Project by KC Law 20190915

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

A LARGE-SCALE EVALUATION OF

WISDOM’S SCALING LAW

Law, Ka Chung

A thesis submitted in partial fulfilment of the requirements of

Liverpool John Moores University

for the degree of

Master of Science in Astrophysics.

September 2019
Abstract

About four decades ago Wisdom (1980) suggested the 2/7 scaling law to separate the

chaotic from regular behaviour of a small body in circular restricted three-body problem.

Until recent decade, literature identifies differentiated critical values of the law where the

max-min ratio is up to 1.5 to 2. The problem is revisited here. A classic setup is solved

numerically to examine the equilibria and long-term evolutions of the orbits. Wide range of

different parameters is worked out with orbital eccentricity allowed to deviate from zero.

These constitute a large-scale of 249 cases to evaluate this 2/7 scaling law. We find that it

is still generally doing a good job in separating the orbital behaviour of these cases, but the

critical value does span a band instead of exhibiting an accurate single point, and the band

is roughly consistent with that found in the literature. In comparing with the recently

suggested “1/5 new scaling law” by Mustill and Wyatt (2012) where eccentricity is

incorporated, the “old Wisdom” still performs better for the all the cases worked out in this

work. Lastly, some recent studies about the equilibrium Lagrange points in the Solar

system are reviewed including the newly discovered Earth moons.

ii
In memory of my father, Hung Fei Law

(1936 – 15th September 2018)

iii
Declaration

The work presented in this thesis was carried out at the Astrophysics Research Institute,

Liverpool John Moores University. Unless otherwise stated, it is the original work of the

author.

While registered as a candidate for the degree of Master of Science in Astrophysics, for

which submission is now made, the author has not been registered as a candidate for any

other award. This thesis has not been submitted in whole, or in part, for any other degree.

Law, Ka Chung

Astrophysics Research Institute

Liverpool John Moore’s University

IC2, Liverpool Science Park

146 Brownlow Hill

Liverpool

L3 5RF

UK

15th September 2019

iv
Acknowledgements

Over the past four months, LJMU, my supervisor Professor Shiho Kobayashi and I were

analogous to a system of primary star, secondary star and test particle respectively. At the

beginning I relied so much on my supervisor that our “separation” was so small and my

work was “chaotic”. By now, my reliance is lesser and our “separation” is hopefully larger

than the “critical value” so that this thesis will become “regular”. Sincerely thanks, Shiho.

v
Table of Contents

1. Introduction 1
2. CR3BP Setup 4
2.1 Notation Definitions 4
2.2 Primary and Secondary Stars 5
2.3 Test Particle Motion 6
2.4 Synodic System 7
2.5 Initial Conditions 9
3. Simulation 11
3.1 Numerical Integration 11
3.2 Coding 12
3.3 Parameters 13
4. Static Results 15
4.1 Lagrange Points 15
4.2 Equilibrium Potential 16
4.3 Horseshoe and Tadpole Orbits 18
4.4 Solar System Examples 20
5. Dynamic Results 23
5.1 Orbital Plots 23
5.2 Chaotic vs Regular Cases 25
6. Elliptic Case 30
6.1 Notation and Setup 30
6.2 Other Conditions 32
6.3 Simulation 34
6.4 Results 36
6.5 Discussion 39
References 41
Appendix A. C++ Code for Poincaré Surface-of-Section 43
Appendix B. GNUPLOT Code for 𝑼∗ (𝒙 ̃)
̃, 𝒚 44
Appendix C. C++ Code for Horseshoe Orbit 44

vi
Section 1. Introduction

Since the Keplerian and Newtonian laws of motions developed in the 17th century,
planetary orbits have once been regarded as predictably simple. Simple as it seems, the
slight deviation of Mercury orbit from its presumed elliptic path as termed perturbation is
yet a predictably unpredictable phenomenon. As the proverb “A secret between two is
God’s secret, a secret between three is everybody’s” goes, the relationship between three
persons is much more complicated than two. The same is also true for the celestial bodies.

Two-body problem is well known to have analytical solution: the orbit is always a
conic section, but nonlinear three-body problem does not have closed-form solution where
the complex behaviour is chaotic. As the mathematician Robert Luke Devaney defines:

A dynamical system is chaotic if it has the following three properties:

1. It has sensitive dependence on initial conditions;


2. It is topological transitive (points from one open set will finally hit the other);
3. Its periodic orbits form a dense set.

The first condition is also commonly known as “butterfly effect”, ever since the paper
titled “Predictability: Does the Flap of a Butterfly’s Wings in Brazil set off a Tornado in
Texas?” published by a mathematician and meteorologist Edward Norton Lorenz in 1972.
Such behaviour is also known as “deterministic chaos” because for any given set of initial
condition the system is scientifically determined, but as the initial condition is varied
slightly the outcome can differ qualitatively. As long as the initial condition is not known
for sure or as calculation error grows, the outcome is indeterminate in advance, i.e., a chaos.

Return to our celestial mechanics context. In any two-body system the variation of
initial conditions such as positions and velocities will not alter the orbits as a family of
conic section. But in three or more bodies different initial conditions can lead to
fundamentally different orbits, namely, chaotic versus regular. Some go with resonances.

In this project, a typical three-dimensional circular restricted three-body problem


(CR3BP) is solved. It comprises a primary star, a secondary star and a test particle of small
mass that does not affect the other orbits. This represents a star-planet-satellite (or any
Trojan body) system. It also suffices to demonstrate the complex behaviour by considering
a coplanar system so that the problem further boils down to a two-dimensional one. This is
the simplest case of a three-body problem as the two heavier bodies constitute a two-body
problem which is solved typically and any chaotic behaviour is confined to the test particle.
1
Since a closed-from solution is unavailable, it must be done numerically. It is not
hard to imagine that positions, velocities and accelerations are the key variables, and the
main interest is how they evolve over time. After suitable mapping, the second-order non-
linear differential equations are mapped as vector form of first-order differential equations.
This form needs to be integrated numerically, but its first-order nature is easier to handle.
A common standard solution to this is done via the fourth-order Runge-Kutta method.

After the orbits are obtained, there can be two aspects of analysis. The first is static
and the second is dynamic. On the static front equilibria will be solved. There are the
famous Lagrange points where the test particle will appear static in the system. Specifically,
there are two such points in stable force equilibria. Interesting nature around these points
as well as the nearby motions will be examined, and some recent studies around these
equilibria in the Solar system will be discussed. As will be seen, the CR3BP framework is
albeit simple, but is very useful in explaining satellites and trojans around planets, and is
especially useful in finding new moons. Very recently two Earth moons are discovered.

On the dynamic front, orbits under different initial conditions will be compared. It
is intuitive that if the test particle is too far from the primary and secondary stars, it will
simply exhibit a near circular orbit. And if it is too close to either one, say the secondary
star, then it will also exhibit a near circular orbit by simply rotating around the closer star.
However, if the test particle is rotating around both stars forming a three-body system, then
as Wisdom (1980) shows that the stability of test particle depends on its separation from
secondary star: orbit would be unstable when the separation is smaller than a critical value.
This is the famous “2/7 scaling law” where the separation threshold is a power law in the
mass ratio of secondary-to-primary star with exponent 2/7. We will test this over ranges of
different parameters, which are widest to our knowledge among the literature. Specifically,
the mass ratio of secondary-to-primary star as well as separation between secondary star
and the small object to be studied are the two dimensions of parameters to be worked on.
The derived thresholds for these cases will be compared with the theoretical scaling law.

Another main aspect of this work is to redo the exercise with slightly eccentric
orbits for the small body. This is a natural extension as no orbits are perfectly circular.
How such change will affect the applicability of the 2/7 scaling law is not well understood.
Despite Wisdom (1980) has briefly touched upon this variation, as we will see, simulation
with a large set of cases do show some patterns as eccentricity varies. In addition, the
interaction between the variations of mass ratio, eccentricity and separation will be studied
altogether, and results will be compared with those implied from a “1/5 new scaling law”.
2
To check whether the behaviour is chaotic or regular, there is simple tool called
Poincaré surface-of-section which is simply a two-dimensional plot yet embeds almost all
information about the system. As we shall see, the split between chaotic and regular is not
a single value but a band. In quite some of the middle cases there are mixed behaviour of
both. These exhibit resonance behaviour which is a special type of regular pattern. The
forms of resonance will be discussed and are compared with the taxonomy in the literature.

Slight variation of these critical values from the literature will also be mentioned.
This will be the key evaluation conducted in this study since the upper bound of the
threshold identified in the literature is almost 1.5 to 2 times of the lower bound identified.
By running simulation with totally 249 cases, we confirm that the 2/7 scaling law is still by
far the best description of the critical threshold, but it does span a range of values which
are by and large consistent with those identified from different studies across the literature.

As echoed with the beginning discussion, the chaotic behaviour has both theoretical
and empirical significance in the field of Astrophysics. On the theoretical side, it confines
the predictability of macro models to a very extent. Although it is always a science to
predict what happens deterministically when inputs are given, the enormous possibility of
inputs render such prediction invalid albeit with high accuracy. On the empirical side, this
explains perturbation of planets such as that of Mercury observed in many centuries ago.
As initial conditions play a key role, what happens to Mercury cannot be projected in other
planets; similar chaotic nature exhibits at the Solar system level in face of the Milky Way.

In what follows, Section 2 sets up the problem and solves it. It is done in a synodic
system rather than a sidereal one, because the interesting chaotic behaviour is only seen
when the test particle is close enough to one of the heavy bodies (here the secondary star).
Section 3 describes the simulation procedures and parameter settings, where a pretty wide
range of mass ratios and separations are worked out. This is our contribution to the
literature. Section 4 presents and discusses the first part of results, namely, the static one.
Some examples of Lagrange points in the Solar system are discussed including some latest
developments. Section 5 presents and discusses the dynamics of orbits from eight-four sets
of parameters. Apart from chaotic and regular cases, three resonance cases will be touched
upon. Section 6 redoes the analysis with orbital eccentricity of the small body allowed.
Specifically, 165 cases are simulated to evaluate the old scaling law as well as comparison
with the new one. Finally, all codes used are given in the three appendices towards the end.

3
Section 2. CR3BP Setup

2.1 Notation Definitions

Consider a three-body system with a primary star, a secondary star and a test
particle, which are denoted by subscripts 𝑝, 𝑠 and nothing (no subscripts) respectively of
any variable. Construct 𝑥-axis so that centre of mass is always at origin (i.e., barycentric)
with primary and secondary stars on negative and positive sides respectively. Without loss
of generality, let the stars or in effect the 𝑥-axis rotates counter-clockwise (if viewed from
opposite side then it rotates clockwise). Such a system is called “in the co-rotation frame”.
Define the followings:

𝐺 ≡ Gravitational constant = 6.67408 × 10−11 m3 kg −1 s −2

𝑖 = 𝑝/𝑠/nil = subscript denoting primary star / secondary star / test particle

𝑚𝑖 ≡ mass of body 𝑖

𝜇 ≡ 𝑚𝑠 ⁄𝑚𝑝 = the ratio of mass of secondary star to mass of primary star

𝐱𝑖 ≡ (𝑥𝑖 , 𝑦𝑖 , 𝑧𝑖 ) = position vector of body 𝑖

𝐯𝑖 ≡ (𝑣𝑖𝑥 , 𝑣𝑖𝑦 , 𝑣𝑖𝑧 ) = velocity vector of body 𝑖

𝑟𝑖𝑗 ≡ distance bodies 𝑖 and 𝑗

𝑎 ≡ distance between primary star and secondary star

𝛿𝑎 ≡ distance between the orbit sphere of secondary star and that of test particle

𝑡 ≡ time

𝜏 ≡ the time where the three bodies are collinear with secondary star in-between

𝑃 ≡ orbital period

Ω ≡ angular velocity of the system

𝜔 ≡ angular velocity of test particle

(𝜉, 𝜂, 𝜁) ≡ sidereal coordinates of the synodic coordinates (𝑥̃, 𝑦̃, 𝑧̃ )

𝜃 ≡ test particle inclination from positive 𝑥-axis

4
𝜙 ≡ test particle inclination from positive 𝑧-axis

𝑒 ≡ orbital eccentricity of test particle

𝑁 ≡ number of rotations by test particle in the co-rotation frame

ℌ ≡ time step in fourth-order Runge-Kutta numerical integration

𝑈 ≡ potential of the system

As both stars are on 𝑥-axis where 𝑥𝑝 < 0, the followings are immediately arrived:

𝐱 𝑝 = (𝑥𝑝 , 0,0), 𝐱𝑠 = (𝑥𝑠 , 0,0) (1a)


𝑥𝑠 − 𝑥𝑝 = 𝑎 (1b)

2.2 Primary and Secondary Stars

Since the test particle does not affect other orbits, the centre of mass of the system
is determined only by the stars. From the above definition of notations, we have:

𝑚𝑝 𝐱 𝑝 + 𝑚𝑠 𝐱 𝑠 𝑚𝑖 𝑚𝑖
̃ 𝑝 𝐱̃ 𝑝 + 𝑚
≡𝑚 ̃ 𝑠 𝐱̃𝑠 = 𝟎, where 𝑚
̃𝑖 ≡ ≡ ,
𝑀𝑟𝑝𝑠 𝑀 ∑𝑖 𝑚 𝑖
(2)
𝐱𝑖 𝐱𝑖 2 2 2
𝐱̃ 𝑖 ≡ = , 𝑖 = 𝑝, 𝑠, 𝑟𝑝𝑠 ≡ ∑ (𝑗𝑝 − 𝑗𝑠 ) = 𝑎 , 𝑗 = 𝑥, 𝑦, 𝑧
𝑟𝑝𝑠 𝑎 𝑗

With (1a), the position vectors in (2) become scalar and be expressed in terms of mass ratio:

(1a) and (2) ⟹ 𝑚 ̃ 𝑠 𝑥𝑠 = 0 ⟹ 𝑥𝑝 + 𝜇𝑥𝑠 = 0, where 𝜇 ≡ 𝑚𝑠 ⁄𝑚𝑝


̃ 𝑝 𝑥𝑝 + 𝑚

With the use of (1b), the positions and masses of both stars are expressed in mass ratio:

𝜇 𝑥𝑝 𝑥𝑝 𝜇
𝑥𝑝 = − 𝑎 𝑥̃𝑝 ≡ = =−
1+𝜇 𝑟𝑝𝑠 𝑎 1+𝜇
{ 1 ⟹ 𝑥𝑠 𝑥𝑠 1 (2a)
𝑥𝑠 = 𝑎 𝑥̃𝑠 ≡ = =
1+𝜇 { 𝑟𝑝𝑠 𝑎 1+𝜇
𝑚𝑝 1 1
𝑚
̃𝑝 ≡ = =
𝑚𝑝 + 𝑚𝑠 1 + 𝑚𝑠 ⁄ 𝑚𝑝 1 + 𝜇
(2b)
𝑚𝑠 𝑚 𝑠 ⁄𝑚 𝑝 𝜇
𝑚
̃𝑠 ≡ = =
{ 𝑚𝑝 + 𝑚𝑠 1 + 𝑚𝑠 ⁄ 𝑚𝑝 1 + 𝜇

The variables with ̃ on top are in normalised-dimensionless form, where variables are
normalised by their companion total with units in numerator and denominator being offset,
and other constants being absorbed as well. Notice that 𝜇 is usually a very small number.

