])

and η = 0.1 ([42]). The value of δ was chosen to be 0.8 throughout all the simulations. The equations

were integrated using a classic Runge-Kutta method of order 8 (Hairer et al. [53]).

In Figures 3.10–3.13, we show several special orbits of the system of equations (3.70). We were

able to numerically verify that most, if not all, initial conditions lead to either ejection of the dumbbell

from the system or collision with one of the primaries in a short amount of time (after a few revolutions

of the primaries). Specifically, if the dumbbell starts in a position outside of the unit circle centred at

the centre of mass of the primaries, that will always, according to our numerical simulations, result in

immediate ejection of the system. The orbits in Figures 3.10–3.13 are special in the sense that they

aren’t short-lived and they show more interesting behaviours of the dumbbell satellite. Nonetheless, all

of them will still lead to either a collison or ejection of the system after a certain amount of time. The fate

of the dumbbell satellite is a collision with one of the primaries in Figures 3.10, 3.11 and 3.13, and an

ejection from the system in Figure 3.12.

In each Figure, we represent the path of the centre of mass of the dumbbell satellite in the inertial

reference frame S by a solid line. The coordinates (U, V ) are dimensionless, cartesian coordinates

associated to frame S. The dashed lines indicate the orbits traced out by the primaries, which are

depicted by black circles at the time of stop of the simulation. Also shown in the Figures is the temporal

evolution of the angular difference θ−φ, which characterises the rotation of the dumbbell. In Figure 3.10,

the dumbbell starts close to one of the fixed points (3.51), with u(0) = u1(η = 0.1, ε = 10−3) ≈ 0.609,

v(0) = 0, v(0) = 0.01, φ(0) = π/2 and φ(0) = 0.99, although the mass parameter δ differs from 1/2,

as required in the fixed point, and it starts moving along a circular arc. However, soon after that the

trajectory of the centre of mass appears to become chaotic, starting to rotate around the outer of the

primaries and eventually colliding with it. This information is corroborated by the graph of the time

evolution of θ − φ, which shows that the coordinate suddenly begins to rotate around all possible values

after a start in which it remained almost constant.

In Figures 3.11 and 3.12, the dumbbell starts close to another of the fixed points (3.51), with u(0) =

u2(η = 0.1, ε = 10−3) ≈ −1.042, v(0) = 0, φ(0) = −π/2, and φ(0) = 0.9 in the case of Figure 3.11

and φ(0) = 0 in the case of Figure 3.12. In both cases it starts by tracing out a quasi-circular orbit in

configuration space, very close to the fixed point orbit, then becomes attracted to the inner, heavier,

primary, revolving around this body, and finally it ends up leaving the region in-between the primaries

(Figures 3.11a, 3.12a). In the case of Figure 3.12, the dumbbell permanently leaves this region and is

eventually ejected from the system (Figure 3.12c). In the orbit of Figure 3.11, it remains in the system,

describing large ellipses around the primaries (Figure 3.11b), until it finally returns to the starting region,

63

Page 88: Spin-Orbit Interaction and Chaos in Celestial Mechanics

-1.0 -0.5 0.5 1.0U

-1.0

-0.5

0.5

1.0

V

(a)

0.5 1.0 1.5 2.0 2.5 3.0θ

1

2

3

4

5

6

θ-ϕ(θ)

(b)

Figure 3.10: Collisional solution of (3.70) for η = 0.1, δ = 0.8, ε = 10−3 and initial conditions close toone of the fixed points (3.51): u(0) ≈ 0.609, u(0) = 0, v(0) = 0, v(0) = 0.01, φ(0) = π/2 and φ(0) = 0.99.(a) The motion of the centre of mass of the dumbbell in configuration space, represented by a solid line,appears to be chaotic. After about 11 revolutions of the primaries, depicted by black circles moving overthe dashed lines, around their barycentre, the dumbbell eventually collides with the lighter one. (b) Timeevolution of the angular difference θ − φ between θ = 0 and θ = π, with information about the rotationof the dumbbell. The coordinate suddenly begins at θ ≈ 1.8 to rotate around all possible values, after astart in which it remained almost constant.

only to collide with the lightest body (Figure 3.11c). The graphs of θ − φ in Figures 3.11d, 3.12b show

similar behaviour to the right portion of the graph on Figure 3.10b.

In Figure 3.13, the dumbbell satellite does not start near any fixed point. Its initial position in the

Lagrange plane is u(0) = −0.81, v(0) = 0.03, and the initial value of the azimuthal angle φ is φ(0) =

π/7.1. All the initial velocities are zero. The course of the centre of mass of the dumbbell (Figures 3.13a–

3.13c) is similar to the one just described on the basis of Figure 3.11. The time evolution of the angular

difference θ − φ in Figure 3.13d appears to be chaotic.

