You are on page 1of 112

Relative Orbital Motion Dynamical Models

for Orbits about Nonspherical Bodies

Item Type text; Electronic Thesis

Authors Burnett, Ethan Ryan

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this material


is made possible by the University Libraries, University of Arizona.
Further transmission, reproduction or presentation (such as
public display or performance) of protected items is prohibited
except with permission of the author.

Download date 29/12/2019 13:21:56

Link to Item http://hdl.handle.net/10150/628098


Relative Orbital Motion Dynamical Models for
Orbits about Nonspherical Bodies

by
Ethan R. Burnett

Copyright Ethan
c R. Burnett 2018

A Thesis Submitted to the Faculty of the


Department of Aerospace and Mechanical Engineering
In Partial Fulfillment of the Requirements
For the Degree of
Master of Science
With a Major in Aerospace Engineering
In the Graduate College
The University of Arizona

2018
2
3

Acknowledgments

I would like to thank my advisor, Dr. Eric Butcher, for his encouragement and
excellent teaching.

I greatly enjoyed working with my research colleagues: Mohammad Maadani, Bharani


Malladi, David Yaylali, Jingwei Wang, Morad Nazari, and Arman Dabiri.

I thank my parents and my brothers for their steadfast support.


4

Vita

Education

Master of Science, Aerospace Engineering


Emphasis: Dynamics and Control
University of Arizona, Tucson, AZ May 2018

Bachelor of Science, Aerospace Engineering


Minor, Physics
University of Arizona, Tucson, AZ May 2016

Publications

Ethan Burnett and Eric Butcher, “Linearized Relative Orbital Motion Dynamics in a Ro-
tating Second Degree and Order Gravity Field,” AAS 18-232, AAS/AIAA Astrodynamics
Specialist Conference, Snowbird, UT, August 19 – 23, 2018 (submitted)

Ethan Burnett, Eric Butcher, and T. Alan Lovell, “Linearized Relative Orbital Motion
Model About an Oblate Body Without Averaging,” AAS 18-218, AAS/AIAA Astrodynam-
ics Specialist Conference, Snowbird, UT, August 19 – 23, 2018 (submitted)

Ethan Burnett and Andrew J. Sinclair, “Interpolation on the Unit Sphere in Laplace’s
Method,” AAS 17-793, AIAA/AAS Astrodynamics Specialist Conference, Columbia River
Gorge, Stevenson, WA, August 20 – 24, 2017

Kristofer Drozd, Ethan Burnett, Eric Sahr, Drew McNeely, Vittorio Franzese, and Natividad
Ramos Morón, “Block-Like Explorer of a Near-Earth Body by achieving Orbital Proximity
(BEEBOP),” AAS 17-846, AIAA/AAS Astrodynamics Specialist Conference, Columbia
River Gorge, Stevenson, WA, August 20 – 24, 2017

Eric Butcher, Ethan Burnett, Jingwei Wang, and T. Alan Lovell, “A New Time-Explicit
J2-Perturbed Nonlinear Relative Orbit Model with Perturbation Solutions,” AAS 17-758,
AIAA/AAS Astrodynamics Specialist Conference, Columbia River Gorge, Stevenson, WA,
August 20 – 24, 2017

Eric Butcher, Ethan Burnett, and T. Alan Lovell, “Comparison of Relative Orbital Motion
Perturbation Solutions in Cartesian and Spherical Coordinates,” AAS 17-202, AIAA/AAS
Spaceflight Mechanics Meeting, San Antonio, TX, February 5 – 9, 2017
5

Table of Contents

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2. Orbital Motion and Relative Motion . . . . . . . . . . . . . . . . 15


2.1. Classical Theory of Satellite Orbits . . . . . . . . . . . . . . . . . . . 15
2.1.1. Gravitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.2. The Two-Body Problem and Elliptic Orbits . . . . . . . . . . 16
2.1.3. Orbital Elements for Elliptic and Circular Orbits . . . . . . . 18
2.2. Analytical Relative Dynamics of Co-orbiting Bodies . . . . . . . . . . 19
2.2.1. Relative Orbital Motion . . . . . . . . . . . . . . . . . . . . . 20
2.2.2. The Hill-Clohessy-Wiltshire Model . . . . . . . . . . . . . . . 21
2.2.3. Representing Model Error . . . . . . . . . . . . . . . . . . . . 26
2.3. Orbital Perturbations . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3.1. Perturbed Orbital Motion . . . . . . . . . . . . . . . . . . . . 28
2.3.2. Types of Perturbations . . . . . . . . . . . . . . . . . . . . . . 29
2.3.3. Perturbed Relative Orbital Motion . . . . . . . . . . . . . . . 30

3. Nonspherical Body Gravitation and Orbital Mechanics . . . . 34


3.1. Gravitational Potential Expressed by Spherical Harmonics . . . . . . 34
3.2. The Shape of the Earth . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3. Orbital Motion under the Influence of J2 . . . . . . . . . . . . . . . . 37
3.4. Moons, Planetoids, and Asteroids . . . . . . . . . . . . . . . . . . . . 38
6

Table of Contents—Continued

3.5. Orbital Motion about Small Asteroids . . . . . . . . . . . . . . . . . 39

4. A J2 -Perturbed Relative Orbit Model . . . . . . . . . . . . . . . 43


4.1. Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.1.1. Case of Vanishing Chief Eccentricity . . . . . . . . . . . . . . 47
4.2. Model Analysis and Validation . . . . . . . . . . . . . . . . . . . . . . 50

5. A Linearized Relative Orbit Model for Orbits in a Rotating


Second Degree and Order Gravity Field . . . . . . . . . . . . . . 55
5.1. Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.1.1. Case of Vanishing Chief Eccentricity . . . . . . . . . . . . . . 70
5.1.2. Extension to Eccentric Chief Orbits . . . . . . . . . . . . . . . 74
5.2. Model Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.2.1. The Case of Equatorial Orbits . . . . . . . . . . . . . . . . . . 77
5.2.2. A Special Case LTI Model . . . . . . . . . . . . . . . . . . . . 78
5.2.3. Analysis of Unstable and Stable Eigenspaces . . . . . . . . . . 82
5.3. Model Validation via Simulation . . . . . . . . . . . . . . . . . . . . . 87
5.3.1. Hypothetical Asteroid for Case Studies . . . . . . . . . . . . . 88
5.3.2. Variable Chief Inclination with Constant Γ . . . . . . . . . . . 88
5.3.3. Variable Γ with Constant Chief Inclination . . . . . . . . . . . 95
5.3.4. Variable Chief Semimajor Axis with Constant Body Rotation
Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.3.5. Interesting Cases . . . . . . . . . . . . . . . . . . . . . . . . . 101

6. Conclusions and Suggestions for Future Work . . . . . . . . . 107

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
7

List of Figures

Figure 2.1. Elliptical Orbit Geometry22 . . . . . . . . . . . . . . . . . . . . 18


Figure 2.2. The 3-1-3 Euler Angles and the Orbit Plane23 . . . . . . . . . . 19
Figure 2.3. Relative Motion Problem Geometry . . . . . . . . . . . . . . . 20
Figure 2.4. Relative Distance for Unperturbed and Perturbed Cases . . . . 32
Figure 2.5. ∆h(t) for Unperturbed and Perturbed Cases . . . . . . . . . . . 32

Figure 3.1. Primary Body Mass Distribution Geometry . . . . . . . . . . . 34


Figure 3.2. Nodal Regression due to J2 21 . . . . . . . . . . . . . . . . . . . 38
Figure 3.3. Asteroids and Comets Visited by Spacecraft24 . . . . . . . . . . 39
Figure 3.4. Spacecraft Orbital Ejection due to SRP Forces . . . . . . . . . 42
Figure 3.5. Spacecraft Orbital Velocity Before and During Ejection . . . . . 42

Figure 4.1. Relative Orbit for 4 Orbits (LEO with J2 ) . . . . . . . . . . . . 51


Figure 4.2. Small Parameter Values for Two Orbits (LEO with J2 ) . . . . . 52
Figure 4.3. Radial Error (LEO) . . . . . . . . . . . . . . . . . . . . . . . . 53
Figure 4.4. Along-Track Error (LEO) . . . . . . . . . . . . . . . . . . . . . 53
Figure 4.5. Cross-Track Error (LEO) . . . . . . . . . . . . . . . . . . . . . 53
Figure 4.6. Error Norm (LEO) . . . . . . . . . . . . . . . . . . . . . . . . . 53
Figure 4.7. Maximum Error Norm over Two Orbits vs. Chief Orbit Radius 54

Figure 5.1. Problem Geometry for Orbiting a Rotating Ellipsoidal Body . . 58


Figure 5.2. Libration Points for Rotating Ellipsoidal Body . . . . . . . . . . 79
Figure 5.3. Superposition of In-Plane Modes at Stable Libration Point . . . 84
Figure 5.4. Simulated In-Plane Motion at Stable Libration Point . . . . . . 84
Figure 5.5. Stable and Unstable Manifolds (Unstable Libration Point) . . . 87
Figure 5.6. Average Modeling Error vs. Inclination, Γ = 2.4 . . . . . . . . . 90
8

List of Figures—Continued

Figure 5.7. Maximum Deviation of Relative Orbit Angular Momentum, Γ =


2.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
Figure 5.8. Average Modeling Error vs. Inclination, Γ = 1.2 . . . . . . . . . 93
Figure 5.9. Maximum Deviation of Relative Orbit Angular Momentum, Γ =
1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Figure 5.10. Average Modeling Error vs. Inclination, Γ = 0.8 . . . . . . . . . 94
Figure 5.11. Maximum Deviation of Relative Orbit Angular Momentum, Γ =
0.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Figure 5.12. Average Modeling Error vs. Γ, i = 10◦ . . . . . . . . . . . . . . 96
Figure 5.13. Maximum Deviation of Relative Orbit Angular Momentum, i = 10◦ 97
Figure 5.14. Average Modeling Error vs. Γ, i = 45◦ . . . . . . . . . . . . . . 97
Figure 5.15. Maximum Deviation of Relative Orbit Angular Momentum, i = 45◦ 98
Figure 5.16. Average Modeling Error vs. Γ, i = 75◦ . . . . . . . . . . . . . . 98
Figure 5.17. Maximum Deviation of Relative Orbit Angular Momentum, i = 75◦ 99
Figure 5.18. Average Modeling Error vs. Semimajor Axis, i = 75◦ . . . . . . 100
Figure 5.19. Max Deviation of Relative Angular Momentum vs. Semimajor
Axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
Figure 5.20. Relative Distance with Relative Orbit Resonance . . . . . . . . 103
Figure 5.21. ∆h(t) with Relative Orbit Resonance . . . . . . . . . . . . . . . 103
Figure 5.22. Model Error (Relative Orbit Resonance Case) . . . . . . . . . . 104
Figure 5.23. Relative Orbit (Resonance Case) . . . . . . . . . . . . . . . . . 104
Figure 5.24. Relative Orbit (Marginal Stability Case) . . . . . . . . . . . . . 106
Figure 5.25. Model Error (Marginal Stability Case) . . . . . . . . . . . . . . 106
9