5
2.3 Test Particle Motion

Let the orbital sphere of test particle be slightly outside that of secondary star by
𝛿𝑎; if it is inside then 𝛿𝑎 < 0. Noting that 𝑟𝑝𝑠 = 𝑎 under our formulation, also define the
normalised-dimensionless form of position, time and period of test particle as follows:

𝐱 𝐱 𝑡 𝑡 𝑃 𝑃
𝐱̃ ≡ = , 𝑡̃ ≡ = , 𝑃̃ ≡ =
𝑟𝑝𝑠 𝑎 3⁄ √𝑎3 ⁄𝐺𝑀 3⁄ √𝑎3 ⁄𝐺𝑀 (3)
√𝑟𝑝𝑠 𝐺𝑀 √𝑟𝑝𝑠 𝐺𝑀

Given the primary and secondary stars have “no movements” in current co-rotation frame,
only the test particle motion needs to be stated. The key set of vector equations is the
classical Newtonian one which equates the centripetal force with the gravitational force:

𝑑2𝐱 𝐺𝑚𝑖 (𝐱 𝑖 − 𝐱)
𝐱̈ ≡ 2
= ∑ (4)
𝑑𝑡 |𝐱 𝑖 − 𝐱|3
𝑖=𝑝,𝑠

With the use of (2), (2b) and (3), (4) can be converted to a normalised-dimensionless form:

𝑑 2 𝐱̃ ̃ 𝑖 (𝐱̃ 𝑖 − 𝐱̃)
𝑚 ̃ 𝑖 (𝐱̃ 𝑖 − 𝐱̃)
𝑚
(2), (2𝑏), (3), (4) ⟹ 𝐱̃̈ ≡ = ∑ ≡ ∑
𝑑𝑡̃ 2 |𝐱̃ 𝑖 − 𝐱̃|3 𝑟̃𝑖3
𝑖=𝑝,𝑠 𝑖=𝑝,𝑠

𝑑2 𝑥̃ 𝑚 ̃ 𝑝 (𝑥̃𝑝 − 𝑥̃) 𝑚̃ 𝑠 (𝑥̃𝑠 − 𝑥̃) 1 1 𝜇 𝜇 1 1


𝑥̃̈ ≡ = 3 + 3 = −1+𝜇 ( 3 + 3 ) 𝑥̃ − (1+𝜇)2 ( 3 − 3 )
𝑑𝑡̃ 2 𝑟̃𝑝 𝑟̃𝑠 𝑟̃𝑝 𝑟̃𝑠 𝑟̃𝑝 𝑟̃𝑠

𝑑2 𝑦̃ 𝑚 ̃ 𝑝 (𝑦̃𝑝 − 𝑦̃) 𝑚̃ 𝑠 (𝑦̃𝑠 − 𝑦̃) 1 1 𝜇


⟹ 𝑦̃̈ ≡ = 3 + = −1+𝜇 ( 3 + 3 ) 𝑦̃ (4a)
𝑑𝑡̃ 2 𝑟̃𝑝 𝑟̃𝑠3 𝑟̃𝑝 𝑟̃𝑠

𝑑2 𝑧̃ 𝑚 ̃ 𝑝 (𝑧̃𝑝 − 𝑧̃ ) 𝑚̃ 𝑠 (𝑧̃𝑠 − 𝑧̃ ) 1 1 𝜇
𝑧̃̈ ≡ = 3 + = −1+𝜇 ( 3 + 3) 𝑧̃
{ 𝑑𝑡̃ 2 𝑟̃𝑝 𝑟̃𝑠3 𝑟̃𝑝 𝑟̃𝑠

where 𝑟̃𝑖2 = ∑𝑗(𝑗̃𝑖 − 𝑗̃)2 , 𝑖 = 𝑝, 𝑠 and 𝑗̃ = 𝑥̃, 𝑦̃, 𝑧̃ (5)


2 2 2 2
𝑟̃𝑝2 = (𝑥̃𝑝 − 𝑥̃) + (𝑦̃𝑝 − 𝑦̃) + (𝑧̃𝑝 − 𝑧̃ ) = (𝑥̃ + 1+𝜇
𝜇
) + 𝑦̃ 2 + 𝑧̃ 2
or { 2 (5a)
𝑟̃𝑠2 2 2
= (𝑥̃𝑠 − 𝑥̃) + (𝑦̃𝑠 − 𝑦̃) + (𝑧̃𝑠 − 𝑧̃ )2 = (𝑥̃ − 1
1+𝜇
2
) + 𝑦̃ + 𝑧̃ 2

The above derivation assumes primary and secondary stars are static on two points
on the 𝑥-axis; but they are actually rotating around each other so that the 𝑥-axis is indeed
rotating. Thus, (4) with (5) or (4a) with (5a) are not the true set of equations to describe the
dynamics of test particle. This will be accounted for immediately in the next sub-section.

6
2.4 Synodic System

Following Murray and Dermott (1999, Section 3.2), consider a sidereal coordinate
system (𝜉, 𝜂, 𝜁) which is in the inertial frame of the normalised co-rotation frame with
synodic coordinate system (𝑥̃, 𝑦̃, 𝑧̃ ) discussed above. Both primary and secondary stars
are on 𝜉-axis at 𝑡̃ = 0 and 𝜁-axis coincides with 𝑧̃ -axis for all 𝑡̃. Let the system rotates
̃ in the
counter-clockwise around 𝜁 -axis at a normalised constant angular velocity Ω
̃ is 2𝜋 divided by its normalised period 𝑃̃Ω . With (3):
inertial frame. By definition, Ω

2𝜋 2𝜋 2𝜋
̃≡
By (3): Ω = = =1 (6)
̃
𝑃Ω 𝑃Ω ⁄√𝑎3 ⁄𝐺𝑀 √4𝜋 2 𝑎3 ⁄𝐺𝑀⁄√𝑎3 ⁄𝐺𝑀
̃ = 1, the variable will be retained throughout which will be useful in derivation
Despite Ω
later on. Let the test particle rotates counter-clockwise at a normalised constant angular
velocity 𝜔 in the inertial frame and at 𝜔′ in the co-rotation frame. Since all bodies rotate
along the same direction, the angular velocity of test particle in the co-rotation frame is
simply the angular velocity of test particle less that of the system, both in the inertial frame:

2𝜋 2𝜋 2𝜋√𝑎3 ⁄𝐺𝑀
̃=
𝜔′ ≡ 𝜔 − Ω − = ̃
−Ω
𝑃̃ 𝑃̃Ω 𝑎
3
√4𝜋 2 (1+𝜇 + 𝛿𝑎) ⁄𝐺𝑀

3
1 𝛿𝑎 −2

⟹ 𝜔 = [( ̃
+ ) − 1] Ω (7)
1+𝜇 𝑎
̃ 𝑡̃ with respect to the
At time 𝑡̃, the synodic system has moved an angle of Ω
sidereal system. Thus the latter can be transformed to the former in the following way:

𝑥̃ cos Ω̃ 𝑡̃ ̃ 𝑡̃
sin Ω 0 𝜉
= ̃
[𝑦̃] [ − sin Ω𝑡̃ cos Ω̃ 𝑡̃ 0] [𝜂 ]
𝑧̃ 0 0 1 𝜁

Invert the matrix to work on the sidereal coordinates, which will ultimately be removed:

𝜉 ̃ 𝑡̃
cos Ω ̃ 𝑡̃ 0 −1 𝑥̃
sin Ω cos Ω ̃ 𝑡̃ ̃ 𝑡̃
− sin Ω 0 𝑥̃
⟹ [𝜂 ] = [ − sin Ω
̃ 𝑡̃ ̃
cos Ω𝑡 0 ̃ ] [ 𝑦
̃ ] = [ ̃
sin Ω𝑡̃ cos Ω̃ 𝑡̃ 0] [𝑦̃] (8)
𝜁 0 0 1 𝑧̃ 0 0 1 𝑧̃
Then differentiate (8) with respect to 𝑡̃ twice:

𝜉̇ cos Ω̃ 𝑡̃ ̃ 𝑡̃
− sin Ω 0 𝑥̃̇ − Ω ̃ 𝑦̃
̃ 𝑡̃
[𝜂̇ ] = [ sin Ω cos Ω̃ 𝑡̃ ̃ 𝑥̃]
0] [𝑦̃̇ + Ω
𝜁 ̇ 0 0 1 𝑧̃̇

7
𝜉̈ cos Ω ̃ 𝑡̃ ̃ 𝑡̃
− sin Ω 0 𝑥̃̈ − 2Ω ̃ 𝑦̃̇ − Ω
̃ 2 𝑥̃
̃ 𝑡̃
⟹ [𝜂̈ ] = [ sin Ω cos Ω̃ 𝑡̃ ̃ 𝑥̃̇ − Ω
0] [𝑦̃̈ + 2Ω ̃ 2 𝑦̃] (9)
𝜁̈ 0 0 1 𝑧̃̈
In the first two rows of rightmost vector in (9), the three terms in sequence are respectively
observed, Coriolis and centrifugal accelerations. The observed accelerations in the synodic
system (𝑥̃̈ , 𝑦̃̈) are of the same form as the sidereal system (𝜉̈ , 𝜂̈ ). The Coriolis acceleration
̃ 𝑦̃̇, +2Ω
(−2Ω ̃ 𝑥̃̇) are due to rotating frame, hence the dependence on cross-dimension terms.
̃ 2 𝑥̃, −Ω
Finally, the centrifugal accelerations (−Ω ̃ 2 𝑦̃) depend negatively on their positions.

Substitute (4a) into (9) by eliminating (𝑥̃̈ , 𝑦̃̈, 𝑧̃̈ ), the sidereal accelerations are expressed in
terms of synodic positions (𝑥̃, 𝑦̃, 𝑧̃ ) only, free of any synodic velocities and accelerations:

1 1 𝜇 𝜇 1 1 1 1 𝜇
𝜉̈ = [−1+𝜇 ( 3 + ) 𝑥̃ − (1+𝜇)2 ( − ̃ 𝑡̃ +
)] cos Ω ( + ̃ 𝑡̃
) 𝑦̃ sin Ω
𝑟̃𝑝 𝑟̃𝑠3 𝑟̃𝑝3 𝑟̃𝑠3 1+𝜇 𝑟̃𝑝3 𝑟̃𝑠3
1 1 𝜇 𝜇 1 1 ̃ 𝑡̃ − 1 1 𝜇 ̃ 𝑡̃
𝜂̈ = [−1+𝜇 ( + ) 𝑥̃ − (1+𝜇)2 ( − )] sin Ω ( + ) 𝑦̃ cos Ω (10)
𝑟̃𝑝3 𝑟̃𝑠3 𝑟̃𝑝3 𝑟̃𝑠3 1+𝜇 𝑟̃𝑝3 𝑟̃𝑠3
1 1 𝜇
𝜁̈ = −1+𝜇 ( + ) 𝑧̃
{ 𝑟̃𝑝3 𝑟̃𝑠3

Equate (9) and (10), the two versions of sidereal acceleration (𝜉̈ , 𝜂̈ , 𝜁 ̈) but itself eliminated:

̃ 𝑦̃̇ − Ω
(𝑥̃̈ − 2Ω ̃ 2 𝑥̃) cos Ω
̃ 𝑡̃ − (𝑦̃̈ + 2Ω
̃ 𝑥̃̇ − Ω
̃ 2 𝑦̃) sin Ω
̃ 𝑡̃
1 1 𝜇 𝜇 1 1 ̃ 𝑡̃ + 1 1 𝜇 ̃ 𝑡̃
= [−1+𝜇 ( + ) 𝑥̃ − (1+𝜇)2 ( − )] cos Ω ( + ) 𝑦̃ sin Ω
𝑟̃𝑝3 𝑟̃𝑠3 𝑟̃𝑝3 𝑟̃𝑠3 1+𝜇 𝑟̃𝑝3 𝑟̃𝑠3