64

Page 89: Spin-Orbit Interaction and Chaos in Celestial Mechanics

-1.0 -0.5 0.5 1.0U

-1.0

-0.5

0.5

1.0

1.5

V

(a)

-60 -50 -40 -30 -20 -10U

-5

5

10

15

V

(b)

-1.0 -0.5 0.5 1.0U

-1.5

-1.0

-0.5

0.5

1.0

V

(c)

5 10 15 20 25 30θ

1

2

3

4

5

6

θ-ϕ(θ)

(d)

Figure 3.11: Collisional solution of (3.70) for η = 0.1, δ = 0.8, ε = 10−3 and initial conditions closeto another of the fixed points (3.51): u(0) ≈ −1.042, u(0) = 0, v(0) = 0, v(0) = 0, φ(0) = −π/2 andφ(0) = 0.9. (a) Initial phase of the motion of the centre of mass of the dumbbell in configuration space,between θ = 0 and θ = 65. (b) Whole trajectory of the dumbbell in configuration space, with the durationof more than 420 revolutions of the primaries (θ = 2640), corresponding in the case of the Nix-Pluto-Charon system to more than 7 years. (c) Ending phase of the motion of the centre of mass of thedumbbell in configuration space, between θ = 2612 and θ = 2640. (d) Time evolution of the angulardifference θ − φ between θ = 0 and θ = 34, with information about the rotation of the dumbbell. Noevidence of periodic motion seems to exist. The coordinate rotates around all possible values.

65

Page 90: Spin-Orbit Interaction and Chaos in Celestial Mechanics

-1.0 -0.5 0.5 1.0U

-1.0

-0.5

0.5

1.0

V

(a)

5 10 15 20 25 30θ

1

2

3

4

5

6

θ-ϕ(θ)

(b)

-1 1 2U

-10

-8

-6

-4

-2

V

(c)

Figure 3.12: Ejection solution of (3.70) for η = 0.1, δ = 0.8, ε = 10−3 and initial conditions close to one ofthe fixed points (3.51): u(0) ≈ −1.042, u(0) = 0, v(0) = 0, v(0) = 0, φ(0) = −π/2 and φ(0) = 0. (a) Initialphase of the motion of the centre of mass of the dumbbell in configuration space, between θ = 0 andθ = 45, time at which the dumbbell starts the ejection (about 7 revolutions of the primaries have passed).(b) Time evolution of the angular difference θ − φ between θ = 0 and θ = 34, with information about therotation of the dumbbell. No evidence of periodic motion seems to exist. The coordinate rotates aroundall possible values. (c) Trajectory of the dumbbell in configuration space up until θ = 60. The dumbbellhas been ejected from the system.

66

Page 91: Spin-Orbit Interaction and Chaos in Celestial Mechanics

-1.0 -0.5 0.5 1.0U

-1.0

-0.5

0.5

1.0

V

(a)

5 10 15 20 25 30U

-5

5

10

15

V

(b)

-1.5 -1.0 -0.5 0.5 1.0U

-1.0

-0.5

0.5

1.0

V

(c)

5 10 15 20 25 30θ

1

2

3

4

5

6

θ-ϕ(θ)

(d)

Figure 3.13: Collisional solution of (3.70) for η = 0.1, δ = 0.8, ε = 10−3 and initial conditions u(0) =−0.81, u(0) = 0, v(0) = 0.03, v(0) = 0, φ(0) = π/7.1 and φ(0) = 0. (a) Initial phase of the motion of thecentre of mass of the dumbbell in configuration space, between θ = 0 and θ = 45. (b) Whole trajectoryof the dumbbell in configuration space, with the duration of more than 194 revolutions of the primaries(θ = 1220), corresponding in the case of the Nix-Pluto-Charon system to more than 3 years. (c) Endingphase of the motion of the centre of mass of the dumbbell in configuration space, between θ = 1208 andθ = 1220. (d) Time evolution of the angular difference θ− φ between θ = 0 and θ = 34, showing signs ofchaotic rotation.

67

Page 92: Spin-Orbit Interaction and Chaos in Celestial Mechanics

68

Page 93: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Chapter 4

Conclusions

In this thesis we proposed to study the problem of spin-orbit interaction using an approach other than

the Averaging Theory, or the Tidal Theory, which rely on poorly determined parameters related to the

internal structure of celestial bodies. Taking that into account, we decided to consider the problem

where a satellite, modelled as a dumbbell, revolves around a point mass, interacting with it through

the gravitational force. This is what we called the Keplerian Dumbbell (KD) system. Despite not being

new, an analysis without approximations of the full dynamics of the KD was lacking. We derived the

exact equations of motion for this system, and then found and analysed its steady states or stationary

orbits. We showed that all the steady states of the KD system are unstable to small variations on the

initial conditions and that the KD is a structurally unstable problem for equal masses of the dumbbell.

For the case where the two masses of the dumbbell are equal and the motion of the centre of mass

of the dumbbell is planar, the steady states are Lyapunov unstable. For the case where the dumbbell

is aligned with the direction connecting its centre of mass to the centre of mass of the KD system and

the motion of the centre of mass of the dumbbell is planar, we have proved that for a sufficiently large

trajectory radius, these steady states are also Lyapunov unstable. Numerical analysis of the eigenvalues

of the Hessian matrix of the effective Hamiltonian associated to these steady states suggests that they

are always unstable, irrespective of the radius of the trajectory. The effective Hamiltonian is at least a

1−saddle near the steady states.