List of Tables

Table 2.1. Orbital Elements and Differences . . . . . . . . . . . . . . . . . . 31

Table 3.1. Asteroid Physical Parameters . . . . . . . . . . . . . . . . . . . . 40

Table 4.1. Orbital Elements for LEO Relative Orbit Scenario with J2 . . . 51

Table 5.1. Unique O Matrix Terms in 5.31, Oij = Oji . . . . . . . . . . . . 62


Table 5.2. Unique E Matrix Terms in 5.32, Eij = Eji . . . . . . . . . . . . 62
Table 5.3. Unstable and Stable Eigenvalues for Example Case . . . . . . . . 83
Table 5.4. Simulated Asteroid Data . . . . . . . . . . . . . . . . . . . . . . 88
Table 5.5. Orbital Elements, Variable Chief Inclination . . . . . . . . . . . 88
Table 5.6. Orbital Elements, Variable Chief Semimajor Axis . . . . . . . . . 99
10

Abstract

RELATIVE ORBITAL MOTION DYNAMICAL MODELS FOR


ORBITS ABOUT NONSPHERICAL BODIES
by

Ethan R. Burnett

Master of Science

University of Arizona

Tucson, Arizona, 2018

Dr. Eric A. Butcher, Chair

Relative orbital motion dynamical models are presented and discussed. Two types
of models are primarily discussed in this work: a linear relative motion model account-
ing for J2 (the gravitational parameter associated with the oblateness of the Earth),
and a new linear relative motion model accounting for both nonzero second degree
and order gravity terms C20 = −J2 and C22 . The latter model, referred to as the
“second-order model,” is useful for simulating and studying spacecraft relative motion
in orbits about uniformly rotating asteroids. This model is derived in two alternate
forms. The first makes use of averaging in the kinematics and the second avoids any
use of averaging. Additional work is devoted to analyzing the stability of relative
orbital motion in rotating second degree and order gravity fields. To facilitate this, a
parameter called the relative orbit angular momentum is introduced.
For commensurate angular rates of primary body rotation and orbital mean mo-
tion, a special case linear time-invariant (LTI) model is obtained from the unaveraged
second-order model. This is connected to the topic of libration points in the body
frame of rotating gravitating triaxial ellipsoids, and shown to successfully predict the
instability of libration points collinear with the long axis and the stability of libration
points collinear with the short axis. Analytical and numerical results confirm the
accuracy of all models discussed.
11

1. Introduction

The need to coordinate the relative orbital motion of multiple spacecraft is present in
many modern scientific, commercial, and manned space missions, as well as in orbital
space station construction. Numerous analytical challenges have been introduced by
the small separation and even smaller margin of error necessary for space vehicle
coordination in these scenarios. This has led to the burgeoning study of spacecraft
relative motion, a demanding and increasingly relevant subject within the larger field
of the analytical mechanics of space systems.
In the short history of spaceflight, much of the research in spacecraft relative
motion has been based on linearized dynamic models that neglect the perturbations
that affect real orbital motion. The Hill-Clohessy-Wiltshire (HCW) model is the
most famous and well-known example of such models.1 While useful for short-term
applications like terminal guidance and rendezvous, this model is less useful for longer
time spans, after the effect of the neglected orbital perturbations (namely higher-order
gravitational terms and drag) have sufficiently accumulated. This makes it inherently
undesirable for use in long-term formation flying missions.
In recent work, researchers have proposed different methods and models for incor-
porating the effect of perturbations on relative orbital motion. Much of this work has
been focused on accounting for the perturbations due to the J2 zonal harmonic, which
are considerable in low-Earth orbits. Two major approaches have been undertaken
for incorporating the effects of J2 : state transition matrices (STMs) and ODE models.
The STMs and ODE models typically describe the relative motion of the spacecraft
in the local-vertical-local-horizontal (LVLH) frame of the chief, but some also use the
orbit element differences or relative orbit elements (ROEs). A comparison of these
approaches was given in Johnson et al.2 Gim and Alfriend developed a well-known
state transition matrix (called the GA STM) that incorporates the J2 perturbation.3
12

It maps orbit element differences and makes use of a spherical coordinate description
of relative motion in the LVLH frame. Use of Brouwer-Lyddane theory4, 5 enables
mapping between the mean and osculating orbital elements such that the STM can
be expressed with either set. The STM is an analytical solution that explicitly de-
pends on the orbit elements of the chief. This is advantageous because the model is
valid for any eccentricity (a common challenge for linear ODE models), but the ex-
plicit dependence on a computed or observed set of chief orbit elements is a potential
disadvantage. The reliance on a continuous update of information about the chief
orbit may not always be desirable or achievable in on-board computers.
Several modifications of the GA STM have been made to include drag6 and all
zonal harmonics.7 Riggi and D’Amico8 proposed a modified STM that enables sepa-
ration of in-plane and out-of-plane motion using a new set of orbital elements. Finally,
Biria and Russell10 also obtained a new STM using Vinti’s intermediary, which incor-
porates both the J2 and J3 perturbations. The linearized restriction of the GA STM
is retained by all of these additional works.
One of the advantages of some ODE models over STM models is that they
sometimes do not rely on a continuous update of information about the chief or-
bit. Instead, they may only make use of the chief orbit elements at some initial
time. Furthermore, ODE models, especially LTI models or LTV models of the form
ẋ = A (t) x = A (t + T ) x can be analyzed with the many useful tools from linear
systems theory and Floquet theory, providing the potential for additional dynamical
insight. Such insights are often more difficult to extract from STM models. Both ODE
models and STM models offer the potential for closed-form, time-explicit solutions,
which have many uses ranging from terminal guidance to relative orbit determina-
tion. However, it is often easier to include nonlinear and perturbation effects in ODE
models than in STM models.
Some researchers have developed nonlinear ODE models to account for larger sepa-
ration or eccentric chief orbits. These can yield time-explicit solutions via straightfor-
13

ward perturbation expansions9 or other means. In 1963 London11 obtained second-


order nonlinear equations in this manner to account for larger separation, and de
Vries12 obtained linear time-varying equations to account for eccentric chief orbits.
Both types of perturbations were included to second order by Anthony and Sasaki
in 1965.13 In 2016, Butcher et al. developed third-order Cartesian relative motion
perturbation solutions for slightly eccentric chief orbits,14 then extended this work
to account for larger chief eccentricity in both Cartesian and spherical coordinate
perturbation solutions.15
In 2002, Schweighart and Sedwick16, 17 obtained a linearized ODE model that at-
tempted to correct for the effects of the J2 perturbation. Their approach was to
time-average the gradient of the J2 potential to obtain constant coefficient linearized
equations. They noted that such averaging resulted in a loss of some information
about the perturbed relative motion, and they made efforts to correct for this. Their
procedure introduced analytic corrections to an initially unperturbed chief orbit, in-
stead of treating the kinematics of the perturbed LVLH frame in the more formal
manner of describing the angular velocity of the frame in terms of the perturbed
orbit element rates, outlined by many sources such as Prussing and Conway,18 and
implemented by Casotto.19 However, Casotto’s implementation still did not result in
a stand-alone model like that of Schweighart and Sedwick. There is thus a need for
developing a relative motion model that incorporates the strengths of both of these
models. Furthermore, a linearized model which accounts for both C20 = −J2 and
C22 has not been previously developed, but could prove to be relevant for spacecraft
rendezvous/docking and formation flying in orbits around asteroids. Asteroids are
increasingly mentioned as targets for exploration by space agencies and even private
companies, due to their mystery to science, potential risk to human civilization, and
opportunity for future resource extraction.
This thesis is organized as follows: Chapters 2 and 3 introduce much of the nec-
essary background material for the work in this thesis, but also contain some new
14

analytical arguments and concepts, and the results of two carefully written numer-
ical simulations. Chapter 4 focuses on the derivation and testing of a linearized
J2 -perturbed relative motion ODE model developed in Butcher et al.20 This model
treats the kinematics more formally (an advantage of Casotto’s work) and the per-
turbation solution for the linear ODEs provides a stand-alone model.
Chapter 5 introduces a completely new linearized relative motion ODE model
which incorporates the effects of C20 and C22 . This model assumes that the primary
body is a triaxial ellipsoid, such that all nonzero higher-order gravitational coefficients
can be expressed as polynomials in terms of C20 and C22 . The model derivation
assumes that these higher-order gravity terms are small enough to ignore, and that
the primary body is in stable rotation about its axis of maximum inertia. Further
analytical work is done with this model, and it is tested extensively with numerical
simulations. Finally, zeroing the C22 terms in this linearized model also provides
a new J2 = −C20 model that seems to offer improved performance over the model
developed in Butcher et al.
15

2. Orbital Motion and Relative Motion

2.1 Classical Theory of Satellite Orbits

This section serves as a basic, non-exhaustive review of introductory orbital mechanics


and the two-body problem. We begin with a discussion of gravitational forces and
then derive the two-body problem and include a brief review of the geometry of
the bound orbits that occur in the two-body problem. There are numerous texts
that discuss this subject, but this discussion is adapted from Prussing and Conway’s
Orbital Mechanics,18 Vallado’s Fundamentals of Astrodynamics and Applications,21
and Battin’s An Introduction to the Mathematics and Methods of Astrodynamics.22

2.1.1 Gravitation

From Newton’s law of gravitation, gravity is an attractive force between two bodies in
proportional to the product of their masses and inversely proportional to the square
of the distance that separates them:

Gm1 m2
F= r̂ (2.1)
r2
where G is the gravitational constant, and G = 6.67259 × 10−11 m3 kg−1 s−2 . For a
system of n masses, for which rij = rj − ri is a vector from mass i to mass j, the
attraction felt by the i th mass is given below, where δij is the Kronecker delta.

n
X mi mj
Fi = mi r̈i = G (1 − δij ) 3
rij (2.2)
j=1
rij

It can be easily shown using this expression that the linear and angular momentum
of this system is constant, along with the total mechanical energy.
16

2.1.2 The Two-Body Problem and Elliptic Orbits

Using equation 2.2 with n = 2:

Gm1 Gm2
r̈2 − r̈1 = 3
r21 − 3 r12 (2.3)
r12 r12

Dropping the subscript so r12 = r and rearranging, we arrive at the two-body equa-
tion:
G (m1 + m2 ) µ
r̈ = − 3
r = − 3r (2.4)
r r
In the case where m1 >> m2 , µ ≈ Gm1 . This standard gravitational parameter is
catalogued for all large well-known celestial bodies in the solar system.
Vector manipulations of equation 2.4 result in transformed forms that are per-
fect differentials, which can be integrated directly. The constants of this integration
are called the integrals of motion or orbital elements.22 There are many formula-
tions of orbital elements, but in this thesis we will use only the most common sets
which describe the geometry of bound orbits in the two-body problem. These will be
introduced after the geometry of such bound orbits has been introduced.
Crossing equation 2.4 with r, we obtain:

µ
r × r̈ + r × r=0 (2.5)
r3

This is integrated to obtain:


r × ṙ = h (2.6)

where h is the angular momentum vector, which is a constant vector. Thus, the
two-body orbital motion (satisfying conservation of angular momentum) lies in the
plane normal to this vector.
To solve equation 2.4, take the cross product with h:
 
−µ −µ ṙ rṙ
r̈ × h = 3 r × h = 3 r × (r × ṙ) = µ − (2.7)
r r r r2
17

Re-arranging equation 2.7:


d r
r̈ × h = µ (2.8)
dt r
This can be integrated to yield:
r 
ṙ × h = µ +e (2.9)
r

The vector e is constant and normal to h, so it is fixed in the orbital plane. Dotting
equation 2.9 with r yields:

r · (ṙ × h) = (r × ṙ) · h = h2 = µ (r + r · e) = µr (1 + e cos f ) (2.10)

where f is the angle between r and e. Rearranging and solving for r, we obtain:

h2
r= (2.11)
µ (1 + e cos f )

Isolating p = h2 /µ and defining this variable (with length dimensions) as the parame-
ter, we note that it can also be defined in terms of the dimensionless quantity e and a
new (for now, undefined) quantity a, which has dimensions of length: p = a (1 − e2 ).
Substituting this into equation 2.11, we obtain an alternate form: r = p/ (1 + e cos f ).
This is a polar coordinate representation of the equation of the orbit. It can be shown
that for e < 1, this describes an ellipse, given in Fig. 2.1 (reproduced from Battin22 ).
Now we define a as the semimajor axis, and b in the figure is the semiminor axis,
p
b = a2 (1 − e2 ). Two important points on the orbit are the periapsis (f = 0) and
the apoapsis (f = π). Note that for e = 0, we have a circular orbit and these points
(along with f ) are undefined. In this case, f must be replaced with an orbit angle
measured from a point of interest in the orbit. This subject will be revisited during
the discussion of orbital elements.
Using the equations already defined in this section, we can obtain many of the
common equations used to describe two-body orbital mechanics. A few more useful
equations will be presented for reference, but without derivation. First, the period
18

Figure 2.1. Elliptical Orbit Geometry22

of an orbit T and mean motion n (the mean angular rate of the orbiting body) are
defined: s
a3
T = 2π (2.12)
µ
r
2π µ
n= = (2.13)
T a3
The mean motion n defines the mean anomaly M , which is a linearly increasing
angular variable that will be useful for several expansions and equations used in the
following chapters:
M = n (t − tp ) (2.14)

where tp is the time of periapsis passage.

2.1.3 Orbital Elements for Elliptic and Circular Orbits

It is necessary to describe the orientation of the elliptical or circular orbit with respect
to an (assumed) inertially fixed coordinate system. This is done with a 3-1-3 Euler
angle rotation through the angles Ω (called the right ascension of the ascending node,
RAAN), i (called the inclination), and ω (called the argument of periapsis).23 The
rotation from an inertially fixed coordinate system to the inertially fixed orbit plane
(using these angles) is depicted in Fig. 2.2 (reproduced from Schaub and Junkins23 ).
19

Figure 2.2. The 3-1-3 Euler Angles and the Orbit Plane23

Note that the argument of latitude θ = ω + f is also defined in the figure. This
is the angle between the ascending node and the current location in the orbit of the
satellite. This angle must be used for inclined circular orbits, since the argument of
periapsis ω is undefined. For equatorial circular orbits, the angular position of the
orbiting satellite must be measured from the inertially fixed reference direction îx .
Since a 3D orbit is a 3 degree-of-freedom problem, we need six quantities to define
the state of a satellite in an elliptical orbit. The most common orbit element set
to be used in this thesis is [a, e, i, ω, Ω, f ], and all of these quantities have already
been defined in this chapter. Note that for circular orbits, this set is undefined, and
the state can be described using fewer than 6 quantities. For example, for a circular
inclined orbit, we would use the orbit element set [a, i, Ω, θ].

2.2 Analytical Relative Dynamics of Co-orbiting Bodies

Often, we are interested in the simultaneous behavior of multiple orbiting spacecraft.


This is especially true if the spacecraft are in similar orbits (either due to coincidence
or to fulfill a certain mission). In this section, we will describe the motion of a
20

secondary spacecraft with respect to a primary spacecraft, in a coordinate system


centered on the primary. This relative orbital motion description can be very useful
for formation flying and rendezvous and docking of multiple spacecraft.

2.2.1 Relative Orbital Motion

In the spacecraft relative motion problem, the motion of the secondary (deputy)
spacecraft is described in a rotating frame centered on the primary (chief) spacecraft.
This frame is called the Hill frame, or equivalently the Local Vertical-Local Horizontal
(LVLH) or R-T-N frame. The problem geometry is depicted in Fig. 2.3 below.

Figure 2.3. Relative Motion Problem Geometry

The positions of the chief and deputy spacecraft with respect to the primary body
are given by rc and rd , respectively. The vector from the chief to the deputy is
∆r = rd − rc , and it is resolved in its LVLH components ∆r = xêr + yêt + zên . The
unit vectors êr , êt , ên are defined:

rc rc × ṙc
êr = , ên = , êt = ên × êr (2.15)
rc krc × ṙc k
21

The relative motion dynamics are obtained by first treating the kinetics of the
problem, by subtracting the accelerations acting on the chief from those acting on
the deputy:
∆r̈ = a (rd ) − a (rc ) (2.16)

where a (r) = − rµ3 r if we are only considering two-body gravitational dynamics.


Treating the kinematics involves resolving the inertial acceleration of the relative
position vector into the Hill frame, which rotates with angular velocity ω H . This is
done by applying the transport theorem twice:
H 2 H
d ∆r d∆r
∆r̈ = 2
+ ω̇ H × ∆r + 2ω H × + ω H × (ω H × ∆r) (2.17)
dt dt
Pure two-body dynamics are not encountered in reality; there are always perturb-
ing accelerations acting on any orbiting satellite. The nature of these perturbations
will determine the exact forms of a (r) and ω H . In Earth orbits, these perturbations
are generally small and two-body dynamics will approximate the behavior of orbiting
satellites very well for short time spans.

2.2.2 The Hill-Clohessy-Wiltshire Model

The Hill-Clohessy-Wiltshire model approximates the relative motion of satellites in


similar orbits in the two-body problem. While George Hill also obtained linearized
relative motion equations to describe the Moon’s orbit in a rotating Earth-centric
frame, the best known introduction of this model is from 1960, in which it was used
to design a preliminary terminal guidance system for satellite rendezvous.1 This
simple but useful model (which is restricted for use with small separations, small
perturbing accelerations, and a circular chief orbit) will now be derived.
We begin by treating the kinetics of the problem, by subtracting the two-body
accelerations acting on the chief from those acting on the deputy:
µ µ
∆r̈ = − r + 3 rc
3 d
(2.18)
rd rc
22

Noting rd = rc + ∆r, we expand the equation:

µ µ
∆r̈ = − 3 (rc + ∆r) + rc (2.19)
((rc + ∆r) · (rc + ∆r)) 2 rc3

Factoring the denominator of the deputy acceleration term, and keeping only terms
linear in ∆r:  − 32
µ (∆r · rc ) µ
∆r̈ ≈ − 3 1+2 (rc + ∆r) + rc (2.20)
rc rc2 rc3
Note that linearization renders the equality in equation 2.19 as an approximation in
equation 2.20, but the approximation notation will be dropped. Using the first term
in a binomial expansion ((1 + )k ≈ 1 + k) to linearize the exponential term, and
factoring:   
µ (∆r · rc )
∆r̈ = − 3 1−3 (rc + ∆r) − rc (2.21)
rc rc2
Expanding the expression in parentheses and keeping only terms linear in ∆r, drop-
ping the subscript notation so rc = r:
 
µ (∆r · r)
∆r̈ = − 3 ∆r − 3 r (2.22)
r r2

Noting r̂ = r/r and expressing this in dyadic notation:

µ  
∆r̈ = − I − 3r̂r̂ · ∆r (2.23)
r3

where I is the unit dyadic, and r̂r̂ is also a dyadic. For readers unfamiliar with dyads,
the pair ab is called a dyad or direct product with antecedent a and consequent b.
A dyad maps any vector v into a vector parallel to a according to the definition
ab · v = a (b · v).27
We resolve ∆r into its LVLH components as ∆r = xêr + yêt + zên , and noting
r̂ = êr , we obtain the linear terms due to the difference in acceleration felt by the
deputy and chief:
µ
∆r̈ = (2xêr − yêt − zên ) (2.24)
r3
23

For a circular chief orbit, r = a, so we can rewrite this using the definition of the
mean motion:
∆r̈ = 2n2 xêr − n2 yêt − n2 zên (2.25)

We must resolve the inertial acceleration of the relative position vector into the LVLH
frame as well, which rotates with angular velocity ω H = nên . This is done by applying
the transport theorem twice:
H 2 H
d ∆r d∆r
∆r̈ = + ω̇ H × ∆r + 2ω H × + ω H × (ω H × ∆r) (2.26)
dt2 dt
If the chief spacecraft is in a circular orbit, ω H = nên is constant:
H 2 H
d ∆r d∆r
∆r̈ = 2
+ 2ω H × + ω H × (ω H × ∆r) (2.27)
dt dt
Resolving this in LVLH components and simplifying:

∆r̈ = ẍ − n2 x − 2nẏ êr + ÿ − n2 y + 2nẋ êt + z̈ên


 
(2.28)

Setting the vector equation 2.28 equal to equation 2.25, simplifying, and separating
by component, we obtain the three well-known HCW equations:
ẍ = 3n2 x + 2nẏ

ÿ = − 2nẋ (2.29)

z̈ = − n2 z
The solution to this set of linear ODEs can be shown to be:18
sin nt 2
x (t) = (4 − 3 cos nt) x0 + ẋ0 + (1 − cos nt) ẏ0
n n
2 4 sin nt − 3nt
y (t) = 6 (sin nt − nt) x0 + y0 − (1 − cos nt) ẋ0 + ẏ0 (2.30)
n n
sin nt
z (t) = cos ntz0 + ż0
n
Taking the time derivative of these equations, we can easily obtain ẋ (t) , ẏ (t) , ż (t)
as well:
ẋ (t) = 3n sin ntx0 + cos ntẋ0 + 2 sin ntẏ0