̃ 𝑦̃̇ − Ω
(𝑥̃̈ − 2Ω ̃ 2 𝑥̃) sin Ω
̃ 𝑡̃ + (𝑦̃̈ + 2Ω
̃ 𝑥̃̇ − Ω
̃ 2 𝑦̃) cos Ω
̃ 𝑡̃
1 1 𝜇 𝜇 1 1 ̃ 𝑡̃ − 1 1 𝜇 ̃ 𝑡̃
= [−1+𝜇 ( + ) 𝑥̃ − (1+𝜇)2 ( − )] sin Ω ( + ) 𝑦̃ cos Ω
𝑟̃𝑝3 𝑟̃𝑠3 𝑟̃𝑝3 𝑟̃𝑠3 1+𝜇 𝑟̃𝑝3 𝑟̃𝑠3

1 1 𝜇
𝑧̃̈ = −1+𝜇 ( + ) 𝑧̃
𝑟̃𝑝3 𝑟̃𝑠3

Since the time variable 𝑡̃ appears only in sine/cosine terms, the coefficient of either must
̃ 𝑡̃ and
be equal on both sides of the first two equations above. Match coefficients of cos Ω
̃ 𝑡̃ from either of the first two equations gives the (same new set of conditions):
sin Ω

1 1 𝜇 𝜇 1 1
̃ 𝑦̃̇ − Ω
𝑥̃̈ − 2Ω ̃ 2 𝑥̃ = − ( 3 + 3 ) 𝑥̃ − ( 3 − 3)
1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠 2
(1 + 𝜇) 𝑟̃𝑝 𝑟̃𝑠
1 1 𝜇
̃ 𝑥̃̇ − Ω
𝑦̃̈ + 2Ω ̃ 2 𝑦̃ = − ( 3 + 3 ) 𝑦̃ , or
1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠
1 1 𝜇
𝑧̃̈ =− ( 3 + 3 ) 𝑧̃
{ 1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠

8
1 1 𝜇 𝜇 1 1
̃ 𝑣̃𝑦 + [Ω
𝑣̃̇𝑥 = 2Ω ̃2 − ( 3 + 3 )] 𝑥̃ − ( 3 − 3)
1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠 2
(1 + 𝜇) 𝑟̃𝑝 𝑟̃𝑠
1 1 𝜇
𝑣̃̇𝑦 ̃ 𝑣̃𝑥 + [Ω
= −2Ω ̃2 − ( 3 + 3 )] 𝑦̃ (11)
1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠
1 1 𝜇
𝑣̃̇𝑧 = − ( 3 + 3 ) 𝑧̃
{ 1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠
(11) is then mapped to a vector form by stacking 𝐱̃ and 𝐯̃ as 𝐗: the upper half is
trivial while the lower half is just (11). This will be the key set of equations in simulation:

𝑥̃ 𝑣̃𝑥

𝑦̃ 𝑣̃𝑦
𝑣̃𝑧
𝑧̃ 1 1 𝜇 𝜇 1 1
𝐗̇ ≡ 𝐅, 𝐗 ≡ ,𝐅 = ̃ 𝑣̃𝑦 + [Ω
2Ω ̃2 − ( + )] 𝑥̃ − (1+𝜇)2 ( − ) (12)
1+𝜇 𝑟̃𝑝3 𝑟̃𝑠3 𝑟̃𝑝3 𝑟̃𝑠3
𝑣̃𝑥
̃ 2 − 1 ( 13 +
̃ 𝑣̃𝑥 + [Ω
−2Ω
𝜇
)] 𝑦̃
1+𝜇 𝑟̃𝑝 𝑟̃𝑠3
𝑣̃𝑦
1 1 𝜇
−1+𝜇 ( + ) 𝑧̃
[ 𝑣̃𝑧 ] [ 𝑟̃𝑝3 𝑟̃𝑠3 ]

2.5 Initial Conditions

Before closing the setup, initial conditions or 𝑡̃ = 0 form of (12) are needed; these
variables are denoted by an additional subscript 0. Consider the test particle in co-rotation
frame initially inclined at 𝜙0 from positive 𝑧 -axis and at 𝜃0 from positive 𝑥 -axis
counting counter-clockwise on 𝑥𝑦-plane. From (2), its original and normalised forms are:

1
𝑥0 = ( 𝑎 + 𝛿𝑎) sin 𝜙0 cos 𝜃0
1+𝜇
1
𝑦0 = ( 𝑎 + 𝛿𝑎) sin 𝜙0 sin 𝜃0
1+𝜇
1
𝑧0 = ( 𝑎 + 𝛿𝑎) cos 𝜙0
{ 1+𝜇

𝑥0 𝑥0 1 𝛿𝑎
𝑥̃0 ≡
= =( + ) sin 𝜙0 cos 𝜃0
𝑟𝑝𝑠 𝑎 1+𝜇 𝑎
𝑦0 𝑦0 1 𝛿𝑎
⟹ 𝑦̃0 ≡ = =( + ) sin 𝜙0 sin 𝜃0 (13)
𝑟𝑝𝑠 𝑎 1+𝜇 𝑎
𝑧0 𝑧0 1 𝛿𝑎
𝑧̃0 ≡ = =( + ) cos 𝜙0
{ 𝑟𝑝𝑠 𝑎 1+𝜇 𝑎
Here confines to the special case where test particle rotates on 𝑥̃𝑦̃-plane, i.e., a coplanar
system and starts from positive 𝑥̃-axis. Set 𝜙0 = 𝜋⁄2 and 𝜃0 = 0 in (13) as special case:

9
1 𝛿𝑎 𝜋 1 𝛿𝑎
𝑥̃0 = (
+ ) sin cos 0 = +
1+𝜇 𝑎 2 1+𝜇 𝑎
𝜋
𝜙 = 1 𝛿𝑎 𝜋
Special case of { 0 2 : 𝑦̃0 = ( + ) sin sin 0 = 0 (13a)
𝜃0 = 0 1+𝜇 𝑎 2
1 𝛿𝑎 𝜋
𝑧̃0 = ( + ) cos =0
{ 1+𝜇 𝑎 2
The initial velocity in the co-rotation frame is orbital angular velocity from (7) times radius,
which is simply the initial position. The velocities for both general and coplanar cases are:

𝐯̃0 = 𝜔′𝐱̃ 0

1
′ 1 𝛿𝑎 −2 1 𝛿𝑎 ̃ sin 𝜙0 sin 𝜃0
𝑣̃𝑥,0 = 𝜔 𝑥̃0 = − [(1+𝜇 + 𝑎
) − (1+𝜇 + 𝑎
)] Ω
1
′ 1 𝛿𝑎 −2 1 𝛿𝑎 ̃ sin 𝜙0 cos 𝜃0
⟹ 𝑣̃𝑦,0 = 𝜔 𝑦̃0 = [(1+𝜇 + 𝑎
) − (1+𝜇 + 𝑎
)] Ω (14)
1
1 𝛿𝑎 −2 1 𝛿𝑎 ̃ cos 𝜙0
𝑣̃𝑧,0 = 𝜔′ 𝑧̃0 = [(1+𝜇 + 𝑎
) − (1+𝜇 + 𝑎
)] Ω
{
𝑣̃𝑥,0 = 0
𝜋 1
𝜙0 = 1 𝛿𝑎 −2 1 𝛿𝑎 ̃
Special case of { 2 : 𝑣̃𝑦,0 = [(1+𝜇 + 𝑎 ) − (1+𝜇 + 𝑎 )] Ω (14a)
𝜃0 = 0
{ 𝑣̃𝑧,0 = 0
The differential vector equations (12), initial conditions (13a) and (14a), with the
̃ from (6) are the key set of equations used in simulation to generate the results.
value of Ω
The numerical solution to (12) involves the numerical integration of vector 𝐅 given input
𝐗, which is the most challenging part and will be discussed immediately in the next section.

10
Section 3. Simulation

3.1 Numerical Integration

There are various ways to handle the first-order ordinary differential equations with
initial value problem given in the previous section. The discussion here draws mainly from
Press et al (2007, Chapter 17). As they put it, for general formulation 𝑑𝑦⁄𝑑𝑥 = 𝑧(𝑥):

The underlying idea of any routine for solving the initial value problem
is always this: Rewrite the 𝑑𝑦’s and 𝑑𝑥’s… as finite steps ∆𝑦 and ∆𝑥,
and multiply the equations by ∆𝑥. This gives algebraic formulas for the
change in the functions when the independent variable 𝑥 is “stepped”
by one “stepsize” ∆𝑥. In the limit of making the stepsize very small, a
good approximation to the underlying differential equation is achieved.
Literal implementation of this procedure results in Euler’s method…,
which is, however, not recommended for any practical use. Euler’s
method is conceptually important, however; one way or another,
practical methods all come down to this same idea: Add small
increments to your functions corresponding to derivatives… multiplied
by stepsizes. Press et al (2007, Chapter 17, Page 900)

The Euler’s method is actually the repeated updating of integral 𝐗 𝑛 from the given
integrand 𝐹(𝑡̃𝑛 , 𝐗 𝑛 ) between 𝑡̃𝑛 and 𝑡̃𝑛+1 over the step size ℌ from following formula:

𝐗 𝑛+1 = 𝐗 𝑛 + ℌ𝐹(𝑡̃𝑛 , 𝐗 𝑛 ), where ℌ ≡ 𝑡̃𝑛+1 − 𝑡̃𝑛

However, in this setting only the beginning of interval is used to obtain the integral, the
information of other parts of interval 𝐹(𝑡̃𝑛+1 , 𝐗 𝑛+1 ) is wasted. There are ways to remedy
this; one of the most popular approaches is done via Runge-Kutta method which utilises
information of the interval midpoint. For second-order Runge-Kutta the following is run:

𝐤1 = ℌ𝐹(𝑡̃𝑛 , 𝐗 𝑛 )
ℌ 𝐤1
{𝐤 2 = ℌ𝐹 (𝑡̃𝑛 + 2 , 𝐗 𝑛 + 2
)
𝐗 𝑛+1 = 𝐗 𝑛 + 𝐤 2 + 𝐎(ℎ3 )

Here 𝐤1 is the ∆𝐗 𝑛 obtained in the first round. Then half of this and half of the ∆𝑡̃𝑛 , i.e.,
(𝑡̃𝑛 + ℌ2, 𝐗 𝑛 + 𝐤21) are input again to obtain ∆𝐗 in the second round, which embeds the
information of beginning as well as midpoint of the interval to yield 𝐗 𝑛+1 . To further
enhance accuracy, the above procedure can be repeated one more time by splitting the two
sub-intervals into four. This is the famous fourth-order Runge-Kutta method but derivation
is tedious which can be found in say, Musa et al (2010). The algorithm runs as follows:

11
𝐤1 = ℌ𝐹(𝑡̃𝑛 , 𝐗 𝑛 )
ℌ 𝐤
𝐤2 = ℌ𝐹 (𝑡̃𝑛 + 2 , 𝐗 𝑛 + 21 )
ℌ 𝐤
𝐤3 = ℌ𝐹 (𝑡̃𝑛 + 2 , 𝐗 𝑛 + 22 ) (15)
𝐤4 = ℌ𝐹(𝑡̃𝑛 + ℌ, 𝐗 𝑛 + 𝐤 3 )
𝐤1 𝐤2 𝐤3 𝐤4
{ 𝐗 𝑛+1 = 𝐗 𝑛 + 6
+ 3
+ 3
+ 6
+ 𝐎(ℌ5 )

In practice, given the explicit integrand 𝐹, 𝐤 𝑖 ’s are obtained in sequence and thus
the integral 𝐗 𝑛 (integrated from 𝐹) is updated to 𝐗 𝑛+1 . The smaller the time step ℌ, the
better is the accuracy of numerical integration and 𝐎(ℌ5 ) is small enough to be neglected.

3.2 Coding

As mentioned previously, the differential vector equations (12) are the key set to
generate the dynamics. These involve the integration of 𝐹 to obtain 𝐗 which is done by
the fourth-order Runge-Kutta method of numerical integration in (15) just discussed above.
C++ is used for its speed advantage; in spite of this, the longest case takes quite some
hours to retrieve the results. The full code for one of the cases is given in Appendix A. The
differential vector equations are formulated as function “F” in the first section of the code;
then are solved by the fourth-order Runge-Kutta numerical integration as function “RK4”
in the second section. Parameters are set in the main part of code in the final section.

In the main part, a full path, name and extension of a file needs to be specified
where output will be generated. As there is no preinstalled plotter in C++, the retrieved
data are plotted in Excel and a freeware gnuplot. Four columns of data are simulated, these
are in order (𝑥̃𝜏 − 𝑥̃0 , 𝑣̃𝑥,𝜏 , 𝑥̃𝜏 , 𝑦̃𝜏 ). The first two are 𝑥̃-component position (adjusted by
initial value) and velocity of test particle whenever it crosses positive 𝑥̃-axis at time 𝜏.
These will be plotted as figures to demonstrate the chaotic and regular behaviour. The last
two columns are 𝑥̃- and 𝑦̃-coordinates of the test particle position, which serve to check
whether it is still under circular orbit. In some cases the locus diverges from circular orbit
after a certain number of rotations, then only the orbital part is plotted. The rows are the
data points simulated at various time 𝑡̃. Constant time interval applies throughout.

12
3.3 Parameters

In our coplanar case, the differential vector equations (12) are in effect reduced to
four as the 𝑧̃ -component position and velocity drop out; hence is the number of equations
𝑛eq in coding. The number of rotations by test particle in the co-rotation frame 𝑁 is the
number of simulated data points expected to be retrieved; it is found from trials that about
1000 is needed to generate stable chart pattern. The final global parameter is time-step in
Runge-Kutta looping, ℌ; it is found that about 0.01 is small enough to obtain accurate
integration results. It follows that the total time travelled by test particle in the co-rotation
frame 𝑡̃max is the total distance travelled per cycle multiplied by the number of rotations:

𝑡̃max = 2𝜋𝑁⁄|𝜔′ | (16)


The absolute value sign appears in the denominator as the angular velocity can be negative
if the test particle rotates clockwise in the co-rotation frame, i.e., its angular speed in the
inertial frame is slower than that of the system. In concluding, the global parameters are:

𝜙 0 = 𝜋 ⁄2 , 𝜃0 = 0, 𝑁 = 1000, ℌ = 0.01 (17)


𝜙0 = 𝜋⁄2 and 𝜃0 = 0 refer to a coplanar system that starts from positive 𝑥̃-axis.