As all the steady states are Lyapunov unstable, we expect the KD to exhibit chaos. Only the steady

states 4 and 5, in which the dumbbell is aligned with the direction connecting its centre of mass to the

centre of mass of the KD, exist for different masses of the dumbbell. Interestingly, some of the unstable

steady states are the Eulerian solutions of the General Three-Body Problem. In this way, we provided a

link between the KD system and this problem.

In the limit when the length of the dumbbell goes to zero, the Kepler problem is recovered. Neverthe-

less, the stability of the system changes abruptly in this limit, since the fixed point of the Kepler problem

is stable of the centre type.

The dumbbell model was then incorporated into the framework of the Restricted Three-Body Problem

(R3BP). Our goal was to model the dynamics of the satellite Nix under the gravitational influence of Pluto

69

Page 94: Spin-Orbit Interaction and Chaos in Celestial Mechanics

and Charon. We derived the exact equations of motion of the R3BP with dumbbell satellite and focused

our attention on two special cases, one in which the dumbbell’s centre of mass is constrained to move

in the Lagrange plane — the Planar Circular Restricted Three-Body Problem (PCR3BP) with dumbbell

satellite — and the other in which it is constrained to move along an axis orthogonal to this plane and

that passes through the barycentre of the primaries — the Sitnikov Problem (SP).

Necessary conditions for these two motions to occur were obtained and, based on that, the steady

states of the system were found. Some of these steady states are the direct analogues or extensions

of the Lagrangian points in the (conventional) CR3BP ([44, 3]). While in the conventional CR3BP there

are three equilibrium points over the “x” axis, when the satellite has a dumbbell shape, there may be

1, 2, 3, 4 or 5 equilibrium points over this axis. On the other hand, there are still only two “equilateral”,

symmetric, equilibrium points in the problem with dumbbell satellite, as in the case of the conventional

CR3BP.

We have determined that these steady states only occur for special configurations of the dumbbell,

namely, when it lies on the Lagrange plane, making either a right angle, or being aligned, with the line that

connects the primaries, or when it is aligned orthogonally to this plane. Moreover, the new equilateral

equilibrium points only exist for configurations in which the dumbbell either lies on the Lagrange plane,

aligned with the line that joins the primaries, or is aligned orthogonally to the plane. And they don’t occur

for different masses of the primaries, unless the dumbbell is aligned vertically to the plane.

Finally, these equilibrium points only exist for equal masses of the dumbbell, therefore we conclude

that the R3BP with dumbbell satellite is a structurally unstable problem, much like the KD system. The

stability of these fixed points will be studied in a future paper.

The special case of the PCR3BP with dumbbell satellite in which the spin axis of the dumbbell is or-

thogonal to the Lagrange plane was also studied numerically, with values of the parameters appropriate

for the Pluto-Charon-Nix system. We verified that most, if not all, initial conditions will lead to either a

collision with a primary or ejection of the dumbbell from the system in finite time. This could mean that

Nix can be just passing by our solar system right now. For instance, we obtained a solution in which Nix

would eventually collide with Charon in a little over 7 years. Some trajectories show definitely signs of

chaoticity or, at least, of non periodicity.

Lastly, we point out that, unlike the conventional CR3BP, the CR3BP with dumbbell satellite doesn’t

in general possess any invariant. We derived sufficient conditions for the conservation of an effective

Hamiltonian in the CR3BP with dumbbell satellite. For instance, whenever the motion of the dumbbell

is such that it is either always perpendicularly oriented to the Lagrange plane or confined to move and

rotate in the mediating plane of the line that connects the primaries, there will be conservation of an

effective Hamiltonian.

In the future we plan to extend the analysis of the dumbbell rigid body into a body constituted by

multiple dumbbells, each aligned along one of the coordinate axes. We also wish to study the inter-

action between two dumbbell rigid bodies, in order to investigate other types of synchronisations, and,

ultimately, to replace the dumbbell by an axisymmetric body.

70

Page 95: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Bibliography

[1] H. A. Weaver, S. A. Stern, M. J. Mutchler, A. J. Steffl, M. W. Buie, W. J. Merline, J. R. Spencer, E. F.

Young, and L. A. Young. Discovery of two new satellites of pluto. Nature, 439:943–945, 2006.

[2] M. R. Showalter and D. P. Hamilton. Resonant interactions and chaotic rotation of pluto’s small

moons. Nature, 522:45–49, June 2015. doi: 10.1038/nature14469.

[3] V. G. Szebehely. Theory of Orbits, The Restricted Problem of Three Bodies. Academic Press,

1967.

[4] C. D. Murray and S. F. Dermott. Solar System Dynamics. Cambridge University Press, 1999. ISBN:

978-0521575973.