ẏ (t) = − 6n (1 − cos nt) x0 − 2 sin ntẋ0 + (4 cos nt − 3) ẏ0 (2.31)

ż (t) = − n sin ntz0 + cos ntż0


24

Since these solutions to the HCW equations are linear in the initial conditions, we
could use them to obtain a state transition matrix Φ (t), mapping the initial state
∆s> (0) = [x0 , y0 , z0 , ẋ0 , ẏ0 , ż0 ]> to the state at any time t, ∆s> (t) = [x, y, z, ẋ, ẏ, ż]> .
There are many uses for the resulting STM for spacecraft relative motion, including
linear impulsive rendezvous for bringing two spacecraft in similar near-circular orbits
close enough for docking or other proximity operations.
Note that by collecting terms linear in time from y(t) in equation 2.30, we obtain
−3 (2nx0 + ẏ0 ) t, which isolates a “no-drift” condition when set to zero:

2nx0 + ẏ0 = 0 (2.32)

If the initial conditions (x0 , ẏ0 ) satisfy this equation, then the HCW equations yield
completely bounded periodic relative motion. It can be shown that this condition is
essentially a linearized approximation of δa = ad − ac = 0, which ensures that both
the chief and deputy orbit have the same orbit period and their relative motion will
remain bounded under two-body dynamics.23 Since it is only an approximation of
the no-drift condition, non-drifting relative orbit initial conditions can (and generally
will) still result in slow drift in the HCW solution.
The linearized no-drift condition reappears in an interesting manner if the HCW
position and velocity equations are used to obtain an expression for what is labeled
here as the “relative orbit angular momentum” ∆h = ∆r × ∆v. If the HCW position
and velocity solutions are substituted for the R-T-N components of ∆r and ∆v, we
obtain ∆hHCW , whose R-T-N components are given in equations 2.33 – 2.35 below,
where Υ = 2nx0 + ẏ0 .
2 
∆hHCW (r) = − 2(3nx0 + 2ẏ0 )z0 + ẋ0 ż0 + 3Υz0 cos nt + n (3tΥ − y0 ) z0 sin nt
 n
  
2 3
+ − ẋ0 + (y0 − 3tΥ) ż0 cos nt + 2ẋ0 + Υż0 sin nt
n n
(2.33)
 
2 2
∆hHCW (t) = ẋ0 z0 + 3x0 + ẏ0 ż0 + Υ (−ż0 cos nt + nz0 sin nt) (2.34)
n n
25

2  cos nt
21n2 x20 + ẋ20 + 24nx0 ẏ0 + 7ẏ02 + 2 ẋ20 + 7ẏ02 + 3ntẋ0 Υ
 
∆hHCW (n) = −
n n
2 2
 cos nt 2
 sin nt
+ 42n x0 + n (49x0 ẏ0 − ẋ0 y0 ) − 3 ẋ0 ẏ0 + n x0 y0
n n
sin nt 2
 sin nt
− (8nx0 ẋ0 + 2ny0 ẏ0 ) + 9n x0 t + 6nẏ0 t Υ
n n
(2.35)
If the linearized no-drift condition is perfectly satisfied, then Υ = 0 and the three
components of ∆hHCW above become bounded periodic functions of time.
Consider cases where Υ = 0 and there is no along-track offset of the relative orbit
of the deputy around the chief (such that δe and δi are the only nonzero orbit element
differences). Then, defining t = 0 to be the time when rd and rc are collinear, we
know y0 , z0 , and ẋ0 are zero. Thus, ∆hHCW reduces to the following constant vector:
 
2 2
21n2 x0 + 24nx0 ẏ0 + 7ẏ02 ên

∆hHCW = 3x0 + ẏ0 ż0 êt − (2.36)
n n

In a non-drifting unperturbed relative orbit defined only by orbit element differences


δe and δi, the deputy spacecraft moves in a periodic elliptical “relative orbit” around
the chief in a plane that is fixed in the LVLH frame and perpendicular to êr . This
type of relative orbit is frequently examined in the literature and in this thesis. HCW
dynamics predict that the relative orbit angular momentum is conserved in this case,
but the full two-body dynamics can be used to show that this is not true.
It can be easily shown (using the definition of orbit angular momentum) that for
two-body dynamics, absent perturbations, the following expression is true:

∆h = (∆r × ∆v) = − (rc × ∆v) − (∆r × vc ) (2.37)

Taking the derivative of both sides, we can show:


 
d µ µ
(∆h) = − (rd × rc ) (2.38)
dt rc3 rd3

For a non-drifting relative orbit (even one parameterized only by δe and δi), the time
variations in ∆h are very small (assuming small separation) but still clearly nonzero.
26

However, we can say that for most non-drifting relative orbits of interest, absent
perturbations, ∆h is a “nearly conserved” quantity: it is subject to small, bounded
variations about some mean value. This is easy to show. If the chief and deputy
orbits are unperturbed and have the same orbit period, then the relative orbit will be
non-drifting and ∆r and ∆v will be orbit-periodic functions. Thus ∆h = ∆r × ∆v
will also be periodic. Later in this chapter, it will be shown that variations in ∆h due
to perturbations will dwarf the small variations in ∆h in unperturbed non-drifting
relative motion. It should be clear to the reader that since ∆h = ∆r × ∆v, significant
changes in ∆h imply significant changes in the relative orbit behavior. This subject
will be revisited later in this chapter.

2.2.3 Representing Model Error

It is important to test approximated relative motion models for various scenarios,


both for validation and to ensure that they are sufficiently accurate for a particular
application. This is done by comparing them to a “truth” model. The truth model
represents the best appropriate mathematical model of the true dynamical behavior.
The “true” relative motion is obtained by numerically simulating the chief and deputy
orbits separately, then resolving the deputy-chief state difference into the moving
LVLH frame at each time step. This is a relatively straightforward procedure, but it
can be numerically costly. Note that the fidelity of the truth model should be adjusted
accordingly depending on the goal of an error study. For example, if an error study
is being done to compare the accuracy of two models that both are derived without
any consideration of drag effects, it would be inappropriate for the truth model to
include drag. However, if an error study is being done to show the error resulting
from the HCW model neglecting drag, the truth model must obviously have a good
drag model.
There are many ways to represent error in a comprehensible manner, but in this
27

thesis, nearly all model error results will be conveyed by only 5 types of plots. First,
note that we will generally ignore the velocity errors of the relative orbit models, since
the position error sufficiently demonstrates the deviation between the truth and any
approximate models. The first two types of plots will represent model position error
vs. time for a single simulation. The latter three are used for large case studies with
many simulations:

1. Relative orbit plots with truth model and approximate model results

2. Absolute position error by component vs. time

3. Max position error norm in a time range (per case) vs. a varied parameter

4. Average position error norm in a time range (per case) vs. a varied parameter

5. Some computed physical quantity (per case) vs. a varied parameter

Let the relative orbit position error from the simulation of a model be given at time t
by subtracting the true relative position from the model-predicted relative position:

E∆r (t) = ∆rM (t) − ∆rT (t) (2.39)

Resolving this error into its LVLH components, taking the absolute values, and plot-
ting vs. time, we would produce the second plot type listed above. To obtain the
third type of plot, we would find max(kE∆r (t) k) for each simulation in a case study,
and plot vs. the varied parameter chosen for that study. The average position error
norm is given by averaging over some length of time (typically one orbit period T ):

1 T
Z
E ∆r = kE∆r (t) kdt (2.40)
T 0

Generally, the variational behaviors of max(kE∆r (t) k) and E ∆r are very similar in
most parameter studies. The main difference between these two quantities is that
average position error norm places equal “weight” on all error over time, while the
28

max position error norm obviously only gives information about the highest error
achieved by the model over a given time span.
The fifth way for representing model error is quite different from the preceding
four. The idea is to use the model results to calculate the value of some time-
varying physical quantity, and to compare the maximum, minimum, or average
value of that estimated quantity with its counterpart calculated using the true data.
This is a potentially useful analytical approach because it can convey how well a
model is representing the true physics of the problem. The most frequently used
physical quantity is the “max deviation of the relative orbit angular momentum”,
max(k∆h (t) − ∆h (t0 ) k). After a discussion of orbital perturbations, it will be
shown that this is very useful for showing the degree to which the relative orbital
motion has been affected by perturbations. Furthermore, if an approximate model is
developed to account for additional perturbations, then we should typically be able
to estimate the value of this quantity with a reasonable degree of accuracy using the
model results. This gives us an additional metric for gauging the relative performance
of approximate models.

2.3 Orbital Perturbations

Two-body orbital mechanics generally fails to predict satellite motion with useful
accuracy over long time spans, since it neglects the many additional forces acting on a
real orbiter. This section serves as a brief introduction to the mechanics of perturbed
orbital motion and relative orbital motion. We consider the effects of a general,
unspecified perturbation before a brief discussion of the types of perturbations.

2.3.1 Perturbed Orbital Motion

It is often convenient to consider the effects of a perturbing acceleration aP on the


two-body orbit elements. Whereas five of these elements were constant for the un-
29

perturbed two-body problem, all six of these elements may vary over time due to the
perturbing acceleration. Given below are Gauss’ variational equations, which describe
the instantaneous variation of the orbital elements, and are valid for conservative and
non-conservative disturbance accelerations.23 First, note that the disturbing acceler-
ation is resolved into its R-T-N components as aP = RP êr + TP êt + NP ên .

da 2a2  p 
= e sin f RP + TP (2.41)
dt h r
de 1
= (p sin f RP + ((p + r) cos f + re) TP ) (2.42)
dt h
di r cos θ
= NP (2.43)
dt h
dΩ r sin θ
= NP (2.44)
dt h sin i
dω 1 r sin θ cos i
= (−p cos f RP + (p + r) sin f TP ) − NP (2.45)
dt he h sin i
df h 1
= 2+ (p cos f RP − (p + r) sin f TP ) (2.46)
dt r he
Recall that θ = ω + f is the argument of latitude.

2.3.2 Types of Perturbations

The common perturbations that act on orbiting satellites are typically very small, but
the precision necessary for spaceflight demands that they be considered and accurately
modeled. Nonspherical body gravity is present to some degree in orbits around any
celestial body. This is an important topic in this thesis, so the deviation from a spher-
ical gravity field will be discussed extensively in the next chapter. The gravitational
forces from other celestial bodies may also be necessary to consider. In Earth orbit,
for example, the gravitational forces of the Moon and Sun exert noticeable influences
on satellite motion.
Some perturbation forces are highly attitude dependent. In low-Earth orbit, drag
forces from the thin upper atmosphere will cause orbits to decay over time. The
30

magnitude of these forces is proportional to the area of the spacecraft projected in the
direction of its orbital velocity. Solar radiation pressure is another perturbation that
will almost always be present to some degree. This is a force exerted on satellites due
to the momentum exerted on the spacecraft by absorption and reflection of incident
solar radiation. It produces much more noticeable effects for high area-to-mass ratio
(HAMR) objects, and is highly dependent on the projected area of the spacecraft
facing the sun.