Other parameters vary from case to case. Wisdom (1980) suggests in his last
equation (56) that for 𝜇 = 10−3 , 10−4 and 10−5 , 𝑠overlap ≡ (𝛿𝑎⁄𝑎)−1
overlap ≃ 0.51𝜇
−2⁄7

is the critical value: when greater than this resonances begin to overlap which lead to a
band-like chaotic orbit or “stochastic behaviour” as he termed, and when smaller than this
regular behaviour exhibits. This can be mapped into our context by inverting the formula,
where value smaller than critical gives chaotic result and larger than gives a regular one:

(𝛿𝑎⁄𝑎)overlap = 𝑘𝜇 2⁄7 , where 𝑘 = 1⁄0.51 = 1.96 by Wisdom (1980) (18)


Take log to his original formula (56) and rearrange, the criterion is charted in his FIG.16:

7 7
− log 𝜇 = −2 log 0.51 + 2 log 𝑠overlap = 1.02 + 3.5 log 𝑠overlap (19)
3 0.56
For − log 𝜇 = {4 , ⟹ log 𝑠overlap = {0.85
5 1.14

The three values of − log 𝜇 are those on the vertical axis, and the corresponding values
log 𝑠overlap on the horizontal axis projected by the scaling law, i.e., the dashed line of (19):

13
Wisdom (1980, FIG.16)

It is clear that the three values of log 𝑠overlap implied by scaling law are slightly
larger than those given by resonance overlap criterion where 𝑘 = 1.307. Yet some quote
even slightly higher values. For example, Malhotra (1998) quotes 1.40, Deck et al. (2013)
obtain 1.46 while Duncan et al. (1989) yield 1.49. In light of the reported wide range from
1.3 to 2.0, here covers the whole range with 0.3 more on each side: 𝑘 = 1.0, 1.1, … , 2.3.

Having discussed the adoption of 𝑘 in (18), now comes to 𝜇. (18) is valid only for
small 𝜇, but what range should it take? The upper limit studied in Wisdom (1980)
corresponds to the case of largest planet in Solar system: Jupiter-to-Sun ratio is up to
~1.0 × 10−3 . For minor body such as the largest dwarf/minor planet however, like Pluto-
to-Sun ratio it can be as low as ~0.7 × 10−8 (the smallest minor body like meteorite has
no effective lower limit). As a result, a range of 𝜇 from 10−8 to 10−3 is adopted here.

Putting together, (16) to (18) in above and (20) below constitute all parameters used:

𝜇 = 10−8 , 10−7 , … , 10−3 , 𝑘 = 1.0, 1.1, … , 2.3 (20)

14
Section 4. Static Results

4.1 Lagrange Points

Before discussing the behaviour of the test particle orbit, first study the equilibrium.
There are five points where the test particle is in force equilibrium, i.e., no movements
over time. These are the famous Lagrange points, and are marked as red dots L1 to L5 in
the below figure, where primary and secondary stars are in yellow and violet respectively:

Here L1 to L3 are unstable but L4 and L5 are stable, for reason to be given shortly.
If the primary, secondary stars and test particles are respectively the Sun, the Earth and an
artificial satellite, then these points will be of great interest for their relative static positions.
L2 is the best place as it is closest to the Earth and being shielded from sunlight, however,
it is now too crowded to add further satellites there. L4 and L5 become a way out for this.

First is to locate these two points: L4 and L5 incline at 60° with the positive 𝑥̃-axis,
or the primary and secondary stars with either L4 or L5 constitutes an equilateral triangle.
To see this, recall the definition of force equilibrium means there are neither motions nor
accelerations acted upon the test particle. Denote the Lagrange points by subscript L. Then:

𝐱̃̇ 𝐿 = 0 𝑥̃̇ = 𝑦̃̇𝐿 = 𝑧̃̇𝐿 = 0


Lagrange points equilibrium conditions: { ⟹ { 𝐿
𝐱̃̈ 𝐿 = 0 𝑥̃̈𝐿 = 𝑦̃̈𝐿 = 𝑧̃̈𝐿 = 0

Set these conditions in (11), and notice that this reduces the context to a coplanar case:

1 1 𝜇
𝑧̃̈𝐿 = 0 ⟹ − ( 3 + 3 ) 𝑧̃𝐿 = 0 ⟹ 𝑧̃𝐿 = 0
1 + 𝜇 𝑟̃𝑝,𝐿 𝑟̃𝑠,𝐿

1 1 𝜇 𝜇 1 1
̃ 2 𝑥̃𝐿 = −
−Ω ( 3 + 3 ) 𝑥̃𝐿 − ( 3 − 3) (21a)
1 + 𝜇 𝑟̃𝑝,𝐿 𝑟̃𝑠,𝐿 2
(1 + 𝜇) 𝑟̃𝑝,𝐿 𝑟̃𝑠,𝐿

15
1 1 𝜇
̃ 2 𝑦̃𝐿 = −
−Ω ( 3 + 3 ) 𝑦̃𝐿 (21b)
1 + 𝜇 𝑟̃𝑝,𝐿 𝑟̃𝑠,𝐿
̃2:
Now 𝑦̃𝐿 ’s on both sides of (21b) cancel out, substitute this into (21a) by eliminating −Ω

𝜇 1 1
− 2
( 3 − 3 ) =0 ⟹ 𝑟̃𝑝,𝐿 = 𝑟̃𝑠,𝐿 (22)
(1 + 𝜇) 𝑟̃𝑝,𝐿 𝑟̃𝑠,𝐿
̃ = 1, the following is arrived:
Substitute (22) into (21b) again, and noting that (6) implies Ω

𝑟̃𝑝,𝐿 = 𝑟̃𝑠,𝐿 = 1

From (2): 𝑟̃𝑝𝑠 ≡ 𝑟𝑝𝑠 ⁄𝑎 = 1 ⟹ 𝑟̃𝑝𝑠,𝐿 = 1 ⟹ 𝑟̃𝑝,𝐿 = 𝑟̃𝑠,𝐿 = 𝑟̃𝑝𝑠,𝐿

Thus, the primary and secondary with test particle on L4 or L5 forms an equilateral triangle,
and these are known as triangular Lagrangian equilibrium points, which are in contrast to
L1 to L3, the collinear Lagrangian equilibrium points. For L1 to L3 on 𝑥̃-axis, the position
can be solved by setting 𝑦̃𝐿 = 0 in (21a), but now 𝑟̃𝑝,𝐿 ≠ 𝑟̃𝑠,𝐿 and both contain 𝑥̃𝐿2 terms
in the denominator of (21a), rending it highly nonlinear in 𝑥̃𝐿 and better solve numerically.

4.2 Equilibrium Potential

In previous section we mentioned L1 to L3 are unstable but L4 and L5 are stable.


To see this, first define two scalar functions 𝑈 ∗ (𝑥, 𝑦, 𝑧) ≡ −𝑈(𝑥, 𝑦, 𝑧) as follows:

𝐺𝑚𝑝 𝐺𝑚𝑠 Ω2 2
𝑈 ∗ (𝑥, 𝑦, 𝑧) ≡ −𝑈(𝑥, 𝑦, 𝑧) ≡ − [ + + (𝑥 + 𝑦 2 )] (23)
𝑟𝑝 𝑟𝑠 2
Inside the bracket is the sum of gravitational (first two terms) and centrifugal potentials
(last term). The 𝑈 ∗ defined in negative way is purely a convention in celestial mechanics.
To check the potential of five Lagrange points, first normalised everything by 𝐺𝑀⁄𝑎:

2
𝑈∗ 𝑚 ⁄𝑀 𝑚𝑠 ⁄𝑀 (Ω√𝑎3 ⁄𝐺𝑀) 𝑥 2 𝑦 2
̃∗ ≡
Define 𝑈 ̃∗ = − 𝑝
⟹ 𝑈 − − [( ) + ( ) ]
𝐺𝑀⁄𝑎 𝑟𝑝 ⁄𝑎 𝑟𝑠 ⁄𝑎 2 𝑎 𝑎

𝑚
̃𝑝 𝑚 ̃𝑠 Ω ̃2
̃ ∗ (𝑥̃, 𝑦̃) = −
⟹ 𝑈 − − (𝑥̃ 2 + 𝑦̃ 2 ) = −𝑈(𝑥̃, 𝑦̃) in (21)
𝑟̃𝑝 𝑟̃𝑠 2

̃ ∗ is dimensionless as all units of masses in numerator and those of distances in


Now 𝑈
denominator cancelled out each other under normalisation. With the use of (2b), (5a) and
(6), also noting that 𝑧̃ = 0, the potential can be expressed in terms of (𝑥̃, 𝑦̃) and 𝜇 only:

16
2 −12 −12
̃ ∗ (𝑥̃, 𝑦̃) = − 1 [(𝑥̃ + 𝜇 ) + 𝑦̃ 2 ] 𝜇 1 2 2 1
𝑈 1+𝜇 1+𝜇
− 1+𝜇
[(𝑥̃ − 1+𝜇
) + 𝑦̃ ] − 2(𝑥̃ 2 + 𝑦̃ 2 )

𝑚𝑠 1 𝜇
For 𝜇 ≡ = 0.25, = 0.8, = 0.2
𝑚𝑝 1+𝜇 1+𝜇

1 1
̃ ∗ (𝑥̃, 𝑦̃)𝜇=0.25 = −0.8[(𝑥̃ + 0.2)2 + 𝑦̃ 2 ]−2 − 0.2[(𝑥̃ − 0.8)2 + 𝑦̃ 2 ]−2 − 0.5(𝑥̃ 2 + 𝑦̃ 2 )
⟹ 𝑈

The above function is plotted below, which is done in gnuplot with code in Appendix B.

One can see there are two wells on the surface; the deeper one is at (−0.2,0) with
heavier weight 0.8 while the shallower one is at (0.8,0) with lighter weight 0.2. From (2a),
these are the positions of primary and secondary stars respectively. As L1 to L3 are all on
𝑥̃-axis, they are located at the top edge of these wells where L1 is in-between the two wells,
L2 on the right-hand side and L3 on the left-hand side. Since all three points are on the top
of ∩-shape along 𝑥̃-direction and at the bottom of ∪-shape along 𝑦̃-direction, they are
saddle points which are unstable equilibria. For L4 and L5, they are located at the top of
the surface (in fact the global max) yielding the potential maximum. Simple geometric
calculation concludes their positions are at (0.3, √3⁄2) and (0.3, − √3⁄2) respectively.

The above equilibrium stability conclusion for L4 and L5 is generally true for most
practical values of 𝜇. Murray and Dermott (1999, Section 3.7) linearise the equations of
motion and carry out a linear stability analysis, then conclude the following condition:

𝜇 1
Stability condition for 𝐿4 and 𝐿5 : ≤ 0.0385 ⟹ 𝜇 ≤ 0.04 =
1+𝜇 24.97

That says, the mass of primary star has to be at least 25 times of the secondary star
for L4 and L5 to be stable. Practically speaking, this condition is trivially satisfied for most
cases in the Solar system even when planet is primary and its satellite is secondary.
17
4.3 Horseshoe and Tadpole Orbits

The final analysis to be conducted in this section is the motion around L4 and L5,
which bring out the famous horseshoe orbit. Consider a test particle motion near the Sun-
Jupiter’s L4 and L5, with initial conditions taken from Taylor (1981, Table 1, Orbit a):

𝑥̃0 = −1.06201
𝑚𝐽 𝑦̃0 = 0
𝜇≡ = 9.53875 × 10−4 , , 𝑡̃max = 200
𝑚⊙ 𝑥̃̇0 = 0
{ 𝑦̃̇0 = 0.10851

1 1
By (2𝑎), 𝑥̃𝑠 = = = 0.999047
1 + 𝜇 1 + 9.53875 × 10−4

With these initial conditions, the test particle orbit is simulated again in C++ with
code given in Appendix C (simplified version of the one in Appendix A). The positions of
the primary and secondary stars could be easily obtained from (1) and (2a). For L4 and L5,
the coordinates are also readily identified given the property of equilateral triangle:

𝜋 𝜇 𝜋 9.53875 × 10−4 1
𝑥̃𝐿 = 𝑥̃𝑝 + cos= + cos = −4
+ = 0.499
3 1+𝜇 3 1 + 9.53875 × 10 2
𝜋 √3
𝑦̃𝐿 = ± sin = ± = ±0.866
{ 3 2

The orbit with positions of L4, L5, primary and secondary stars, are all shown in Orbit 1:

The solution to the orbit involves solving a fourth-order characteristic equation


resulting in some sine/cosine terms in time, which is the source of sausage-shape receded
libration. But for particular initial position, say, taking Taylor (1981, Table 1, Orbit b):

18
𝑥̃0 = −1.02745, 𝑦̃0 = 0, 𝑥̃̇0 = 0, 𝑦̃̇0 = 0.04032
where other parameters being unchanged. Then a smooth horseshoe is obtained, as shown
in Orbit 2 above, where the epicyclic motion is almost completely suppressed. This is in
fact the orbit where a maximum value of Jacobi constant (to be discussed in the following
subsection) is attained, when the system potential is highest with respect to test particle’s
motion (velocity squared) –– a so-called most stable orbit under the horseshoe family.