[5] R. Fitzpatrick. An Introduction to Celestial Mechanics. Cambridge University Press, 2012. ISBN:

978-1107023819.

[6] J. Wisdom, S. J. Peale, and F. Mignard. The chaotic rotation of hyperion. Icarus, 58:137–152, May

1984. doi: 10.1016/0019-1035(84)90032-0.

[7] J. J. Klavetter. Rotation of hyperion. I. observations. Astronomical Journal, 97:570–579, Feb. 1989.

doi: 10.1086/115006.

[8] T. Grav et al. The short rotation period of nereid. The Astrophysical Journal Letters, 591:L71, June

2003. doi: 10.1086/377067.

[9] B. E. Schaefer et al. Nereid: Light curve for 1999–2006 and a scenario for its variations. Icarus,

196:225–240, July 2008. doi: 10.1016/j.icarus.2008.02.025.

[10] G. H. Pettengill and R. B. Dyce. A radar determination of the rotation of the planet mercury. Nature,

206:1240, 1965. doi: 10.1038/2061240a0.

[11] G. J. F. MacDonald. Tidal friction. Rev. Geophys., 2:467–541, 1964. doi: 10.1029/

RG002i003p00467.

[12] S. J. Peale and T. Gold. Rotation of the planet mercury. Nature, 206:1240–1241, 1965. doi:

10.1038/2061240b0.

[13] P. Goldreich and S. J. Peale. Spin-orbit coupling in the solar system. Astronomical Journal, 71:

425–438, 1966. doi: 10.1086/109947.

71

Page 96: Spin-Orbit Interaction and Chaos in Celestial Mechanics

[14] P. Goldreich. Final spin states of planets and satellites. Astronomical Journal, 71:1–7, Feb. 1966.

doi: 10.1086/109844.

[15] A. E. H. Love. Some Problems of Geodynamics. Cambridge University Press, 1911. ISBN: 978-

1332316601.

[16] J. Wisdom. Urey prize lecture: Chaotic dynamics in the solar system. Icarus, 72:241–275, Nov.

1987. doi: 10.1016/0019-1035(87)90175-8.

[17] C. F. Yoder. Astrometric and geodetic properties of earth and the solar system. In T. J. Ahrens,

editor, Global Earth Physics: A Handbook of Physical Constants. AGU Reference Shelf, volume 1,

pages 1–31. American Geophysical Union, Feb. 1995.

[18] A. Celletti, G. Della Penna, and C. Froeschle. Analytical approximation of the solution of the dis-

sipative standard map. International Journal of Bifurcation and Chaos, 8:2471–2479, Dec. 1998.

doi: 10.1142/S0218127498001984.

[19] P. Goldreich and S. J. Peale. Dynamics of planetary rotations. Annu. Rev. Astron. Astrophys., 6:

287–320, Sept. 1968. doi: 10.1146/annurev.aa.06.090168.001443.

[20] W. M. Kaula. Tidal dissipation by solid friction and the resulting orbital evolution. Rev. Geophys., 2:

661–685, Nov. 1964. doi: 10.1029/RG002i004p00661.

[21] M. Efroimsky and V. V. Makarov. Tidal friction and tidal lagging. applicability limitations of a popular

formula for the tidal torque. The Astrophysical Journal, 764(1):26 (10pp), Jan. 2013. doi: 10.1088/

0004-637X/764/1/26.

[22] A. C. M. Correia and J. Laskar. Mercury’s capture into the 3/2 spin-orbit resonance as a result of

its chaotic dynamics. Nature, 429:848–850, June 2004. doi: 10.1038/nature02609.

[23] H. Poincare. Les Methodes nouvelles de la Mecanique Celeste, volume Tome I. Gauthier-Villars,

Paris, 1892.

[24] M. Henon and C. Heiles. The applicability of the third integral of motion: Some numerical experi-

ments. Astronomical Journal, 69:73–79, Feb. 1964. doi: 10.1086/109234.

[25] B. V. Chirikov. A universal instability of many-dimensional oscillators systems. Phys. Rep., 52:

263–379, May 1979. doi: 10.1016/0370-1573(79)90023-1.

[26] J. Wisdom. Chaotic behaviour in the solar system. Proc. R. Soc. London A, 413:109–129, Sept.

1987. doi: 10.1098/rspa.1987.0104.

[27] K. Sitnikov. The existence of oscillatory motions in the three-body problem. Doklady Akademii Nauk

SSSR, 133:303–306, 1960.

[28] G. Jacobi. Sur le mouvement d’un point et sur un cas particulier du probleme des trois corps.

Comptes Rendus, 3:59–61, 1836.

72

Page 97: Spin-Orbit Interaction and Chaos in Celestial Mechanics

[29] D. Scheeres. Orbits close to asteroid 4769 castalia. Icarus, 121(1):67–87, 1996.

[30] D. J. Scheeres, B. G. Williams, and J. K. Miller. Evaluation of the dynamic environment of an

asteroid: Applications to 433 eros. J. Guid. Control. Dyn., 23(3):466–475, 2000.