2.3.3 Perturbed Relative Orbital Motion

For studying relative orbital motion in a perturbed environment, we must keep track
of how the chief LVLH frame is moving with respect to the inertial frame. The time-
varying angular velocity of the chief LVLH frame can be expressed in its components
along the third inertial axis K̂, the nodal vector along K̂ × ên , and the normal axis
ên :19
di K̂ × ên
ω H = Ω̇K̂ + + θ̇ên (2.47)
dt kK̂ × ên k
This angular velocity should be substituted into equation 2.17 to treat the kinematics
of the relative motion problem. The orbit element rates can be obtained by substitut-
ing the particular perturbing acceleration aP into Gauss’ variational equations. This
is a simple procedure in principle, but in practice it can be challenging.
Recall that for unperturbed relative motion, we have the following expression for
the time derivative of the relative orbit angular momentum:
 
d µ µ
(∆h) = ∆r × ∆a = − (rd × rc ) (2.48)
dt rc3 rd3

where ∆a = (ad − ac ) is the difference in accelerations felt by the deputy and chief.
For the unperturbed two-body problem, the relative orbit angular momentum is
bounded and orbit-periodic. The addition of a perturbing acceleration to the problem
31

Table 2.1. Orbital Elements and Differences


a (km) e i (◦ ) Ω (◦ ) ω (◦ ) f (◦ )
œc 40.0 0.0005 10.0 0.0 0.0 0.0
δœ 0.0 0.002 0.2 0.0 0.0 0.0

dynamics adds an additional term to equation 2.48:


 
d µ µ 
(∆h) = 3
− 3 (rd × rc ) + ∆r × aP (d) − aP (c) (2.49)
dt rc rd

Importantly, with the addition of perturbing accelerations, the orbital behavior of


the chief and deputy is typically no longer periodic, thus ∆r (t) 6= ∆r (t + T ) and
∆v (t) 6= ∆v (t + T ). The consequence of this is that the relative orbit angular
momentum loses its periodicity, and we have the possibility for much larger changes
than the bounded periodic oscillations in the unperturbed problem. This can be
demonstrated with an example.
Consider two cases for relative motion with the initial orbit elements and differ-
ences given in Table 2.1. In the first case, the chief and deputy satellites will be in
unperturbed orbits (with the same period) about a spherical body with a radius of
3.56 km and a density of 2.6 g/cm3 . The second case shares the same initial con-
ditions, but the body is changed to a rotating ellipsoid with semi-axes of 6, 3, and
2.5 km and a rotation period of 48 hours. In both cases, the primary bodies have
the same density and volume (and hence, the same mass). This example serves as
a demonstration of the potentially dramatic effects of a nonspherical body’s gravity
field on relative motion. The relative configuration of the orbit frame and body frame
is still under-defined since any additional details would be irrelevant to the present
discussion, but all data for this example will be given when this problem is revisited
in Section 5.3.5.
Fig. 2.4 shows the magnitude of separation between the chief and deputy for 4
unperturbed chief orbit periods (T = 77.2 hrs). Note that the unperturbed relative
motion (dashed line) is completely periodic, as expected. However, the addition of the
32

0.45

0.4

0.35

0.3

0.25

0.2

0.15

0.1

0.05
0 0.5 1 1.5 2 2.5 3 3.5 4

Figure 2.4. Relative Distance for Unperturbed and Perturbed Cases

10-6
1.5

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4

Figure 2.5. ∆h(t) for Unperturbed and Perturbed Cases

gravitational perturbations from the primary body clearly produces wildly perturbed
relative motion behavior (solid black line). The relative motion magnitude has grown
considerably starting around 0.5 chief orbit periods, and soon after the relative motion
33

has completely diverged from the unperturbed case.


Fig. 2.5 similarly shows the unperturbed and perturbed relative orbit angular mo-
mentum magnitude ∆h (t). Note that the unperturbed relative orbit angular momen-
tum magnitude oscillates periodically on a scale far too small to be seen in this figure.
The unperturbed ∆h oscillates by just ±5.4 × 10−10 about a mean of 3.8406 × 10−7 .
Since drastic changes in ∆h only appear due to perturbations, the maximum devi-
ation of the relative orbit angular momentum, max(k∆h (t) − ∆h (t0 ) k), is useful
for estimating the degree to which the relative orbit behavior has been affected by
the perturbations. Furthermore, we can use how well the model results predict this
quantity as an additional means for evaluating the model’s accuracy.
34

3. Nonspherical Body Gravitation and Orbital


Mechanics

3.1 Gravitational Potential Expressed by Spherical Harmonics

In this section, the gravitational potential expressed in terms of the spherical harmon-
ics will be introduced. This short review is adapted from Scheeres’ Orbital Motion in
Strongly Perturbed Environments 24 and Schaub and Junkins’ Analytical Mechanics of
Space Systems.23 Readers should consult these sources for more detailed information.
We begin by considering the gravitational potential for an arbitrary mass distri-
bution. The problem geometry is depicted in Fig. 3.1. Note that the Brillouin sphere
is a sphere whose radius is given by the distance of the furthest point on the body
from the expansion center (the origin of the coordinate system depicted).

Figure 3.1. Primary Body Mass Distribution Geometry

The overall body potential U is given by integrating the potential of a differential


35

mass element over the entire body:


Z
dm (ρ)
U (r) = G (3.1)
B kr − ρk

where B is the collection of all mass elements in the body, ρ is the position vector
for the differential mass element dm, and G is the gravitational constant. The grav-
itational potential satisfies Laplace’s equation outside of the body: ∇2 U = 0. For a
homogeneous sphere with constant density, mass M, and radius R, equation 3.1 yields
U = GM/r.
Laplace’s equation can be resolved in spherical coordinates:

∂ 2U
   
1 ∂ 2 ∂U 1 ∂ ∂U 1
r − 2 cos φ + 2 =0 (3.2)
r2 ∂r ∂r r cos φ ∂φ ∂φ r cos2 φ ∂λ2

The spherical coordinates can be defined for the position vector r = xêx + yêy + zêz :
p
r = x2 + y 2 + z 2 , λ = tan−1 (y/x) is the longitude, and φ = sin−1 (z/r) is the
latitude. Solution of equation 3.2 via separation of variables yields the spherical
harmonics, which are an orthogonal set of solutions that form a basis for describing
any other function satisfying Laplace’s equation.31 The potential for any arbitrary
physical gravity field can thus be expressed in a series in terms of the spherical
harmonics:
∞ l  l
µXX R
U (r, φ, λ) = Plm (sin φ)[Clm cos mλ + Slm sin mλ] (3.3)
r l=0 m=0 r

where R is the radius of the Brillouin sphere and µ is the gravitational parameter.
Plm are the associated Legendre functions, and Clm and Slm are the gravity field
harmonic coefficients, which are defined by the mass distribution of the body. Note
that this expansion is only valid outside of the Brillouin sphere of radius R. For an
expansion that is valid inside the sphere, we must express the potential in terms of
an orthogonal series of eigenfunctions satisfying Poisson’s equation: ∇2 U = −4πGσ,
where σ is the local density.25 This is not an important topic for this thesis.
36

In equation 3.3, note that m = n = 0 gives the “zeroth degree and order term”.
This is the largest contribution to the gravitational potential; a spherically symmetric
“bulk” term equivalent to assuming that the body is replaced by a homogeneous
sphere of equal mass. The higher-order gravitational terms add and subtract “mass”
to particular regions of the body to account for local departures from a spherical
shape. However, a few of the additional terms can be eliminated with prudent choice
of the body-fixed coordinate system. First, if the origin of the coordinate system is
chosen to coincide with the body center of mass, it can be shown that the first degree
and order terms are all zero: C11 = S11 = C10 = 0. Next, choosing the body-fixed
coordinate system to be aligned with the principal axes of inertia, we can eliminate
some of the second degree and order gravity coefficients: C21 = S21 = S22 = 0. Thus,
a simple second degree and order gravitational potential can be described with just
the coefficients C20 and C22 . It has been found that in general, the second degree
and order gravity field accounts for the majority of the perturbations on orbital
dynamics.24
Large bodies like the Earth and Venus are very nearly symmetric about their axis
of rotation. The gravitational potential of an axially symmetric body can be well-
approximated in terms of the so-called zonal gravitational harmonics Jl = −Cl0 alone.
These describe contributions to the gravitational potential that vary with latitude
but not longitude. However, more accurate models of the Earth’s gravity field will
include additional harmonics: the sectoral harmonics Cll , which vary with longitude
but not latitude, and the tesseral harmonic, which trace a “checkerboard” pattern
with dependence on both latitude and longitude. For an excellent depiction of these
three types of terms in the spherical harmonic expansion, see Vallado’s Fundamentals
of Astrodynamics and Applications.21
37

3.2 The Shape of the Earth

It has been known since the time of Newton that the best simple shape to approximate
the figure of the Earth is not that of a sphere but of an oblate spheroid. The centrifugal
acceleration due to the Earth’s rotation causes it to bulge slightly around the equator.
The gravitational effect of this bulge (which has noticeable effects on the motion of
low Earth-orbiting satellites) is well-captured by considering the J2 zonal harmonic
from the spherical harmonic expansion of the gravitational field. The first six zonal
harmonics for the Earth are given by J2 = 1082.63 × 10−6 , J3 = −2.51 × 10−6 ,
J4 = −1.61 × 10−6 , J5 = −0.15 × 10−6 , and J6 = 0.57 × 10−6 . It is clear that J2 is by
far the largest, causing a noticeable precession of satellite orbits.23
While modeling the Earth’s gravitational potential with J2 alone provides a vast
improvement in simulation accuracy, the number of terms used will depend on the
fidelity required. Sufficiently high-fidelity models may make use of hundreds of grav-
itational coefficients. For example, the official Earth Gravitational Model EGM2008
provides the Earth’s gravity field spherical harmonics to degree and order 2159.

3.3 Orbital Motion under the Influence of J2

By inserting the J2 perturbing acceleration (see equations 4.15 and 4.16 in Chapter
4) into Gauss’ variational equations (equations 2.41 - 2.46), and orbit-averaging the
results, we obtain the well-known nonzero secular orbit element rates due to J2 :22
 2
dΩ 3 R
= − J2 n cos i (3.4)
dt 2 p
 2
dω 3 R
n 5 cos2 i − 1

= J2 (3.5)
dt 4 p
Equation 3.4 provides the average rate of the regression of the node: the plane of the
orbit rotates about the Earth’s polar axis in the opposite direction of the motion of
the satellite. Equation 3.5 describes the mean rate of rotation of the line of apsides.
38

Figure 3.2. Nodal Regression due to J2 21

From equation 3.5, it can be easily shown that a critical angle of i = 63.4◦ determines
whether the line of apsides regresses or advances. A common (and physically similar)
analogy to the precession of the orbit plane is the precession of a spinning top in a
uniform gravitational field. The precession of the orbit is clearly shown in Fig. 3.2, a
figure borrowed from Vallado’s Fundamentals of Astrodynamics and Applications.21
Note that the value of J2 was exaggerated by a factor of 20 to produce this figure.