Figure Adopted from Wikipedia:


https://upload.wikimedia.org/wikipedia/commons/d/d6/Lagrange_Horseshoe_Orbit.jpg

The key feature of the horseshoe orbit is that it embeds both L4 and L5 which can
be explained below. Refer to the above figure. Suppose the test particle starts at point A,
the inner part near L5 as well as L1. As L5 is a potential maximum and any neighbourhood
is having lower potential, forces are exerted outward from L5 and the test particle is
squeezed out. However, it moves towards rather than away from L1 as it rotates relatively
faster than the system along inner orbit with shorter radius in the co-rotation frame.

Recall the saddle point property of L1 just discussed which is repulsive (∩-shape)
along 𝑥̃-direction and attractive (∪-shape) along 𝑦̃-direction, the particle would turn right
(attracted to north but repulsed to east) until it moves to the outer orbit and finally U-turns.
At point B, there is no net force acting along 𝑥̃-direction, so that the test particle moves
from B to C under inertia. Now the test particle is along outer orbit moving relatively
slower than the system in co-rotation frame and “receding” to D. Same argument applies
from D to E (as from A to C) and from E to A. Thus, the whole cycle is explained.

There are companion to the horseshoe orbit if the initial separation from L4 or L5 is
further reduced a bit. Taking the parameters from Murray and Dermott (1999, Fig. 3.16),
here replicate both cases with one slightly near L4 and another slightly farther away:

19
0.001 𝑥̃̇ = 0
For both cases: 𝜇 = , { 0 , where − 𝑥̃𝑝 = 𝑚
̃ 𝑠 = 0.001
1 + 0.001 𝑦̃̇0 = 0

𝑥̃0 = 1⁄2 − 0.001 + 0.0065 𝑥̃0 = 1⁄2 − 0.001 + 0.008


Nearer case: { 𝑦̃0 = √3⁄2 + 0.0065 , Farther case: { 𝑦̃0 = √3⁄2 + 0.008
𝑡̃max = 15 × 2𝜋 𝑡̃max = 15.5 × 2𝜋

In above, L4 is located at (1⁄2 − 0.001, √3⁄2). The two initial positions are distanced
respectively at 0.009 and 0.011 away from L4, differed only by about ~20%. As the codes
are very similar to the horseshoe cases in Appendix C, they are therefore omitted here.

Both cases result in tadpole orbits, with the farther case spanning a more elongated arc.
Compared with horseshoe orbits, Murray and Dermott (1999, Section 3.11) find that the
⁄3 ⁄
̃ 𝑠1
width of the horseshoe region is ~𝑚 ̃ 𝑠1 2 .
whereas that if the tadpole region is ~𝑚

4.4 Solar System Examples

In our Solar system there is a typical example of tadpole and horseshoe orbits,
which are those respectively by Saturn’s satellites Janus (Saturn X) and Epimetheus
(Saturn XI). Given their masses (with Saturn’s) and semi-major axes obtained from online:

𝑚𝐽
𝑚𝑆 = 5.6834 × 1026 kg 𝜇𝐽 ≡ = 3.34 × 10−9
𝑚𝑆 𝑎𝐽 = 151460km
{ 𝑚𝐽 = 1.8975 × 1018 kg ⟹ { 𝑚𝐸 and {
𝜇𝐸 ≡ = 9.27 × 10−10 𝑎𝐸 = 151410km
𝑚𝐸 = 5.2660 × 1017 kg 𝑚𝑆

The libration widths of both are given in Murray and Dermott (1999, Section 3.10) as:

20
8 1⁄2 12
tadpole 𝛿𝑎𝐽 = ( ) 𝜇𝐽 𝑎𝐽 = 14.3km
Libration width of { orbit: 3
horseshoe 0.02 1⁄2 13
𝛿𝑎𝐸 = 2 ( ) 𝜇𝐸 𝑎𝐸 = 24.1km
{ 3

These are too thin to be observed such that the orbital types were empirically confirmed
until relatively recently. Yet the conjecture was made far earlier given the two satellites
sharing almost the same orbit: 𝑎𝐽 ≈ 𝑎𝐸 . But since the masses of them are comparable
(𝑚𝐽 = 3.6𝑚𝐸 ), some kind of mutual perturbations is expected. It turns out that they orbit
on each side and rebound upon meeting each other. With conservation of the total orbital
angular momentum, their widths of arc in angle ϑ are given by this simple relationship:

𝑚𝐽 ϑ𝐽 = 𝑚𝐸 ϑ𝐸 , where ϑ𝐽 + ϑ𝐸 = 360°

𝑚𝐸 °
5.2660 × 1017 × 360°
⟹ ϑ𝐽 = 360 = 18 17
= 78° , ϑ𝐸 = 282°
𝑚𝐽 + 𝑚𝐸 1.8975 × 10 + 5.2660 × 10

Tiscareno et al (2009, Fig.1) illustrates the above orbital descriptions, where points 1, 2
and 3 are positions of 1st July 2004, 21st May 2005 and 9th September 2006 respectively:

Tiscareno et al (2009, Fig.1)

Image from Ohio State University

21
For L4 and L5 in the Solar system, a classic example is Jupiter and its two co-
rotating clusters –– Greeks in L4 and Trojans in L5, as shown in the above image adopted
from Ohio State University. Despite these two are supposed to be on two points, both span
wide arcs and have comparable number of objects, more than 1 million on each swamp.

Over the years many objects in the Lagrange points have been identified in the
Solar system, some in Solar-planet and others in planet-satellite context. Until recently,
Slíz-Balogh et al (2019) confirmed two more dust moons in the Solar-Earth L5 point, the
Kordylewski clouds (KDC). By applying imaging polarimetry they observed two dust
clusters inclined at 73.0° and 87.3° between the Earth and the Moon, both in L5 region:

Slíz-Balogh et al (2019, Figures 1 and 2)

From the pictures above, the lighter orange regions denote a specific range of angle
which is polarised by dust cluster. The two pictures denote the two clusters respectively.

As a final remark before closing this section, it is evident that millions of trojans
are on the Jupiter orbit. While Saturn is also a gas giant (~30% of Jupiter mass), it does not
have any known Trojan asteroids. For the remaining planets in mass order, Neptune has 22
trojans, Uranus has 2, both Earth and Venus have 1, and Mars has 9. Hou et al (2014)
suggest it is secular resonances in Saturn that leads to the absence of trojans in L4 and L5.
The resonances and chaotic behaviour will be discussed immediately in the next section.

22
Section 5. Dynamic Results

5.1 Orbital Plots

It is mentioned previously that (𝑥̃𝜏 − 𝑥̃0 , 𝑣̃𝑥,𝜏 ) will be plotted to check whether the
orbit is chaotic or regular; the rationale of doing so can be explained by manipulating the
function 𝑈 introduced earlier. With the use of (2b) and (5a), (23) can be re-written as:

1 1
−2 −2
̃ 2 (𝑥̃ 2 +𝑦̃ 2 )
Ω 1 𝜇 2 2 2 1 2 2 2
𝑈= 2
+ 1+𝜇
{[(𝑥̃ + 1+𝜇
) + 𝑦̃ + 𝑧̃ ] + 𝜇 [(𝑥̃ − 1+𝜇
) + 𝑦̃ + 𝑧̃ ] } (23a)

Then differentiate 𝑈 partially with respect to all its arguments, the followings are arrived:

1 𝜇 𝜇 1
𝜕𝑈 1+𝜇
(𝑥̃ + 1+𝜇 ) 1+𝜇
(𝑥̃ − 1+𝜇 )
𝑈𝑥̃ ≡ ̃ 2 𝑥̃ −
=Ω +
𝜕𝑥̃ 2 3⁄ 2 2 3⁄2
𝜇 1
[(𝑥̃ + 1+𝜇 ) + 𝑦̃ 2 + 𝑧̃ 2 ] [(𝑥̃ − 1+𝜇 ) + 𝑦̃ 2 + 𝑧̃ 2 ]
{ }
1 𝜇
𝜕𝑈 1+𝜇 1+𝜇
⟹ 𝑈𝑦̃ ≡ ̃ 2 𝑦̃ −
=Ω + 𝑦̃
𝜕𝑦̃ 2 3⁄ 2 2 3⁄2
𝜇 1
[(𝑥̃ + 1+𝜇 ) + 𝑦̃ 2 + 𝑧̃ 2 ] [(𝑥̃ − 1+𝜇 ) + 𝑦̃ 2 + 𝑧̃ 2 ]
{ }
1 𝜇
𝜕𝑈 1+𝜇 1+𝜇
𝑈𝑧̃ ≡ = − 3⁄ 2
+ 3⁄2
𝑧̃
𝜕𝑧̃ 𝜇
2
1
2
[(𝑥̃ + 1+𝜇 ) + 𝑦̃ 2 + 𝑧̃ 2 ] [(𝑥̃ − 1+𝜇 ) + 𝑦̃ 2 + 𝑧̃ 2 ]
{ { }

With the use of (11), the above derivatives can be greatly simplified as the followings:

1 1 𝜇 𝜇 1 1
̃2 −
𝑈𝑥̃ = [Ω( 3 + 3 )] 𝑥̃ − ̃ 𝑦̃̇
( 3 − 3 ) = 𝑥̃̈ − 2Ω
1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠 2
(1 + 𝜇) 𝑟̃𝑝 𝑟̃𝑠
1 1 𝜇
̃2 −
⟹ 𝑈𝑦̃ = [Ω ( 3 + 3 )] 𝑦̃ ̃ 𝑥̃̇
= 𝑦̃̈ + 2Ω
1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠
1 1 𝜇
𝑈𝑧̃ = − ( 3 + 3 ) 𝑧̃ = 𝑧̃̈
{ 1 + 𝜇 𝑟̃𝑝 𝑟̃𝑠

Then, multiply these three equations by 2𝑥̃̇, 2𝑦̃̇, 2𝑧̃̇ respectively and sum them all up:

⟹ 2(𝑈𝑥̃ 𝑥̃̇ + 𝑈𝑦̃ 𝑦̃̇ + 𝑈𝑧̃ 𝑧̃̇ ) = 2𝑥̃̇𝑥̃̈ + 2𝑦̃̇𝑦̃̈ + 2𝑧̃̇ 𝑧̃̈

Yet, the above equation is nothing but the total derivative form with respect to 𝑡̃:

𝑑𝑈 𝑑 𝑑 2 𝑑
⟹ 2 = (𝑥̃̇ 2 + 𝑦̃̇ 2 + 𝑧̃̇ 2 ) = |𝐱̃̇| = |𝐯̃|2
𝑑𝑡̃ 𝑑𝑡̃ 𝑑𝑡̃ 𝑑𝑡̃

23
The last two equalities in above are straight forward from the definitions of variables.
Lastly, integrate both ends of the above equation over 𝑡̃, a simple elegant form is arrived:

⟹ 2𝑈 = |𝐯̃|2 + 𝐶

The integration constant 𝐶 here is Jacobi constant. Now, recall the original definition of
𝑈 in (23a), with the use of (2b) and direct expansion of |𝐯̃|2 , the above equation becomes:

1 1
−2 −2
̃ 2 (𝑥 2 2) 2 𝜇 2 2 2 1 2 2 2
⟹ Ω ̃ + 𝑦̃ + 1+𝜇
{[(𝑥̃ + 1+𝜇
) + 𝑦̃ + 𝑧̃ ] + 𝜇 [(𝑥̃ − 1+𝜇
) + 𝑦̃ + 𝑧̃ ] }

− (𝑣̃𝑥2 + 𝑣̃𝑦2 + 𝑣̃𝑧2 ) = 𝐶

For coplanar case where 𝑧̃ = 𝑣̃𝑧 = 0, the above becomes a function of four variables only:

̃) = 𝐶
𝑓(𝑥̃, 𝑦̃, 𝑣̃𝑥 , 𝑣̃𝑦 ; 𝜇, Ω

Since the plots employ only data points when test particle crosses positive 𝑥̃-axis at 𝑡̃ = 𝜏,
it follows that 𝑦̃𝜏 = 0 by construction, i.e., one variable is saved from the above function.
So, plotting under 𝑡̃ = 𝜏 reduces the number of variables in above function to three only:

̃) = 𝐶
𝑓(𝑥̃𝜏 , 𝑣̃𝑥,𝜏 , 𝑣̃𝑦,𝜏 ; 𝜇, Ω

Finally, one of the three variables can always be expressed in terms of the remaining two.
By expressing 𝑣̃𝑦,𝜏 in terms of (𝑥̃𝜏 , 𝑣̃𝑥,𝜏 ), the number of variables is further down to two:

Express 𝑣̃𝑦,𝜏 = 𝑔(𝑥̃𝜏 , 𝑣̃𝑥,𝜏 ) ⟹ 𝑓 (𝑥̃𝜏 , 𝑣̃𝑥,𝜏 , 𝑔(𝑥̃𝜏 , 𝑣̃𝑥,𝜏 )) = 𝐹(𝑥̃𝜏 , 𝑣̃𝑥,𝜏 ) = 𝐶,

where 𝑓, 𝑔, 𝐹 are functions

Thus, the scatter plot of 𝑣̃𝑥 against 𝑥̃ already contains full information of position
and velocity (two components each), and is termed Poincaré surface-of-section where each
𝐶 gives a surface. Finally, the adjustment of 𝑥̃-dimension position to its initial value as
𝑥̃𝜏 − 𝑥̃0 simply translates the plot laterally to around the origin. Notice that the same
adjustment is not needed for velocity as its initial value is already zero: 𝑣̃𝑥,0 = 0.