[31] V. V. Beletsky and E. M. Levin. Dynamics of space tether systems, volume 83 of Adv. Astronaut.

Sci. AAS Publication, 1993.

[32] A. J. Maciejewski, M. Przybylska, L. Simpson, and W. Szuminski. Non-integrability of the dumb-

bell and point mass problem. Celest. Mech. Dyn. Astr., 117(3):315–330, 2013. doi: 10.1007/

s10569-013-9514-7.

[33] A. A. Burov. Non-integrability of planar oscillation equation for satellite in elliptical orbit. Vestnik

Moskov. Univ. Ser. Mat. Mekh., 1:71–73, 1984. in Russian.

[34] A. Celletti and V. Sidorenko. Some properties of the dumbbell satellite attitude dynamics. Celest.

Mech. Dyn. Astr., 101:105–126, May 2008. doi: 10.1007/s10569-008-9122-0.

[35] A. Elipe, M. Palacios, and H. Pre¸tka-Ziomek. Equilibria of the three-body problem with rigid dumb-

bell satellite. Chaos, Solitons & Fractals, 35:830–842, Mar. 2008. doi: 10.1016/j.chaos.2006.05.

071.

[36] H. Goldstein. Classical Mechanics. Addison-Wesley, 2nd edition, 1980. ISBN: 978-0201029697.

[37] L. A. Pars. A Treatise on Analytical Dynamics. Heinemann, London, 1965.

[38] Z. E. Musielak and B. Quarles. The three-body problem. Reports on Progress in Physics, 77:

065901 (30pp), June 2014. doi: 10.1088/0034-4885/77/6/065901.

[39] M. W. Hirsch and S. Smale. Differential equations, Dynamical systems, and Linear Algebra. Aca-

demic Press, San Diego, 1974. ISBN: 978-0123495501.

[40] V. I. Arnold. Mathematical Methods of Classical Mechanics, volume 60 of Graduate Texts in Math-

ematics. Springer-Verlag New York, 2nd edition, 1989.

[41] T. Poston and I. Stewart. Catastrophe Theory and its Applications. Pitman London, 1978.

[42] S. A. Stern, F. Bagenal, K. Ennico, G. R. Gladstone, et al. The pluto system: Initial results from its

exploration by new horizons. Science, 350:aad1815–aad1815, Oct. 2015. doi: 10.1126/science.

aad1815.

[43] J. A. Vera. On the periodic solutions of a rigid dumbbell satellite placed at L4 of the restricted

three body problem. International Journal of Non-Linear Mechanics, 51:152–156, May 2013. doi:

10.1016/j.ijnonlinmec.2013.01.013.

[44] L. Lagrange. Essai sur le probleme des trois corps. Euvres, 6:229–331, 1772.

[45] M. Brozovic, M. R. Showalter, R. A. Jacobson, and M. W. Buie. The orbits and masses of satellites

of pluto. Icarus, 246:317–329, Jan. 2015. doi: 10.1016/j.icarus.2014.03.015.

73

Page 98: Spin-Orbit Interaction and Chaos in Celestial Mechanics

[46] L. Euler. De moto rectilineo trium corporum se mutuo attrahentium. Novo Comm. Acad. Sci. Imp.

Petrop., 11:144–151, 1767.

[47] H. Poincare. Les Methodes nouvelles de la Mecanique Celeste, volume Tome III. Gauthier-Villars,

Paris, 1899.

[48] W. S. Koon, M. W. Lo, J. E. Marsden, and S. D. Ross. Dynamical systems, the three-body problem

and space mission design. Interdisciplinary Applied Mathematics. Springer, 2011. ISBN: 978-

0387495156.

[49] H. Povenmire. Dumbbell shaped objects in the solar system. Meteoritics and Planetary Science

Supplement, 37:A119, July 2002.

[50] R. Dvorak and Y. Sui Sun. The phase space structure of the extended sitnikov-problem. Celest.

Mech. Dyn. Astr., 67:87–106, Jan. 1997. doi: 10.1023/a:1008265310911.

[51] R. Dvorak and C. Lhotka. Sitnikov problem. Scholarpedia, 9(12):11096, 2014. revision #142999.

[52] M. W. Buie, W. M. Grundy, E. F. Young, L. A. Young, and S. A. Stern. Orbits and photometry of

pluto’s satellites: Charon, s/2005 p1, and s/2005 p2. Astronomical Journal, 132(1):290, 2006.

[53] E. Hairer, C. Lubich, and G. Wanner. Geometric Numerical Integration, Structure-Preserving Algo-

rithms for Ordinary Differential Equations. Springer-Verlag, 2002. ISBN: 978-3662050200.

[54] R. Dilao. Uma introducao informal a teoria dos sistemas dinamicos e do caos. url: http://sd.ist.utl.pt,

2016.

[55] G. Birkhoff and G. Rota. Ordinary Differential Equations. Wiley, 3rd edition, 1978.

[56] P. Hartman. Ordinary Differential Equations. Classics in applied mathematics 38. Society for In-

dustrial and Applied Mathematics, 2nd edition, 1982.