3.4 Moons, Planetoids, and Asteroids

Large celestial bodies are nearly spherical because their gravitational forces are strong
enough that their constitutive material is compressed and coalesced into a nearly
uniform sphere. There is a common oblate distortion of the body due to centrifugal
effects resulting from rotation. However, these forces are not sufficiently powerful to
smooth out the considerable shape distortions seen in smaller bodies. Generally, the
smaller the celestial body, the more it deviates from a spherical (or at least axially
39

Figure 3.3. Asteroids and Comets Visited by Spacecraft24

symmetric) shape.
Small asteroids typically have highly irregular shapes. Fig. 3.3 is a compilation
of asteroids and comets imaged by spacecraft (borrowed from Scheeres24 ). The shape
of an asteroid can be roughly estimated by using light curve data. Light curves
are photometric observations of a body that allow an observer to estimate the total
projected area of a body over time. From this data, it is possible to estimate a
triaxial ellipsoid that best fits the asteroid. This best-fit ellipsoid is sufficient to
roughly estimate the dominant second degree and order gravitational field terms,24
with the estimate error depending on how much the body shape deviates from its
best-fit ellipsoid.

3.5 Orbital Motion about Small Asteroids

Orbital motion about small asteroids is challenging to model and predict accurately.
A particular challenge is the near-parity of solar radiation pressure (SRP) forces
and the gravitational forces of the asteroid. There is also an interesting interplay
40

between the time-varying nature of the asteroid gravity field felt by the orbiter (due
to rotation of a complex non-spherical gravity field) and the orbital dynamics. This
introduces a potential for orbital resonances which may eject an orbiting satellite,
or even cause it to impact the asteroid. It is also common for SRP forces to slowly
perturb the orbit until it is ejected. The terminator plane is a plane normal to the line
connecting the sun and the asteroid center of mass. Orbiting in this plane minimizes
the destabilization of the orbit due to SRP forces. However, the initial orbital elements
must be chosen carefully so that Gauss’ variational equations (with aP composed of
the SRP disturbance acceleration and, ideally, the second-order gravity terms as well)
predict that the orbit semimajor axis and eccentricity will not be subject to secular
changes.
The highly perturbed and irregular nature of orbits around small bodies can be
illustrated with an example. Consider a small asteroid whose properties are given in
Table 3.1. This asteroid has properties chosen to roughly match the estimates for
2016 HO3, and the simulation uses ephemeris data from July 31, 2016 (Julian date
2457600.5) from JPL’s Small-Body Database. Note that its shape is approximated
by a triaxial ellipsoid, and its gravity field is truncated to second degree and order
terms for the purpose of simulation.

Table 3.1. Asteroid Physical Parameters


Semi-axes A,B,C (m) 50, 25, 20
3
Density (g/cm ) 2.0
Rotation Period (hr) 5.56
Mass (kg) 2.094 ×108
Gravitational Parameter 1.3978×10-11
Gravity Constants C20 = −0.093, C22 = 0.0375

We are interested in simulating the motion of a spacecraft orbiting the asteroid.


The spacecraft has a mass-to-area ratio of 33 kg/m2 , and a low reflectance. We use a
simple cannonball SRP force model, which assumes that the SRP force is not attitude
41

dependent.26 We define the “Sun-asteroid frame” as a moving frame whose x-axis is


parallel with the line connecting the Sun and the asteroid center of mass, and whose
y-axis lies in the plane of the asteroid orbit. The z-axis is thus perpendicular to
the asteroid orbital plane. Assume that the asteroid is initially oriented by a π/2
counterclockwise z-axis rotation from the Sun-asteroid frame to its body frame, and
rotates counterclockwise about its axis of maximum inertia, which is aligned with the
inertial z axis.
With the sun-asteroid frame thus defined, the initial position and velocity of the
orbiter are (in meters and m/s) r0 = 120êy and v0 = 0.011êz . Thus, the spacecraft
starts in the terminator plane with a very slow orbital velocity due to the small mass
of the asteroid. The orbital velocity is so low that the initial orbital period is just
over one full day. The negligible gravitational forces of this asteroid make the orbit
particularly susceptible to SRP forces, and after 32 days, the spacecraft is ejected
from orbit around the asteroid. Before ejection, the spacecraft alternates between
revolutions near the terminator plane and revolutions near a plane offset from the
terminator plane in the +x direction. Fig. 3.4 clearly shows the orbital behavior
around the asteroid for 32.4 days, and clearly shows the ejection due to SRP forces.
The sphere depicted is the Brillouin sphere, and the asteroid rotates within this
sphere. Fig. 3.5 shows the spacecraft orbital velocity (in the asteroid perifocal frame)
by component and by its norm over time. The quasi-periodic nature of this dynamical
system is evident from this plot.
The offset plane in which the spacecraft performs many revolutions is due to a
near-balancing of the SRP forces and the asteroid gravity in this region. However, the
gravity field of the asteroid is time-varying, which also influences the spacecraft orbital
behavior in a complicated manner. Overall, this example shows that the perturbed
orbit behavior around small bodies is very complex. Relative orbital motion in such an
environment is similarly dynamically interesting, however the relative motion models
discussed and developed in this thesis neglect the influence of SRP forces.
42

0.25

0.2

0.15

0.1

0.05

-0.05

-0.1

-0.15

0.2
0 0.7
-0.2 0.4 0.5 0.6
-0.4 0.1 0.2 0.3
-0.1 0

Figure 3.4. Spacecraft Orbital Ejection due to SRP Forces

0.02

-0.02
0 5 10 15 20 25 30 35
0.01

-0.01
0 5 10 15 20 25 30 35
0.02

-0.02
0 5 10 15 20 25 30 35
0.015

0.01

0.005
0 5 10 15 20 25 30 35

Figure 3.5. Spacecraft Orbital Velocity Before and During Ejection


43

4. A J2-Perturbed Relative Orbit Model

4.1 Formulation

This chapter focuses on the derivation and testing of a linearized J2 -perturbed relative
motion model. In addition to being linear in the state variables, it retains only terms
linear in J2 . The derivation is borrowed from part of the content in Butcher et
al.,20 but the notation has been adapted for consistency with the rest of this thesis.
Furthermore, the derivation was originally performed in a dimensionless manner,
but we will repeat the derivation in dimensional form for clarity and consistency with
Chapter 5. However, to discuss the generation of perturbation solutions, the resulting
ODE model will be put in a dimensionless form.
The J2 -perturbed two-body potential function is given in terms of the gravitational
parameter µ, J2 , the equatorial radius R, the position vector r, and the polar axis
unit vector K̂:22
J2 R2
  
µ 3  2
U (r) = 1− r · K̂ − 1 (4.1)
r 2r2 r2
Using r̈ = ∇U (r), we obtain the equations of motion for a satellite orbiting a body
with J2 , where êr = r/r:
3µJ2 R2
 2   
r  
r̈ = −µ 3 − 1 − 5 êr · K̂ êr + 2 êr · K̂ K̂ (4.2)
r 2r4
Note that in the absence of perturbations, the orbit radius can be expressed as r = ρa,
where ρ is given below in terms of the mean anomaly M = n (t − tp ), and tp is the
time of periapsis crossing:22
1 3
ρ(e, M ) = 1 − e cos M + e2 (1 − cos 2M ) + e3 (cos M − cos 3M ) + ... (4.3)
2 8
This equation remains instantaneously valid in terms of the osculating orbit elements
e(t) and M (t), which change due to the perturbation from J2 . This subject will be
revisited later in the model derivation.
44

Recall from Section 2.2.1 that the relative dynamical model is obtained by sub-
tracting the dynamical equations of a “chief” satellite from those of a “deputy” satel-
lite and linearizing: ∆r̈ ≈ r̈d − r̈c , where the resulting linearized equations are in
terms of ∆r = rd − rc , the position of the deputy relative to the chief:20

( 
µ   3µJ2 R2  2 
∆r̈ = − 3 I − 3r̂r̂ · ∆r − 1 − 5 K̂ · r̂ I
r 2r5
  2 
+ 2K̂K̂ + 5 7 K̂ · r̂ − 1 r̂r̂ (4.4)
)
  
− 10 K̂ · r̂ K̂r̂ + r̂K̂ · ∆r

Note this is in a dyadic form (cf. Chapter 2).


We wish to resolve these kinetic equations in matrix-vector form. To do this, first
note that êr · K̂ = sin θ sin i, where θ = ω + f is the argument of latitude and i is
the inclination. Also, we can resolve ∆r into its chief LVLH frame components as
∆r = xêr + yêt + zên . To resolve K̂ into its LVLH components, we use the 3-1-3
rotation matrix from the inertial frame to the LVLH frame, which is defined by the
chief orbit element angles Ω, i, θ:
 
cos Ω cos θ − sin Ω sin θ cos i sin Ω cos θ + cos Ω sin θ cos i sin θ sin i
RHN = − cos Ω sin θ − sin Ω cos θ cos i − sin Ω sin θ + cos Ω cos θ cos i cos θ sin i
sin Ω sin i − cos Ω sin i cos i
(4.5)
Thus H
K̂ = [RHN ] N K̂, where K̂ = [0, 0, 1]> :
N

 
sin θ sin i
H
K̂ =  cos θ sin i  (4.6)
cos i

The right side of equation 4.4 is now resolved in LVLH frame components and
2
expressed in a matrix-vector form, after collecting α = J2 Ra :
45

( 
2x
µ  −y 
∆r̈ =
r3
−z
6 − 9 sin2 i (1 − cos 2θ) 6 sin2 i sin 2θ
  
6 sin 2i sin θ x )
2
+ α 6 sin2 i sin 2θ − 23 + sin4 i (9 − 21 cos 2θ) − 32 sin 2i cos θ  y 
3 3 2 15 2 2 z
6 sin 2i sin θ − 2 sin 2i cos θ − 2 − 3 cos i + 2 sin θ sin i
(4.7)