Before closing the discussion of plot, it remains to describe how data points are
retrieved as test particle crosses the positive 𝑥̃-axis. First, we need to know how it crosses:
from positive 𝑦̃ to negative or from negative to positive, i.e., the sign of 𝜔′ . From (7):


1 𝛿𝑎 −2 1 𝛿𝑎 𝜇 𝛿𝑎
𝜔 ≷0 ⟺ ( + ) ≷1 ⟺ 1≷ + ⟺ ≷
1+𝜇 𝑎 1+𝜇 𝑎 1+𝜇 𝑎
24
For 𝛿𝑎⁄𝑎 taking the format in (18), the above inequality relationship becomes:

𝜇 2 5
𝜔′ ≷ 0 ⟺ ≷ 𝑘𝜇 7 ⇐≈⇒ 𝜇 7 ≷ 𝑘, where 1 + 𝜇 ≈ 1
1+𝜇

5
From the range of parameters given in (20), obviously 𝜇 7 < 𝑘 and hence 𝜔′ < 0.
This means the test particle rotates in clockwise direction in the co-rotation frame. Thus in
coding, when 𝑦̃ changes sign from positive to negative, the weighted average of (𝑥̃, 𝑣̃𝑥 )
from these two points are recorded with the relative magnitude of 𝑦̃ being weights, where
the farther from 𝑥̃-axis or the larger the |𝑦̃|, the smaller is the weight and vice versa:

−𝑦̃𝑡̃ 𝑦̃𝑡̃−1
𝑥̃𝜏 = 𝑥̃𝑡̃−1 + 𝑥̃ ̃
𝑦̃𝑡̃−1 > 0 𝑦̃𝑡̃−1 +(−𝑦̃𝑡̃ ) 𝑦̃𝑡̃−1 +(−𝑦̃𝑡̃ ) 𝑡
If { , then { (note − 𝑦̃𝑡̃ > 0)
𝑦̃𝑡̃ < 0 𝑣̃𝑥,𝜏 =
−𝑦̃𝑡̃
𝑣̃𝑥,𝑡̃−1 +
𝑦̃𝑡̃−1
𝑣̃𝑥,𝑡̃
𝑦̃𝑡̃−1 +(−𝑦̃𝑡̃ ) 𝑦̃𝑡̃−1 +(−𝑦̃𝑡̃ )

The above condition is included in the main part (final section) of the code.

5.2 Chaotic vs Regular Cases

For the fourteen cases of 𝑘 and six cases of 𝜇 given in (20), the Poincaré surface-
of-section of these eighty-four cases are presented as fourteen rows by six columns of
figures on the following page. This mimics FIG.16 of Wisdom (1980) with two axes
interchanged, but the critical divide is now around horizontal panels of 𝛿𝑎⁄𝑎 = 1.3𝜇 2⁄7
instead of the dashed line of slope 7/2. Recall from Section 1 the third chaotic property is a
dense set of orbits. Thus, the chaotic case can be identified from the filled area of phase
space. For regular cases circle- or ellipse-like shapes should be observed; at least only lines
are shown without filled area. Some cases exhibit island loops, which are characteristic of
resonant motion. These are classified as special kinds of the regular patterns.

To facilitate analysis, the fourteen by six cases are summarised in a table afterwards,
where the letters C, I and R denote chaotic, resonant (island) and regular cases respectively.

It is clear from above table that the double thick line roughly splits the chaotic case
from regular, with the former above and latter below. This split is consistent with the
resonance overlap criterion 𝑘 = 1.307 suggested by Wisdom (1980). For his extensively
studied cases 𝜇 = 10−3 and 10−4 (the two rightmost columns), the split is in fact clearer.

25
𝑘\𝜇 10−8 10−7 10−6 10−5 10−4 10−3

1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

2.1

2.2

2.3
26
Summary of Cases: C = Chaotic, I = Resonant (Islands), R = Regular
𝑘 𝜇 10−8 10−7 10−6 10−5 10−4 10−3
1.0 C C C C C C
1.1 C C R R C C
1.2 R R C C C C
1.3 C C C C C C

1.4 R C R I R I
1.5 C R R R R R
1.6 R C R R C C
1.7 I R R R R R
1.8 C R R R R R
1.9 R R R R R C
2.0 R R R R R R
2.1 R R R R R R
2.2 R R R R R R
2.3 R R R R R R

Another split happens at 𝑘 between 1.9 and 2.0, where the below of this gives
straight regular cases. This corresponds to (18) where 𝑘 = 1.96, the theoretical case
implied directly from Wisdom (1980)’s linear scaling law. Yet one more split is located at
𝑘 between 1.0 and 1.1, where the above of this gives straight chaotic cases. It seems the
threshold is not a clear cut value at 𝑘 = 1.307 but a region spanned between 1.0 and 2.0.
Therefore, some subsequent papers simply quote the critical value as 𝜇 2⁄7 or 2𝜇 2⁄7 .

One interesting observation is the critical band is wider towards the left- or right-
hand side, i.e., when 𝜇 is small or large; the band is narrower when 𝜇 = 10−6 or 10−5 .

If 𝜇 is too small (towards 10−8 ), variables including both position and velocity of
test particle are also small (see the axis scale) so that numerical error is relatively large. To
see this, a small value of 𝜇 in (7) would make the first term in bracket very close to 1,
then |𝜔′ | would be exceptionally small so that it takes very long (i.e., many loops more
than other cases) to numerically integrate a given number of cycles (𝑁) according to (16).
As numerical error increases with all these, the critical band could be widened this way.

27
By contrast, if 𝜇 is too large, then as Wisdom (1980) mentions at the beginning
(Abstract) that the criterion would fail to hold true. Probably 𝜇 = 10−3 is the upper limit.

There are three interesting cases of resonant where islands (the loops) are observed.
These cases are rare because the position of the test particle has to satisfy a strict criterion.
By definition, resonance occurs when orbital periods of two close bodies (here the test
particle and secondary star) have a nontrivial least common multiple (i.e., other than direct
multiplication of the two). That is, the ratio between them constitutes a rational number:

𝑃𝜔 𝑝
= (24)
𝑃Ω 𝑝 + 𝑞
Here 𝑝 and 𝑞 are non-zero integers, where 𝑞 is the order of resonance which is negative
for external mean motion resonance (test particle being outside the system), and |𝑞| gives
the number of distinct visible islands on the Poincaré surface-of-section plot. As orbital
periods are normalised by uniformly adjusting the factor of √𝑎3 ⁄𝐺𝑀 according to (3),
̃=1
which is inversely proportional to normalised angular velocity by (6); also notice Ω
and with the use of (7) to eliminate 𝜔, now 𝑝 and 𝑞 are in terms of only parameters:

3
̃ 1
𝑃𝜔 𝑃̃𝜔 Ω 1 𝛿𝑎 2 𝑝
= = = =( + ) = (24a)
𝑃Ω 𝑃̃Ω 𝜔 𝜔 1+𝜇 𝑎 𝑝+𝑞
Recall the three cases with islands in above table; the companion figures are recapped here:

Three Cases with Distinct Islands: From Left to Right 𝒒 = −𝟕, −𝟐, −𝟑

To see their exact forms of resonance, just insert 𝜇 and 𝛿𝑎⁄𝑎 = 𝑘𝜇 2⁄7 into (24a)
then identify 𝑝 (𝑞 is already determined by the number of islands). In all our hypothetical
cases here however, the test particle is very close to the system such that 𝜔 is close to 1.
So one can imagine the implied value of 𝑝 would be exceptionally large for exceptionally
small value of 𝜇 –– as it decreases from 10−3 to 10−8 , 𝑝 increases from 13 to 536:

28
𝛿𝑎 2 1 536
𝜇 = 10−8 , = 1.7𝜇 7 , 𝑞 = −7 = 1.0132 ≈
𝑎 𝜔 536 + (−7)
𝑝 = 536
𝛿𝑎 2 1 27
−5
𝜇 = 10 , = 1.4𝜇 7 , 𝑞 = −2 ⟹ = 1.0793 ≈ ⟹ { 𝑝 = 27
𝑎 𝜔 27 + (−2) 𝑝 = 13
−3
𝛿𝑎 2 1 13
{𝜇 = 10 , 𝑎 = 1.4𝜇 , 𝑞 = −3 = 1.3039 ≈
7
{𝜔 13 + (−3)

Now back to the “regular” cases. Apart from the cases of islands just discussed,
resonance produces certain types of shapes other than the standard circular or elliptic ones.
Following Murray and Dermott (1999, Section 8.9), there are no cases of exact resonance,
medium-amplitude libration, nor apocentric libration. However, there are examples of the
remaining two cases: large-amplitude libration, inner circulation and outer circulation:

Cases from Left: Large-amplitude Libration, Inner Circulation, Outer Circulation

Finally, for chaotic cases sometimes resonance is found. Consider the followings:

Cases of More Chaotic (Upper) against More Resonant (Lower)

For these three pairs from left to right, the upper and lower panels of each column are of
neighbourhood cases with similar parameters. The upper panels are more chaotic while the
lower ones are less with resonant region(s), verifying again that the split is not clear cut.

29
Section 6. Elliptic Case

6.1 Notation and Setup

Previously the problem has been solved based on the circular orbit of test particle,
but the case of slightly elliptical orbit is interesting as the variation of eccentricity with
separation together has implications to the orbital behaviour. Specifically, the impact on
critical value of scaling law that separates chaotic from regular behaviour will be studied.

Consider the test particle rotating on a coplanar elliptical orbit in the sidereal frame
where the left focus is on centre of mass of the system while the right focus is on positive
𝑥̃-axis. The test particle elliptical orbit is completely inside the circular orbit of secondary
star but is completely outside that of primary star to avoid orbit crossing (otherwise objects
may crash with each other). As in previous section, the points where test particle gets close
to secondary star will be examined. It is customary to define the parameters as follows:

𝛼 = semi-major axis of the test particle elliptical orbit

𝛽 = semi-minor axis of the test particle elliptical orbit

𝐴 = area swept by the test particle elliptical orbit

𝑒 = eccentricity of the test particle elliptical orbit

𝑓 = true anomaly of test particle from positive 𝑥-axis in anticlockwise direction

ℎ = specific angular momentum of the test particle

𝑟 = distance of the test particle from origin

It is standard to derive elliptical orbit equation in spherical coordinates. Consider a


modified figure taken from Pelogia and Brasil (2017, Figure 1). Here 𝑎 (𝛼 in our context)
is semi-major axis with eccentricity 𝜀 (𝑒 in our context) where test particle is at (𝑥̃, 𝑦̃).

(𝑥̃, 𝑦̃)
𝜋 − 𝑓̃ 𝑓̃
𝑟̃ ′
𝑟̃

30
From the basic properties of ellipse, the followings are immediately arrived:

𝛽2 = 𝛼 2 (1 − 𝑒 2 )
{ ⟹ 𝐴 = 𝜋𝛼 2 √1 − 𝑒 2 (25)
𝐴 = 𝜋𝛼𝛽
Consider the right-angled triangle formed by hypotenuse 𝑟̃ ′, 𝑥-axis and dotted blue line.
By Pythagoras Theorem:

2 2 2 2 2 2
𝑟̃ ′ = [𝑟̃ sin(𝜋 − 𝑓̃)] + [2𝛼𝑒 + 𝑟̃ cos(𝜋 − 𝑓̃)] ⟹ 𝑟̃ ′ = (𝑟̃ sin 𝑓̃) + (2𝛼𝑒 − 𝑟̃ cos 𝑓̃)

∵ 𝑟̃ + 𝑟̃ ′ = 2𝛼, ∴ (2𝛼 − 𝑟̃ )2 = 𝑟̃ 2 + 4𝛼𝑒(𝛼𝑒 − 𝑟̃ cos 𝑓̃)

𝛼(1 − 𝑒 2 )
⟹ 𝑟̃ = (26)
1 − 𝑒 cos 𝑓̃
This is the orbit equation. The next task is to transform the positions and velocities from
spherical coordinates to our Cartesian coordinates context. For position this is simple:

(𝑥̃, 𝑦̃) = (𝑟̃ cos 𝑓̃ , 𝑟̃ sin 𝑓̃) (27)


But we also need to work with (𝑥̃̇, 𝑦̃̇) as well, so the time derivative terms including 𝑟̃̇ ,

𝑟̃ 𝑓̃̇ and 𝑓̃̇ will appear. Let’s tackle them one by one. Differentiate (26) with respect to 𝑡̃:

𝑟̃̇ − 𝑟̃̇ 𝑒 cos 𝑓̃ + 𝑟̃ 𝑓̃̇𝑒 sin 𝑓̃ = 0

𝑟̃ 𝑓̃̇𝑒 sin 𝑓̃
⟹ 𝑟̃̇ = − (28)
1 − 𝑒 cos 𝑓̃
To remove the term 𝑟̃ 𝑓̇ in (28), first recall the definition of specific angular momentum:

ℎ = 𝑟̃ 2 𝑓̃̇ (29)
Next, the Kepler’s second law states that equal area is being swept out in equal time.
Alternatively, this means the change of area over time is constant. With the use of (29):

𝑑𝐴 1 2 ̇ 1
= 𝑟̃ 𝑓̃ = ℎ
𝑑𝑡̃ 2 2

On the other hand, the area is equal to its change over time multiplied by orbital period:

𝑑𝐴 1 𝑃 1 √4𝜋 2 (𝛼𝑎)3 ⁄𝐺𝑀 3


𝐴≡ ⋅ 𝑃̃ ≡ ℎ ⋅ = ℎ⋅ = 𝜋ℎ𝛼 2
𝑑𝑡̃ 2 √𝑎3 ⁄𝐺𝑀 2 √𝑎3 ⁄𝐺𝑀

Note in above the pre-normalised semi-major axis is 𝛼𝑎. Eliminate 𝐴 in above with (25):

ℎ = √𝛼(1 − 𝑒 2 )

31
Next, eliminate ℎ with the use of (29), then eliminate 𝑟̃ with the use of (26):

√𝛼(1 − 𝑒 2 ) 1 − 𝑒 cos 𝑓̃
𝑟̃ 𝑓̃̇ = = (30)
𝑟̃ √𝛼(1 − 𝑒 2 )
Also, substitute (30) into (28) to eliminate 𝑟̃ 𝑓̇ :

𝑒 sin 𝑓̃
𝑟̃̇ = − (28a)
√𝛼(1 − 𝑒 2 )
Then use (26) to eliminate 𝑟̃ in (30) to obtain 𝑓̇ :

2
̇ 1 − 𝑒 cos 𝑓̃ (1 − 𝑒 cos 𝑓̃)
̃
𝑓= = (31)
2 3⁄2
𝑟̃ √𝛼(1 − 𝑒 2 ) [𝛼(1 − 𝑒 )]
Notice that this is simply the angular velocity, corresponds to 𝜔 in (7) of the circular case.