[57] J. E. Mansilla. Stability of hamiltonian systems with three degrees of freedom and the three-body

problem. Celest. Mech. Dyn. Astr., 94:249–269, 2006.

[58] C. D. Moore. Braids in classical gravity. Physical Review Letters, 70:3675–3679, 1993.

[59] A. Chenciner and R. Montgomery. A remarkable periodic solution of the three-body problem in the

case of equal masses. Annals of Mathematics, 152:881–901, 2000.

[60] R. Montgomery. A new solution to the three-body problem. Notices of the American Mathematical

Society, 48:471–481, 2001.

[61] M. Suvakov and V. Dmitrasinovic. Three classes of newtonian three-body planar periodic orbits.

Physical Review Letters, 110(11):114301 (4pp.), 2013.

74

Page 99: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Appendix A

Fixed Points and Steady States of

Differential Equations

We follow the work of Dilao [54] in this appendix.

A.1 Differential equations as dynamical systems

Let S be a dynamical system characterised by a finite number of state variables, x =(x1, x2, . . . , xn

).

The set of all possible values of the state variables(x1, x2, . . . , xn

)is called the phase space of the

system S .

Many systems in Nature may be modelled by differential equations, like, for instance, the demo-

graphic evolution of a certain country, or the phase transition of liquid helium from the normal to the

superfluid phase. Let f(x, t) =(f1(x, t), . . . , fn(x, t)

): Rn × R → Rn be a continuous function defined

on an open set U of Rn × R, and consider the system of ordinary differential equations

dx

dt= f(x, t), (A.1)

describing system S . Here t ∈ R, and Rn is the phase space of S and of (A.1). An application

φ(t; t0,x0) : I ⊂ R→ Rn is said to be a solution of (A.1) if it satisfies

dt= f(φ(t; t0,x0), t),

x0 being the value of the solution at the time instant t0. If, additionally, the application φ is such that

φ(t; t0,x0) = φ(t+ T ; t0,x0),

for all t ∈ I, where T > 0 is a positive constant, then it is called a periodic solution or periodic orbit of

(A.1). The image of the solution φ(t; t0,x0) in the phase space of (A.1) is called a phase curve of the

differential equation.

75

Page 100: Spin-Orbit Interaction and Chaos in Celestial Mechanics

The system of differential equations (A.1) defines a vector field or a phase flow in the phase space,

whose components along the directions x1, . . . , xn are the functions f1(x, t), . . . , fn(x, t), respectively.

The image of this vector field at a point x∗ ∈ Rn, that is, the vector f(x∗, t) =(f1(x∗, t), . . . , fn(x∗, t)

),

is tangent to the phase curve associated to φ(t; t0,x0) at x∗. Conversely, a vector field always deter-

mines, or is always associated to, a differential equation. This property makes it possible to qualitatively

construct the solutions of (A.1) in the phase space, at least in dimensions 1, 2 and 3. The function f

may not depend on time. In that case, the system (A.1) is called autonomous and is invariant for time

translations. We now define the concept of fixed point of a differential equation:

Definition 1. A point x∗ ∈ Rn is said to be a fixed point for the phase flow of the differential equation

(A.1) if f(x∗, t) = 0.

If x∗ is a fixed point of (A.1), thendx

dt= 0 at that point, and thus φ(t; t0,x

∗) = x∗ for all t ∈ R. This

means that there is no phase flow at x∗. The fixed points are thus particular solutions of a differential

equation. The fixed point solutions are also known by the names equilibrium points or steady points.

Nonetheless, not all equilibrium solutions or steady states are fixed points, even though all fixed points

may be classified as equilibrium solutions or steady states. For instance, the periodic orbits of (A.1)

correspond to steady state configurations of the dynamical system S . Due to the coordinates adopted

throughout Chapters 2 and 3, the fixed points we determine there are in fact periodic orbits, and thus

steady states, of the KD system and of the R3BP with dumbbell satellite.

The existence of fixed points and periodic orbits of a differential equation constrains the topology of

the phase curves of that equation in phase space. Through the knowledge of the vector field associated

to the equation and its fixed points, one can determine, at least qualitatively, all topologies of the phase

curves. That is why it is so crucial to obtain the fixed points of a differential equation. Furthermore, fixed

points are, in general, the only easily obtainable solutions of a nonlinear differential equation.

A.2 Stability of the fixed points

Let us analyse the behaviour of the solutions of a differential equation in the neighbourhood of the fixed

points of the associated vector field.

Definition 2. (Lyapunov stability). A fixed point x∗ of a differential equation is said to be Lyapunov stable

if:

1. There exists a neighbourhood U (x∗) of x∗ such that φ(t; t0,x0), with x0 ∈ U (x∗), is defined for

all t ≥ 0.

2. For all sufficiently small neighbourhoods V (x∗) ⊂ U (x∗), there exists a neighbourhood V1(x∗) ⊂

V (x∗) such that, for all t ≥ 0, every solution φ(t; t0,x0) with origin in V1(x∗) is in V (x∗) (Fig-

ure A.1a).