We have now completed the kinetic equations for linearized relative motion sub-
ject to the J2 perturbation. Treating the kinematics involves resolving the inertial
acceleration of the relative position vector into the Hill frame, which rotates with
angular velocity ω H . This is done by applying the transport theorem twice as in
equation 2.26. The angular velocity is defined in a hybrid coordinate system in terms
of the rates of change of the orbit element angles:
di K̂ × ên
ω H = Ω̇K̂ + + θ̇ên (4.8)
dt kK̂ × ên k
The angular velocity can be expressed in its LVLH components by dotting into the
LVLH frame:
di
ωr = ω H · êr = Ω̇ sin θ sin i + cos θ (4.9)
dt
ωt = ω H · êt = 0 (4.10)
ωn = ω H · ên = θ̇ + Ω̇ cos i (4.11)
Note that equation 4.10 can be shown to yield the following constraint:
di cos θ sin i
= Ω̇ (4.12)
dt sin θ
In LVLH components, equation 2.26 becomes:
∆r̈ = ẍ − ω̇n y − 2ωn ẏ − ωn2 x + ωn ωr z êr


+ ÿ + ω̇n x + 2ωn ẋ − ωn2 + ωr2 y − ω̇r z − 2ωr ż êt


 
(4.13)
+ z̈ + ωn ωr x + ω̇r y + 2ωr ẏ − ωr2 z ên


We need the perturbed chief orbit element rates Ω̇, di/dt, and θ̇ to obtain the
LVLH frame angular velocity terms. The variational equations in their Gaussian
form yield the osculating rates for Ω, i, θ:
r sin θ
Ω̇ = NP
h sin i
rcosθ (4.14)
di/dt = NP
h
θ̇ =θ̇u + θ̇p = θ̇u − Ω̇ cos i
46

The perturbed argument of latitude rate θ̇ is the sum of two terms: θ̇u , representing
the rate of change of θ as measured from an unperturbed and unmoving ascending
node, and θ̇p , which appears due to the regression of the ascending node from which
θ is measured. These two terms of θ are sometimes referred to as its “unperturbed”
and “perturbed” components, respectively. By resolving the perturbing acceleration
as aP = RP êr + TP êt + NP ên , we can obtain NP referenced above. In this particular
case, the perturbing acceleration is due to J2 alone:

aJ2 = RJ2 êr + TJ2 êt + NJ2 ên (4.15)

3µJ2 R2 2 2

RJ2 = − 1 − 3 sin i sin θ
2r4
3µJ2 R2
2 sin2 i sin θ cos θ

TJ2 = − (4.16)
2r 4

3µJ2 R2
NJ2 = − (2 sin i cos i sin θ)
2r4
Substituting NJ2 , we obtain:

3µJ2 R2
Ω̇ = − cos i sin2 θ
hr3 (4.17)
3µJ2 R2
di/dt = − sin 2i sin 2θ
4hr3
The average J2 acceleration is given by:
Z 2π
1 3µJ2 R2
aJ2 dθ = − (1 + 3 cos 2i) êr = RJ2 êr (4.18)
2π 0 8r4

We obtain θ̇u by balancing the centripetal acceleration with the sum of the two-body
and average J2 acceleration, using a binomial expansion to retain only terms linear
in J2 , and simplifying:
s    2 !
1 R J 3 3 R
θ̇u = n2 − 2 ≈ nρ− 2 1 + J2 (1 + 3 cos 2i) (4.19)
ρ3 r 16 r

Now that the orbit element rates Ω̇, di/dt, and θ̇ have been obtained, we can define
the nonzero angular velocity terms appearing in equation 4.13. Using the definition
of θ̇ and the constraint 4.12, equations 4.9 and 4.11 can be rewritten:
sin i
ωr = Ω̇ (4.20)
sin θ
ωn = θ̇u (4.21)
47

To first order in J2 , assuming ȧ ≈ 0 and ė ≈ 0, it can be shown that the angular


accelerations are:  
sin i d   cos θ
ω̇r = n Ω̇ − Ω̇ (4.22)
sin θ dθ sin θ
d  
ω̇n = n θ̇u = 0 (4.23)

To first order, ωn ωr = nρ−3/2 Ω̇ sin
sin i
θ
, and the only other nonzero angular velocity
2
squared term is ωn :
  2 
2 2 3 3 R
ωn = n ρ 1 + J2 (1 + 3 cos 2i) (4.24)
8 r
4.1.1 Case of Vanishing Chief Eccentricity
If the eccentricity remains small enough to ignore, we substitute ρ = 1, e = 0, r = a,

and h = µa wherever they appear in the model equations. After applying these
2
substitutions, and collecting the dimensionless parameter α = J2 Ra , the nonzero
kinematic angular velocity and acceleration terms are presented below in their final
forms, retaining only terms linear in J2 .
3
ωr = − nα sin 2i sin θ
2 
3
ωn = n 1 + α (1 + 3 cos 2i)
16
3 2
ω̇r = − n α sin 2i cos θ (4.25)
2
3
ωn ωr = − n2 α sin 2i sin θ
2 
2 2 3
ωn = n 1 + α (1 + 3 cos 2i)
8
Substituting these equations into the kinematic equations 4.13, and setting the
result equal to the kinetic equations 4.7, we obtain the linearized model we sought
to derive. We then re-arrange the resulting model equation so that the O (α0 ) terms
(which are just the HCW equations) are factored out and equated with the time-
varying J2 perturbations. The result is given by equation 4.26.

3
ẍ − 2nẏ − 3n2 x
    
0 8 (1 + 3 cos 2i) 0 ẋ
 ÿ + 2nẋ  = αn − 3 (1 + 3 cos 2i) 0 −3 sin 2i sin θ  ẏ 
8
z̈ + n2 z 0 3 sin 2i sin θ 0 ż
2
6 sin2 i sin 2θ
3  15  
8 5 + 15 cos 2i + 24 cos 2θ sin i 2 sin 2i sin θ x
+ αn2  6 sin2 i sin 2θ − 21 2
4 sin i cos 2θ −3 sin 2i cos θ  y 
15 9 sin2 i z
2 sin 2i sin θ 0 − 2 + 4 (27 − 15 cos 2θ)
(4.26)
48

To numerically simulate the model as an LTV system, substitute θ = nt + θ0 into


equation 4.26, and use the initial values of n and i for the chief orbit.
The difference between the dimensional and non-dimensional forms is simple.
First, all variables with distance units are divided by the chief semimajor axis. The
normalized time is τ = nt, so the time derivative transformation is dtd ( ) = n dτd ( ),
where differentiation with respect to normalized time is represented by ( )0 . Applying
these changes to the model equations, we obtain the non-dimensional form:
x00 − 2y 0 − 3x
   3
 0 
0 8 (1 + 3 cos 2i) 0 x
 y 00 + 2x0  = α − 3 (1 + 3 cos 2i)
8 0 −3 sin 2i sin θ  y 0 
z 00 + z 0 3 sin 2i sin θ 0 z0
2
6 sin2 i sin 2θ
3  15  
8 5 + 15 cos 2i + 24 cos 2θ sin i 2 sin 2i sin θ x
+ α 6 sin2 i sin 2θ − 21 2
4 sin i cos 2θ −3 sin 2i cos θ  y 
15 9 sin2 i z
2 sin 2i sin θ 0 − 2 + 4 (27 − 15 cos 2θ)
(4.27)

This non-dimensional form is convenient for solving via straightforward perturbation


expansion.9 The solutions to these O (α) equations can be found efficiently using
symbolic software to evaluate the following:
   2 −1 
x2 (τ ) s − 3 −2s 0
 y2 (τ )  = L−1  2s s2 0  L (F1 (τ )) (4.28)
2
z2 (τ ) 0 0 s +1

where the inverted matrix is the transfer matrix of the HCW system and F1 (τ ) is
obtained by substituting the O (α0 ) (HCW) solution into the right side of equation
4.27. The resulting perturbation solution to the normalized equations is given as the
normalized HCW solution plus the J2 correction terms x (τ ), y (τ ), z (τ ) in equations
4.29 - 4.31 on the following pages.
49


1
x(τ ) = − α [(−96 − 288 cos(2i))x0 − (36 + 108 cos(2i))y 0 0 ]+
32
[(96 + 288 cos(2i))x0 + (6 + 18 cos(2i))τ x0 0 + (36 + 108 cos(2i))y 0 0 ] cos(τ )+
[(−180 + 180 cos(2i) + 264sin2 (i))x0 + (−36 + 36 cos(2i) + 24sin2 (i))τ x0 0 +
(−72 + 72 cos(2i) + 120sin2 (i))y 0 0 ] cos(τ − 2τn )+
[(192(1 − cos(2i)) − 152sin2 (i))x0 + (96(1 − cos(2i)) − 76sin2 (i))y 0 0 ] cos(2τ − 2τn )+
[(−27 + 27 cos(2i) − 6sin2 (i))x0 + (−18 + 18 cos(2i) − 4sin2 (i))y 0 0 ] cos(3τ − 2τn )+
[24 sin(2i)z 0 0 ] cos(τ − τn ) + [−8 sin(2i)z 0 0 ] cos(2τ − τn )+
[−24 sin(2i)z 0 0 ] cos(τn ) + [−168sin2 (i)x0 − 68sin2 (i)y 0 0 ] cos(2τn )+
[8 sin(2i)z 0 0 ] cos(τ + τn ) + [(15(1 − cos(2i)) + 62sin2 (i))x0 + (−6 + 6 cos(2i)+
28sin2 (i))y 0 0 ] cos(τ + 2τn )+
[(18 + 54 cos(2i))τ x0 − (6 + 18 cos(2i))x0 0 + (12τ + 36 cos(2i))τ y 0 0 ] sin(τ )+
[(108 − 108 cos(2i) − 72sin2 (i))τ x0 + 72sin2 (i)y0 + (72(1 − cos(2i))−
48sin2 (i))τ y 0 0 ] sin(τ − 2τn )+
[−48sin2 (i)τ x0 − 16sin2 (i)x0 0 + 8sin2 (i)y0 − 24sin2 (i)τ y 0 0 ] sin(2τ − 2τn )+
[(9 − 9 cos(2i) + 2sin2 (i))x0 0 ] sin(3τ − 2τn ) + [96 sin(2i)z0 ] sin(τ − τn )+
[8 sin(2i)z0 ] sin(2τ − τn ) + [216 sin(2i)z0 ] sin(τn )+
[−16sin2 (i)x0 0 + 168sin2 (i)y0 ] sin(2τn ) + [−112 sin(2i)z0 ] sin(τ + τn )+