Now, 𝑟̃̇ , 𝑟̃ 𝑓̃̇ and 𝑓̃̇ are expressed solely in terms of basic constants and true anomaly 𝑓̃,
but the latter can be recovered to Cartesian coordinates easily from the basic conversion:

𝑥̃ 𝑥̃ 𝑦̃ 𝑦̃
cos 𝑓̃ = = , sin 𝑓̃ = = (27a)
𝑟̃ √𝑥̃ 2 + 𝑦̃ 2 𝑟̃ √𝑥̃ 2 + 𝑦̃ 2

The above equations about elliptical orbits are sufficient for the rest of analysis.

6.2 Other Conditions

Equations (1) to (12) of Sections 2.1 to 2.4 are all about gravitational forces in a
three-body synodic system, these are generic to all kinds of orbits so are applicable here.
Thus, the two more sets of conditions needed before completing the setup are initial
conditions corresponding to (13) and (14), as well as the no orbit crossing conditions.

The initial conditions needed are position and velocity. For position, from (27):

𝛼(1 − 𝑒 2 )
𝑥̃ = 𝑟̃ cos 𝑓̃ = cos 𝑓̃
1 − 𝑒 cos 𝑓̃
(27b)
𝛼(1 − 𝑒 2 )
̃
𝑦̃ = 𝑟̃ sin 𝑓 = sin 𝑓̃
{ 1 − 𝑒 cos 𝑓̃
For velocity, first of all the difference between angular velocity of test particle and
that of the system is needed, as is the circular case in (7). With the use of (31):

2
̇ (1 − 𝑒 cos 𝑓̃)
𝜔 ≡ 𝑓̃ − Ω
′ ̃= ̃
−Ω (32)
[𝛼(1 − 𝑒 2 )]3⁄2

32
̃ = 1 from (6) still applies. The velocity is given by the distance from the centre of
Note Ω
mass times angular velocity 𝑟̃ 𝜔′ . To decompose it into 𝑥̃- and 𝑦̃-components, the tangent
of elliptical orbit is needed. Differentiate the Cartesian orbit equation with respect to time:

𝑑 𝑥̃ 2 𝑦̃ 2 𝑑
( 2 + 2) = 1
𝑑𝑡̃ 𝛼 𝛽 𝑑𝑡̃

𝑑 𝑥̃ 2 𝑦̃ 2
By (25): ( 2+ 2 )=0
𝑑𝑡̃ 𝛼 𝛼 (1 − 𝑒 2 )

2 )𝑥
𝑑𝑥̃ 𝑑𝑦̃ 𝑣̃𝑦 (1 − 𝑒 2 )𝑥̃ (1 − 𝑒 2 ) cos 𝑓̃
⟹ 2(1 − 𝑒 ̃ + 2𝑦̃ =0 ⟹ = = , by (27)
𝑑𝑡̃ 𝑑𝑡̃ 𝑣̃𝑥 −𝑦̃ − sin 𝑓̃

With a little geometry, the 𝑥̃- and 𝑦̃-components of velocity are given by the followings:

− sin 𝑓̃
𝑣̃𝑥 = 𝑟̃ 𝜔′
√(1 − 𝑒 2 )2 cos2 𝑓̃ + sin2 𝑓̃
(1 − 𝑒 2 ) cos 𝑓̃
𝑣̃𝑦 = 𝑟̃ 𝜔′
√(1 − 𝑒 2 )2 cos 2 𝑓̃ + sin2 𝑓̃
{

With the use of (26) and (32):

2
− sin 𝑓̃ 𝛼(1 − 𝑒 2 ) (1 − 𝑒 cos 𝑓̃)
𝑣̃𝑥 = { ̃}
−Ω
1 − 𝑒 cos 𝑓̃ [𝛼(1 − 𝑒 2 )]3⁄2
√(1 − 𝑒 2 )2 cos 2 𝑓̃ + sin2 𝑓̃
⟹ 2 (33)
(1 − 𝑒 2 ) cos 𝑓̃ 𝛼(1 − 𝑒 2 ) (1 − 𝑒 cos 𝑓̃)
𝑣̃𝑦 = { ̃}
−Ω
1 − 𝑒 cos 𝑓̃ [𝛼(1 − 𝑒 2 )]3⁄2
√(1 − 𝑒 2 )2 cos 2 𝑓̃ + sin2 𝑓̃
{
The signs are chosen to match the anticlockwise direction. Without loss of generality, let
the test particle starts on positive 𝑥̃-axis at 𝑓̃0 = 0. With Ω
̃ = 1, (27b) and (33) become:

𝛼(1 − 𝑒 2 )
𝑥̃0 = cos 0 = 𝛼(1 + 𝑒)
̃
For 𝑓0 = 0: 1 − 𝑒 cos 0 (27c)
𝛼(1 − 𝑒 2 )
{ 𝑦̃0 = 1 − 𝑒 cos 0 sin 0 = 0
− sin 0 𝛼(1 − 𝑒 2 ) (1 − 𝑒 cos 0)2
𝑣̃𝑥,0 = { ⁄
− 1}
√(1 − 𝑒 2 )2 cos2 0 + sin2 0 1 − 𝑒 cos 0 [𝛼(1 − 𝑒 2 )]3 2
For 𝑓̃0 = 0:
(1 − 𝑒 2 ) cos 0 𝛼(1 − 𝑒 2 ) (1 − 𝑒 cos 0)2
𝑣̃𝑦,0 = { ⁄
− 1}
{ √(1 − 𝑒 2 )2 cos 2 0 + sin2 0 1 − 𝑒 cos 0 [𝛼(1 − 𝑒 2 )]3 2

33
𝑣̃𝑥,0 = 0
⟹ { 1−𝑒 (33a)
𝑣̃𝑦,0 = √ − 𝛼(1 + 𝑒)
𝛼(1 + 𝑒)
1
If 𝑒 = 0, (33a) boils down to the circular case similar to (7) with 𝛼 = 1+𝜇 + 𝛿𝑎
𝑎
. If 𝑒 = 1,

the orbit becomes a parabola with 𝑣̃𝑦,0 = −2𝛼. (27c) and (33a) are the only modifications
to previous formulation to generate an elliptical orbit, where other parts being unchanged.

Recall in above that the test particle orbit is constructed to be in-between those of
primary and secondary stars. It follows that the semi-major axis of test particle elliptical
orbit should be strictly shorter than that (or radius) of the secondary star, while the semi-
minor axis of elliptical orbit should be strictly longer than the semi-major axis (or radius)
of the primary star. With (2a) and (25), the “no orbit crossing condition” can be stated as:

1 1
𝛼< 𝛼 <
1+𝜇 1+𝜇
{ 𝜇 ⟹ { 𝜇
𝛽> 𝛼 √1 − 𝑒 2 >
1+𝜇 1+𝜇

𝜇 1
⟹ <𝛼<
(1 + 𝜇)√1 − 𝑒 2 1+𝜇

As the critical separation between chaotic and regular orbits happens when the test
particle is close to secondary star, this means 𝛼 is much closer to the upper limit than the
lower one. As long as 𝑒 is small, the first inequality is very likely to be satisfied. This can
be seen by rearranging terms around the first inequality to express 𝑒 on one side:

𝜇 2
𝑒 < √1 − [ ]
𝛼(1 + 𝜇)

With the previous critical value of 𝛼 = 1.3𝜇 2⁄7 and a large value of 𝜇 = 10−3 to lower
the cap, we still get 𝑒 < 0.9999847. Hence the above condition is almost surely satisfied.

6.3 Simulation

Since the dynamics in (12) are unchanged, the procedure of numerical simulation is
also the same, thus Section 3.1 applies here. For coding, the functions “F” and “RK4” in
Appendix A also apply; for main part, replace the initial conditions with (27c) and (33a).
Same format of data files are generated and similar Poincaré surface-of-sections are plotted.
34
The parameters used in (16) to (18) of Section 3.3 will be retained here. But as the
test particle is now inside the orbit of secondary star, its initial distance from origin is
subtracted from that of the secondary star rather than added to. From (2a), the secondary
star is located at 1⁄(1 + 𝜇); and from the previous ellipse chart, the test particle is initially
located at 𝛼(1 + 𝑒). Then the initial separation between the two is simply:

1 2
𝛿𝛼 = − 𝛼(1 + 𝑒) = 𝑘𝜇 7
1+𝜇

Given the parameters 𝜇, 𝑒 and 𝑘, the value of 𝛼 can be passively determined:

1 1 2
𝛼= ( − 𝑘𝜇 7 )
1+𝑒 1+𝜇

And the values of these three parameters used in simulation are given in the followings,
where the ranges of 𝜇 and 𝑒 follow those from Wisdom (1980, FIG. 8 – FIG.10):

𝜇 = 10−3 , 10−4 , 10−5 , 𝑒 = 0.4, 0.8, 1.2, 1.6, 2.0, 𝑘 = 1.0, 1.1, … , 2.1 (34)
With these ranges, the angular velocity in co-rotation frame from (32) is negative;
indeed, calculation shows that −0.3 < 𝜔′ < −0.1. This means the test particle is again
rotating clockwise as before. Thus, the coding operation in retrieving values as it crosses
𝑥̃-axis mentioned in the end of Section 5.1 (also in the end of the main part of the code)
again applies. Before closing, as test particle is rotating on elliptic orbit slower than the
two stars on circular orbits, a rotating ellipse over time is observed under co-rotation frame.

As the above figures illustrate, for small 𝑒 where the elliptic orbit is near circular,
after some time the rotation sweeps out a thin donut-like shape. As eccentricity increases,
say, to 0.16 which is still a small value, a pretty thick donut shape is obtained. This is why
the eccentricity in simulation cannot take large values. Literature takes a max 𝑒 of 0.2.

In the following pages, the Poincaré surface-of-section for 𝜇 = 10−3 , 10−4 , 10−5
are shown, one set on each page, with 𝑘 varied across rows and 𝑒 varied across columns.

35
6.4 Results 𝝁 = 𝟏𝟎−𝟑
𝒌\𝒆 0.04 0.08 0.12 0.16 0.20

1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

2.1

36
𝝁 = 𝟏𝟎−𝟒
𝒌\𝒆 0.04 0.08 0.12 0.16 0.20

1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

2.1

37
𝝁 = 𝟏𝟎−𝟓
𝒌\𝒆 0.04 0.08 0.12 0.16 0.20

1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

2.1

38
Summary of Cases: C = Chaotic, I = Resonant (Islands), R = Regular
𝜇 10−3 10−4 10−5
e
k 0.04 0.08 0.12 0.16 0.20 0.04 0.08 0.12 0.16 0.20 0.04 0.08 0.12 0.16 0.20
1.0 C C C C C I C C C I C C C C C
1.1 I C I C I C C C C C C C C C C
1.2 R C C C C I C C C I C C C I C
1.3 C R I R C I I C C I R C C C C
1.4 R I C C C I R C R C C C C C C
1.5 R R R R R R R C R C R C C C C
1.6 R R R R R I R C I C C C C I C
1.7 R R R R R R I C R C C C C I C
1.8 R R R R R R R I C R C C C I C
1.9 I R I R R R R R I R R R I I C
2.0 R R R R R R R R I R R C R I C
2.1 I R R R R R R R R I R R R R R

6.5 Discussion

The chaotic versus regular behaviours are basically the same as before, here won’t
repeat the discussion in Section 5.2. However, the patterns across both eccentricity and
initial separation (between test particle and secondary star) are clear from the above table,
which summarises the 165 plots in previous three pages. Here are the patterns observed:

1. For each mass ratio 𝜇, the critical value that separates chaotic and regular
behaviour generally increases as eccentricity 𝑒 increases;
2. For each eccentricity 𝑒, the critical value increases as mass ratio 𝜇 decreases;
3. The critical values are between 𝑘 = 1 and 2, which agrees with Section 5.2;
4. Resonant cases are generally found around the critical value.

Finding “3” indeed reconfirms the scaling law on a rough scale. Quillen and Faber
(2006, Abstract) “find that the distance in semimajor axis between the planet and boundary
depends on the planet mass to the 2/7 power and is independent of the planet eccentricity,
at least for planet eccentricities below 0.3”, and Giuppone et al (2013, Abstract) “show that
this criterion fits the stability regions in real exoplanet systems quite well.”