Additionally, if limt→∞ φ(t; t0,x0) = x∗ and x0 ∈ V1(x∗), then x∗ is asymptotically stable (Figure A.1b).

If a fixed point is not Lyapunov stable, then it is (Lyapunov) unstable.

76

Page 101: Spin-Orbit Interaction and Chaos in Celestial Mechanics

U (x∗)

x∗

V (x∗)

V1(x∗)

x0

(a)

U (x∗)

x∗

V (x∗)

(b)

Figure A.1: (a) Lyapunov stable fixed point. (b) Asymptotically stable fixed point.

In the definition of Lyapunov stability, the condition limt→∞ φ(t; t0,x0) = x∗ is not a sufficient condi-

tion for the asymptotic stability of the fixed point x∗. In fact, the solution φ(t; t0,x0) may tend to the fixed

point x∗ at infinity without being bounded from above for finite values of t, Birkhoff and Rota [55].

The analysis of the stability of a fixed point may be a difficult problem for nonlinear equations. In

some cases though, we may apply the stability criterium formulated in the following theorem:

Theorem 3. (Lyapunov). Let x∗ ∈ Rn be an isolated fixed point of the differential equationdx

dt= f(x)

and V : V (x∗) → R be a differentiable function in V (x∗) − {x∗}. Moreover, suppose that V (x∗) = 0

and that V (x) > 0 for x 6= x∗. Then, if V is such thatdV

dt≤ 0 in V (x∗) − {x∗}, the fixed point x∗ is

Lyapunov stable. If V is such thatdV

dt< 0 in V (x∗) − {x∗}, the fixed point x∗ is asymptotically stable.

The function V is called the Lyapunov function.

We note that the Lyapunov theorem is valid only for autonomous differential equations. The Lyapunov

function, if it exists, has a simple geometrical meaning: if x∗ is a Lyapunov stable fixed point, the

Lyapunov function has a local minimum at x∗, and the vector field, restricted to the level sets of V ,

V (x) = z, points inward to these level sets, that is to say, it makes either an obtuse or a right angle with

the direction defined by the gradient of V at each point (Figure A.2).

In addition to the Lyapunov stability criterium, one may also study the stability of a fixed point of a

differential equation by linearising the associated vector field around the fixed point. If this approach is

taken, a matrix equation is obtained, and, due to the dependence of the solutions of a linear system of

differential equations on the eigenvalues of this matrix, the stability of the fixed point of the linear system

may be inferred from these eigenvalues. Specifically, for the two-dimensional case, the fixed point of the

linear system is Lyapunov stable if the real part of both eigenvalues is negative, or, this being equal to

zero, the imaginary parts are both nonvanishing. If the real part of one of the eigenvalues is positive, the

fixed point of the linear system is Lyapunov unstable.

Definition 3. Fixed points of differential equations with pure imaginary eigenvalues are said to be elliptic

fixed points or centres. Fixed points with eigenvalues outside of the imaginary axis of the complex plane

77

Page 102: Spin-Orbit Interaction and Chaos in Celestial Mechanics

(a)

V −1(c)

(ii)

(i)

(iii)

(b)

Figure A.2: (a) Graph of a Lyapunov function V : R2 → R, depicted in the neighbourhood of a stablefixed point at (x, y) = (0, 0). Also illustrated is a plane that intersects the Lyapunov function at constantz, V (x, y) = c. (b) Level set of a Lyapunov function, projected onto the (x, y) plane. This projection is the

graph of the function V −1(c). Three phase vectors are sketched over the level set V −1(c): (i)dV

dt> 0;

(ii)dV

dt= 0; (iii)

dV

dt< 0. The sign of the derivative is determined by the angle between the vector field

and the direction of the gradient of V , which is always normal to the level set.

are called hyperbolic fixed points.

The above analysis is valid only in a neighbourhood of the fixed point. Moreover, if the nonlinear

equation has multiple fixed points, the linear analysis of the stability has to be done around each fixed

point. The following theorem states the conditions under which the stability of the fixed points of the

linear and of the nonlinear systems are the same:

Theorem 4. (Hartman-Grobman [56]). If the matrix of the linear approximation of a nonlinear system of

differential equations around a fixed point x∗, or Jacobian matrix, has no pure imaginary eigenvalues or

eigenvalues equal to zero, then there is an homeomorphism h, defined in an open neighbourhood U of

x∗, which takes orbits of the nonlinear system into orbits of the linear system. The homeomorphism h

preserves the sense of the orbits and may be chosen such that it also preserves the parameterisation

of time1.

This means that in the neighbourhood of the hyperbolic fixed points of a nonlinear system, the phase

curves of the nonlinear system and of the local, linear systems are topologically equivalent. For instance,

Figure A.3 shows that the orbits in phase space of a nonlinear system of differential equations in R2,

around an hyperbolic fixed point of the saddle type (one positive and one negative eigenvalue), and the

orbits of the associated linear system are homeomorphic to one another. The Hartman-Grobman theo-

rem justifies the qualitative construction of the orbits of the nonlinear system in phase space, discussed

in Section A.1.

We end this appendix with the definition of structural stability:

1An homeomorphism preserves the sense of the orbits if its Jacobian matrix has positive determinant.

78

Page 103: Spin-Orbit Interaction and Chaos in Celestial Mechanics

U

h

Figure A.3: Orbits of a nonlinear dynamical system in phase space (left), and corresponding approximatelinear system (right), in a neighbourhood U of an hyperbolic fixed point. The Hartman-Grobman theoremguarantees the existence of the homeomorphism h.

Definition 4. (Structural stability). A dynamical system is said to be structurally stable if any infinitesi-

mally small perturbation in the equations that define the phase flow doesn’t change the topology of the

phase curves.

Other concepts of stability include, for instance, Lagrange or Poisson stability. The motion of a point

x is Poisson stable if it returns infinitely many times to positions arbitrarily close to x. Poincare [47]

established that there are an infinite number of such motions in the R3BP. The trajectory of a point x

is Lagrange stable if it is contained in a totally bounded set2. Interestingly, both of these concepts were

actually introduced by Poincare in [47]. We shall not delve further into them.

2A totally bounded set is a set that can be covered by finitely many open balls of radius ε, for every ε > 0.

79

Page 104: Spin-Orbit Interaction and Chaos in Celestial Mechanics

80

Page 105: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Appendix B

Particular Solutions of the General

Three-Body Problem

In the Three-Body Problem, three bodies of arbitrary masses move in three-dimensional space under

their mutual gravitational interactions, according to Newton’s law of universal gravitation. When there

are no restrictions on the masses, this system is referred to as the General Three-Body Problem. In this

appendix we make a brief review of some of the special, periodic solutions to this problem.

B.1 Eulerian solutions

These special solutions were discovered by Euler [46] in 1767, who considered an initial straight line

configuration for three arbitrary masses. Euler proved that this initial straight line configuration is main-

tained for future times if the ratio of the distances between the masses has a certain value that depends

on them, and if suitable initial conditions are chosen. Furthermore, he showed that this ratio is also

conserved during the motion and that the line that joins the three masses rotates about their centre

of mass. This in turn results in all the masses travelling along ellipses (Figure B.1a). On the other

hand, for the special case where all the masses are equal, two of the masses rotate around the third

one, in identical ellipses or in a circle, always remaining in phase opposition relative to the central mass

(Figures B.1b–B.1c).

The Eulerian solutions are unstable to small deviations from the initial conditions. As the masses

may be ordered in three different ways along the line (if they are different), there are three solutions, one

for each ordering of the masses, corresponding to the case depicted in Figure B.1a.

B.2 Lagrangian solutions

In 1772, Lagrange [44] found a second class of periodic orbits of the General Three-Body Problem, now

known as Lagrangian solutions. In these solutions the three masses are initially positioned at the vertices

of an equilateral triangle. For suitable initial conditions, the masses travel along ellipses around their

81

Page 106: Spin-Orbit Interaction and Chaos in Celestial Mechanics

(a) (b)

(c)

Figure B.1: Eulerian solutions of the General Three-Body Problem: the three masses move in ellipsesaround their common centre of mass, maintaining at all times a collinear configuration represented bythe dashed line. In (a), all the masses are different. In (b) and (c), all masses are equal, and both theouter masses are at the same distance from the central mass. In (c), the outer masses move along thesame circular orbit. This configuration is obtained for suitable initial velocities of these masses.

centre of mass in configuration space, while retaining this special configuration. The triangle changes its

size and orientation periodically, but it remains equilateral for as long as the masses move. In Figure B.2

we depict two Lagrangian solutions for the case of equal masses.

(a) (b)

Figure B.2: Lagrangian solutions for the case of equal masses: the three masses are initially located atthe vertices of an equilateral triangle, which is represented by a dashed line, and this configuration ismaintained for future times. In (a), the masses move along identical ellipses. In (b), the masses travelalong the same circular orbit. This configuration is obtained for suitable initial velocities of the threemasses.

82

Page 107: Spin-Orbit Interaction and Chaos in Celestial Mechanics

Unlike the Eulerian solutions, Lagrangian solutions have regions of stability under certain conditions

[57]. By virtue of the limitation on the mutual positions of the masses in the Eulerian and Lagrangian

solutions, these are called particular solutions.

B.3 Other periodic solutions

More recently, new periodic solutions have been discovered, such as the famous figure-eight solution for

equal masses, in which the three masses trace the same figure-eight orbit, the intersection point being

the centre of mass. This solution was numerically discovered in 1993 by Moore [58] and independently

rediscovered in 2000 by Chenciner and Montgomery [59], who also rigorously proved its existence in

2001, [60].

Suvakov and Dmitrasinovic [61] have also found 13 new periodic, exotic solutions in the planar case

of the General Three-Body Problem.

83

Page 108: Spin-Orbit Interaction and Chaos in Celestial Mechanics

84