39
In spite of all these, Mustill and Wyatt (2012, Abstract) “shows that, above the
critical eccentricity, the chaotic zone width is given by (𝛿𝑎⁄𝑎)chaos ≈ 1.8𝑒 1⁄5 𝜇 1⁄5 , ……
[t]he critical eccentricity is given by 𝑒𝑐𝑟𝑖𝑡 = 0.21𝜇 3⁄7 .” This 1/5-th law is labelled as the
“new scaling law” in the literature. There seems no further follow-up about this to date of
this writing. To verify this, first obtain the critical eccentricity from our studied cases:

10−3 0.21 × (10−3 )3⁄7 = 0.0109


𝜇 = {10−4 ⟹ 𝑒𝑐𝑟𝑖𝑡 = {0.21 × (10−4 )3⁄7 = 0.0041
10−5 0.21 × (10−5 )3⁄7 = 0.0015

Obviously all the eccentricity values used here are larger than the above critical
values. If the 1/5-th law applies, for our fifteen cases (three values of 𝜇, five values of 𝑒)
we can compute the (𝛿𝑎⁄𝑎)chaos mentioned above and then deduce the theoretical critical
M&W
values of 𝑘 (𝑘crit ) under the old Wisdom scaling law. These are in turn compared with
Law
the simulated critical values of 𝑘 obtained here (𝑘crit , the critical divide in above table),
where widest possible ranges are reported with resonant cases (I) regarded as regular (R):

Critical Value of 𝒌 Implied by Mustill and Wyatt (2012) 1/5-th Law


𝜇 10−3 10−4 10−5
𝑒 0.04 0.08 0.12 0.16 0.20 0.04 0.08 0.12 0.16 0.20 0.04 0.08 0.12 0.16 0.20
𝛿𝑎⁄𝑎 0.24 0.27 0.30 0.31 0.33 0.15 0.17 0.19 0.20 0.21 0.09 0.11 0.12 0.12 0.13
M&W 1.71 1.96 2.13 2.26 2.36 2.08 2.39 2.59 2.75 2.87 2.54 2.91 3.16 3.35 3.50
𝑘crit
Law
𝑘crit 1.0-1.4 1.2-1.3 1.0-1.5 1.2-1.5 1.0-1.5 <1.2 1.2-1.3 1.7-1.8 1.3-1.9 <1.8 1.2-1.9 1.8-2.1 1.8-1.9 1.5-1.6 2.0-2.1

By comparing the last two rows of above table, the theoretical value of critical 𝑘
based on the 1/5-th law differ quite a lot from the simulated results here; in fact the latter
are closer to the flat value of 𝑘 = 1.3 under Wisdom 2/7 law. Among the four findings
mentioned at the beginning of this section, “1” is consistent with 1/5-th law but “2” is not.
We find the critical separation decreases with mass ratio instead of one being proportional
to the other. From the summary table in Section 5.2, the case of circular orbit, the critical
separation decreases with mass ratio for 𝜇 ≤ 10−6 but increases for 𝜇 ≥ 10−5 . Hence
there may not be a simple log-linear relationship despite the old scaling law works roughly.

Future research could be along the line of missing variable(s) that results in this
non-linear pattern; one potential candidate is time. As Nesvold and Kuchner (2014) show,
under the collision context the critical separation is a power law as a function of time. Thus,
a natural trial in next step would be something in the form of (𝛿𝑎⁄𝑎)chaos = 𝛾0 𝑡 𝛾1 𝑒 𝛾2 𝜇 𝛾3 .
40
References

Deck, K. M.; Payne, M.; Holman, M. J. 2013. First-order Resonance Overlap and the
Stability of Close Two-planet Systems. Astrophysical Journal, 774 (2), 129 – 141.

Duncan, M., Quinn, T.; Tremaine, S. 1989. The Long-term Evolution of Orbits in the Solar
System: A Mapping Approach. Icarus, 82 (2), 402 – 418.

Giuppone, C. A.; Morais, M. H. M.; Correia, A. C. M. 2013. A Semi-empirical Stability


Criterion for Real Planetary Systems with Eccentric Orbits. Monthly Notices of the
Royal Astronomical Society, 436 (4), 3547 – 3556.

Hou, X. Y.; Scheeres, D. J.; Liu, L. 2014. Saturn Trojans: A Dynamical Point of View.
Monthly Notices of the Royal Astronomical Society, 437 (2), 1420 – 1433.

Malhotra, R. 1998. Orbital Resonances and Chaos in the Solar System. In ASP Conference
Series 149, Solar System Formation and Evolution, ed. D. Lazzaro, R. Vieira
Martins, S. Ferraz-Mello, J. Fernandez, and C. Beaugé (San Francisco), 37.

Murray, C. D., Dermott, S. F. 1999. Solar System Dynamics. Cambridge University Press.

Musa, H.; Saidu, Ibrahim; Waziri, M. Y. 2010. A Simplified Derivation and Analysis of
Fourth Order Runge Kutta Method. International Journal of Computer
Applications (0975 – 8887), 9 (8), 51 – 55, November.

Mustill, Alexander J.; Wyatt,Mark C. 2012. Dependence of a Planet’s Chaotic Zone on


Particle Eccentricity: The Shape of Debris Disc Inner Edges. Monthly Notices of
the Royal Astronomical Society, 419 (4), 3074 – 3080.

Nesvold, Erika R.; Kuchner, Marc J. 2014. Gap Clearing by Planets in a Collisional Debris
Disk. The Astrophysical Journal, 798 (2), 83 – 92.

Pelogia, Karla; Brasil, Carlos Alexandre 2017. “The universal meaning of the quantum of
action”, by Jun Ishiwara. arXiv:1708.03706.

Press, William H.; Teukolsky, Saul A.; Vetterling, William T.; Flannery, Brian P. 2007.
Numerical Recipes: The Art of Scientific Computing, Third Edition. Cambridge
University Press.

41
Quillen, Alice C.; Faber, Peter 2006. Chaotic Zone Boundary for Low Free Eccentricity
Particles Near an Eccentric Planet. Monthly Notices of the Royal Astronomical
Society, 373 (3), 1245 – 1250.

Slíz-Balogh, Judit; Barta, András; Horváth, Gábor 2019. Celestial Mechanics and
Polarization Optics of the Kordylewski Dust Cloud in the Earth-Moon Lagrange
Point L5 - Part II. Imaging Polarimetric Observation: New Evidence for the
Existence of Kordylewski Dust Cloud. Monthly Notices of the Royal Astronomical
Society, 482 (1), 762 – 770.

Taylor, D. B. 1981. Horseshoe Periodic Orbits in the Restricted Problem of Three Bodies
for a Sun-Jupiter Mass Ratio. Astronomy & Astrophysics, 103 (2), 288 – 294.

Tiscareno, Matthew S.; Thomas, Peter C.; Burns, Joseph A. 2009. The Rotation of Janus
and Epimetheus. Icarus, 204 (1), 254 – 261.

Wisdom, J. 1980. The Resonance Overlap Criterion and the Onset of Stochastic Behavior
in the Restricted Three-Body Problem. Astronomical Journal, 85 (8), 1122 – 1133.

Word count before Reference Section:

Front cover to page vi + Pages 1 to 40

= 555 + 9,445 = 10,000

42
Appendix A. C++ Code for Poincaré Surface-of-Section
// Astrophysics Project 7011ASTPHY, by KC Law, Ka Chung PN817379, dated July 2019
// This is the C++ code running the circular restricted three-body problem (CR3BP)
// Parameters should be set in the main() section below, after functions F and RK4
// Upon running no. of rotations (N) is shown, finally data file will be generated

//#include "pch.h" // include for Visual Studio 2017, block for Visual Studio 2019
#include <iostream> // i/o stream, for input from keyboard and output to screen
#include <fstream> // file stream, allow write to and retrieve from disk file
#include <iomanip> // input/output parametric manipulators
#include <cassert> // assert if file is opened or closed

// Differential vector equations F = dX/dt


double* F(int neq, double h, double t, double mu, double* x)
{
double rp3 = pow(pow(x[0] + mu / (1 + mu), 2) + pow(x[1], 2), 1.5); // r{p}^3
double rs3 = pow(pow(x[0] - 1 / (1 + mu), 2) + pow(x[1], 2), 1.5); // r{s}^3
double* dxdt = new double[neq];
dxdt[0] = x[2];
dxdt[1] = x[3];
dxdt[2] = 2 * x[3] + (1 - (1 / rp3 + mu / rs3) / (1 + mu)) * x[0] - (1 / rp3 - 1 / rs3) * mu / pow(1 + mu, 2);
dxdt[3] = -2 * x[2] + (1 - (1 / rp3 + mu / rs3) / (1 + mu)) * x[1];
return dxdt;
}

// 4th order Runge-Kutta numerical integration


double* RK4(int neq, double h, double t, double mu, double* x)
{
double* k1 = new double[neq];
double* k2 = new double[neq];
double* k3 = new double[neq];
double* k4 = new double[neq];
double* x2 = new double[neq];
double* x3 = new double[neq];
double* x4 = new double[neq];
double* F1 = F(neq, h, t, mu, x);
for (int i = 0; i < neq; ++i)
{
k1[i] = h * F1[i];
x2[i] = x[i] + k1[i] / 2;
}
double* F2 = F(neq, h, t, mu, x2);
for (int i = 0; i < neq; ++i)
{
k2[i] = h * F2[i];
x3[i] = x[i] + k2[i] / 2;
}
double* F3 = F(neq, h, t, mu, x3);
for (int i = 0; i < neq; ++i)
{
k3[i] = h * F3[i];
x4[i] = x[i] + k3[i];
}
double* F4 = F(neq, h, t, mu, x4);
for (int i = 0; i < neq; ++i)
{
k4[i] = h * F4[i];
x[i] = x[i] + k1[i] / 6 + k2[i] / 3 + k3[i] / 3 + k4[i] / 6;
}
delete[] F1;
delete[] F2;
delete[] F3;
delete[] F4;
delete[] k1;
delete[] k2;
delete[] k3;
delete[] k4;
delete[] x2;
delete[] x3;
delete[] x4;
return x;
}

// Main part of the programme, set parameters here


int main()
{
char filename[100] = "D:/Astrophysics/Project/Out-4 1.3.dat"; // output file path, name and extension
int neq = 4; // number of equations in differential vector equations F = dX/dt
int N = 1000; // number of rotations by test particle
double mu = pow(10, -4); // mu, mass ratio of secondary star to primary star
double da = 1.3 * pow(mu, (double)2 / (double)7); // delta{a}/a, secondary-test separation to secondary-primary separation
double h = 0.01; // timestep under 4th order Runge-Kutta numerical integration method
double omegap = pow(1 / (1 + mu) + da, -1.5) - 1; // omega', test particle angular velocity under corotation frame
double* x = new double[neq];
x[0] = 1 / (1 + mu) + da; // x{ t}, test particle x-dimension position under corotation frame
x[1] = 0; // y{ t}, test particle y-dimension position under corotation frame
x[2] = 0; // v{x,t}, test particle x-dimension velocity under corotation frame
x[3] = (1 / (1 + mu) + da) * omegap; // v{y,t}, test particle y-dimension velocity under corotation frame
double x0 = x[0]; // x{ 0}, record initial position x{t} at t = 0 for later use
double xt = x[0]; // initialise to record x{ t} just before crossing positive x-axis
double yt = x[1]; // initialise to record y{ t} just before crossing positive x-axis
double vxt = x[2]; // initialise to record v{x,t} just before crossing positive x-axis
const double pi = acos(-1); // define pi = 3.14159
double tmax = 2 * pi * N / abs(omegap); // t{max}, final time
double t = 0; // t{0} , initial time
std::ofstream output(filename); // prepare to write to file
assert(output.is_open()); // make sure file is opened
while (t < tmax)
{
std::cout << std::setprecision(3) << N * t / tmax << "\n"; // print nth rotations on screen, optional
if (yt > 0 && x[1] < 0) // y{t-1} > 0 & y{t} < 0, crossing x-axis
{
double x_mean = (xt * x[1] - x[0] * yt) / (x[1] - yt) - x0; // x{ mean} = mean of x{t-1} and x{t}, weighted by |y|
double vx_mean = (vxt * x[1] - x[2] * yt) / (x[1] - yt); // v{x,mean} = mean of vx{t-1} and vx{t}, weighted by |y|
output << std::setprecision(15) << x_mean << " " << vx_mean << " " << x[0] << " " << x[1] << "\n"; // write to file
}
xt = x[0]; // record x{ t} for next x-axis crossing
yt = x[1]; // record y{ t} for next x-axis crossing
vxt = x[2]; // record v{x,t} for next x-axis crossing
x = RK4(neq, h, t, mu, x); // update x by running RK4
t = t + h; // update timestep by adding h
}
output.close();
return 0;
}

43
Appendix B. GNUPLOT Code for 𝑼∗ (𝒙 ̃)
̃, 𝒚
set title "U*(x,y)"
set grid
set xlabel "x"
set ylabel "y"
set zlabel "U*(x,y)"
set xrange [-1.5:1.5]
set yrange [-1.5:1.5]
set zrange [-3.5:-1.5]
splot -0.8*((x+0.2)**2+y**2)**-0.5-0.2*((x-0.8)**2+y**2)**-0.5-0.5*(x**2+y**2)

Appendix C. C++ Code for Horseshoe Orbit


The first two parts “F” and “RK4” are same as Appendix A. Here gives only the main part.
// Main part of the programme, set parameters here
int main()
{
char filename[100] = "D:/Astrophysics/Project/Horseshoe.dat"; // output file path, name and extension
int neq = 4; // number of equations in differential vector equations F = dX/dt
double mu = 9.53875 * pow(10, -4); // mu, mass ratio of secondary star to primary star
double h = 0.01; // timestep under 4th order Runge-Kutta numerical integration method
double* x = new double[neq];
x[0] = - 1.06201; // x{ t}, test particle x-dimension position under corotation frame
x[1] = 0; // y{ t}, test particle y-dimension position under corotation frame
x[2] = 0; // v{x,t}, test particle x-dimension velocity under corotation frame
x[3] = 0.10851; // v{y,t}, test particle y-dimension velocity under corotation frame
const double pi = acos(-1); // define pi = 3.14159
double tmax = 200; // t{max}, final time
double t = 0; // t{0} , initial time
std::ofstream output(filename); // prepare to write to file
assert(output.is_open()); // make sure file is opened
while (t < tmax)
{
std::cout << std::setprecision(3) << t << "\n"; // print nth rotations on screen, optional
output << std::setprecision(15) << x[0] << " " << x[1] << "\n"; // write x{mean} and v{x,mean}
x = RK4(neq, h, t, mu, x); // update x by running RK4
t = t + h; // update timestep by adding h
}
output.close();
return 0;
}

44

You might also like