2 0 2
[(9 − 9 cos(2i) + 2sin (i))x 0 − 88sin (i)y0 ] sin(τ + 2τn )
(4.29)
50

1
y(τ ) = α([(−114 − 342 cos(2i))x0 − (48 + 144 cos(2i))y 0 0 ]τ + [(−18 − 54 cos(2i))x0 0 ]+
16
[(−18 − 54 cos(2i))τ x0 + (18 + 54 cos(2i))x0 0 − (12 + 36 cos(2i))τ y 0 0 ] cos(τ )+
[(−108 + 108 cos(2i) + 72sin2 (i))τ x0 + (−36 + 36 cos(2i) + 60sin2 (i))x0 0 −
72sin2 (i)y0 + (−72 + 72 cos(2i) + 48sin2 (i))τ y 0 0 ] cos(τ − 2τn )+
[−102sin2 (i)τ x0 − 34sin2 (i)x0 0 + 17sin2 (i)y0 − 51sin2 (i)τ y 0 0 ] cos(2τ − 2τn )+
[(−3 + 3 cos(2i) + 14sin2 (i))x0 0 ] cos(3τ − 2τn ) + [−96 sin(2i)z0 ] cos(τ − τn )+
[8 sin(2i)z0 ] cos(2τ − τn ) + [−24 sin(2i)z0 − 48 sin(2i)τ z 0 0 ] cos(τn )+
[−126sin2 (i)τ x0 + (48 − 48 cos(2i) − 38sin2 (i))x0 0 − 33sin2 (i)y0 − 51sin2 (i)τ y 0 0 ] cos(2τn )+
[112 sin(2i)z0 ] cos(τ + τn ) + [(−9 + 9 cos(2i) − 2sin2 (i))x0 0 + 88sin2 (i)y0 ] cos(τ + 2τn )+
[(132 + 396 cos(2i))x0 + (6 + 18 cos(2i))τ x0 0 + (60 + 180 cos(2i))y 0 0 ] sin(τ )+
[(−72 + 72 cos(2i) + 84sin2 (i))x0 − (36 − 36 cos(2i) − 24sin2 (i))τ x0 0 ] sin(τ − 2τn )+
[(96 − 96 cos(2i) − 58sin2 (i))x0 + (48 − 48 cos(2i) − 29sin2 (i))y 0 0 ] sin(2τ − 2τn )+
[(−9 + 9 cos(2i) + 42sin2 (i))x0 − (6 − 6 cos(2i) − 28sin2 (i))y 0 0 ] sin(3τ − 2τn )+
[24 sin(2i)z 0 0 ] sin(τ − τn ) + [192 sin(2i)τ z0 + 24 sin(2i)z 0 0 ] sin(τn )+
[6sin2 (i)x0 − 12sin2 (i)τ x0 0 + 126sin2 (i)τ y0 + (48 − 48 cos(2i) − 29sin2 (i))y 0 0 ] sin(2τn )+
[8 sin(2i)z 0 0 ] sin(τ + τn ) + [8 sin(2i)z 0 0 ] sin(2τ − τn )
[(15 − 15 cos(2i) + 62sin2 (i))x0 − (6 − 6 cos(2i) − 28sin2 (i))y 0 0 ] sin(τ + 2τn ))
(4.30)

1
z(τ ) = α([(144 − 216sin2 (i))τ z00 ] cos(τ )
64
+ [−60sin2 (i)z0 − 60sin2 (i)τ z00 ] cos(τ − 2τn ) + [15sin2 (i)z0 ] cos(3τ − 2τn )
+ [−384 sin(2i)τ x0 − 48 sin(2i)x00 − 192 sin(2i)τ y00 ] cos(τ − τn )
+ [16 sin(2i)x00 ] cos(2τ − τn ) + [48 sin(2i)x00 ] cos(τn ) − [16 sin(2i)x00 ] cos(τ + τn )
+ [45sin2 (i)z0 ] cos(τ + 2τn ) + [15sin2 (i)z00 ] sin(3τ − 2τn ) + [15sin2 (i)z00 ] sin(τ + 2τn ))
+ [(−144 + 216sin2 (i))τ z0 − (144 − 216sin2 (i))z00 ] sin(τ ) − [60sin2 (i)τ z0 ] sin(τ − 2τn )
+ [192 sin(2i)x0 + 96 sin(2i)y00 ] sin(τ − τn ) + [48 sin(2i)x0 + 32 sin(2i)y00 ] sin(2τ − τn )
+ [144 sin(2i)x0 + 96 sin(2i)y00 ] sin(τn ) + [96 sin(2i)x0 + 32 sin(2i)y00 ] sin(τ + τn )
(4.31)

4.2 Model Analysis and Validation


The errors in the HCW and J2 models were estimated by comparing their results to
the results of a truth model, which separately simulated the motions of both a chief
51

and deputy spacecraft around the Earth with J2 , then subtracted the chief state from
the deputy, and rotated the result into the moving LVLH frame. We simulate a chief
and deputy spacecraft in LEO, with the chief orbit elements and deputy orbit element
differences given by Table 4.1. The resulting relative orbital motion is in a perturbed
ellipse, with an average separation between the chief and deputy of 54.6 km. The
out-of-plane relative motion range is ±6 km, while the in-plane relative motion range
is ±35 km radial and ±70 km along-track.

Table 4.1. Orbital Elements for LEO Relative Orbit Scenario with J2
a (km) e i (◦ ) Ω (◦ ) θ (◦ )
œc 7100 0.0 2.0 0.0 0.0
δœ 0 0.005 0.05 0.0 0.0

The relative orbit for this case is plotted for 4 orbits in Fig. 4.1.

-2

-4

-6

-40
-20
0
20 0 20 40 60 80
40 -80 -60 -40 -20

Figure 4.1. Relative Orbit for 4 Orbits (LEO with J2 )

Note the relative orbit plane rotation due to J2 . Fig. 4.2 shows that the osculating
eccentricity e(t) exceeds the value of α for much of the orbits. This will affect model
accuracy, since the J2 model was derived assuming α >> e. Despite this, Figs. 4.3-4.6
show that the J2 model has lower error than the HCW model as expected. It is clear
that the along-track error constitutes the majority of the error, because both lin-
earized models drift in the y-direction. The along-track and cross-track performance
is consistently improved with this model, but the radial error is comparable. If an
52

10-3
3

2.5

1.5

0.5

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 4.2. Small Parameter Values for Two Orbits (LEO with J2 )

orbit could be found for which the osculating eccentricity did not oscillate to con-
siderable values, the improvements of including J2 terms would become much more
obvious. Note that the difference between the J2 ODE model and its perturbation
solution is so small that the results from the perturbation solution are omitted in the
error plots.
In Fig. 4.7, the chief orbit radius was raised from 7,100 km to a radius of over
20,000 km, but the relative orbit dimensions are kept the same by linearly re-scaling
the deputy orbit element differences δœ in Table 4.1. As the radius of the chief orbit
is increased, the maximum of the error norm over two orbits decreases for both the
HCW solution and the linearized J2 model. The errors from the two models begin to
converge as the radius becomes larger, due to the reduced effect of J2 .
This J2 model clearly improves over the HCW solution, but the effects of Earth’s J2
perturbation are somewhat small for the relative motion. However, for some bodies
that deviate more dramatically from a spherical shape, the effects on the relative
motion can be much more pronounced, and inclusion of these terms in the relative
motion models becomes a necessity even for short time scales.
53

1400 7000

1200 6000

1000 5000

800 4000

600 3000

400 2000

200 1000

0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 4.3. Radial Error (LEO) Figure 4.4. Along-Track Error (LEO)

350 7000

300 6000

250 5000

200 4000

150 3000

100 2000

50 1000

0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 4.5. Cross-Track Error (LEO) Figure 4.6. Error Norm (LEO)
54

Figure 4.7. Maximum Error Norm over Two Orbits vs. Chief Orbit Radius
55

5. A Linearized Relative Orbit Model for


Orbits in a Rotating Second Degree and Order
Gravity Field

5.1 Formulation

Many of the asteroids and other small bodies in the solar system are of a shape that
can be approximated by a triaxial ellipsoid. Furthermore, most of the small bodies in
the solar system undergo stable rotation about their axis of maximum inertia.24 Here,
a dynamical model will be derived for the relative orbital motion in the proximity of a
stably rotating ellipsoidal body. The efficacy of the model is restricted to near-circular
orbits, but it will be particularly applicable for modeling relative motion dynamics
in desirable “science orbits”, which maintain consistent distance from the primary
body, and whose form conveniently persists for long time spans. Note that since
this is a completely new model, references to literature for some of these derivations
are limited or unavailable, although the overall derivation bears some procedural
similarity to Chapter 4.
Recall from Chapter 3 the formula for the gravitational potential of an arbitrary
body, given in terms of the spherical harmonic series, is:
∞ l  l
µXX R
U (r, φ, λ) = Plm (sin φ)[Clm cos mλ + Slm sin mλ] (5.1)
r l=0 m=0 r

The potential is given in terms of the spherical coordinates r (range), φ (latitude), λ


(longitude), where R is the radius of the Brillouin sphere and µ is the gravitational
parameter. Plm are the associated Legendre functions, and Clm and Slm are the
gravity field harmonic coefficients, which are defined by the mass distribution of the
body.
For a perfect ellipsoidal body, it can be shown that the gravitational potential
reduces to a simpler form, with all Slm = 0, and Clm nonzero only for even values of
56

l and m:24
∞ l  l
µXX R
U (r, φ, λ) = Plm (sin φ)[Clm cos mλ] (5.2)
r l=0 m=0 r
Furthermore, the remaining gravitational coefficients can be explicitly defined in
terms of the axes A ≥ B ≥ C:29, 30
m
  2l − m2 −2i
int( l−m ) 2 2 +2i 2 A 2 +B 2
l X (A − B ) 2
C − 2
 4
3 (2 − δ0m ) 2
! (l − m)!
Clm =
16i 2l − m2 − 2i ! m2 + i !i!
 
Rl 2m (l + 3)(l + 1)! i=0
(5.3)

The two second-order gravitational coefficients (the body “oblateness” and “ellip-
ticity” terms, respectively) are given in terms of the body geometry:

A2 + B 2
 
1 2
C20 = C − (5.4)
5R2 2
1 2 2

C22 = A − B (5.5)
20R2
Substituting these equations into 5.3 and simplifying, we obtain a simpler and more
interesting expression for all of the higher-order gravitational coefficients:
int( l−m ) l
− m +2i) m +2i
2 ( 2
l
5 2 2l ! (l − m)! X4

C20 C222
Clm = 3 (2 − δ0m ) l
(5.6)
− m2 − 2i ! m2 + i !i!
 
(l + 3)(l + 1)! i=0 2

The implications of 5.6 are that any lth -order gravitational coefficients will be
explicit polynomial functions of C20 and C22 , with all terms being of order l/2. For
example, consider the following 4th and 6th order gravitational coefficients:

15 2 2

C40 = C20 + 2C22 (5.7)
7
5
C42 = C20 C22 (5.8)
7
5 2
C44 = C22 (5.9)
28
125 3 2

C60 = C20 + 6C20 C22 (5.10)
21
57

From 5.2, it is clear that in the gravitational potential function, for increasing
l
values of l, the Clm Rr terms will become increasingly small, since r > R and all
|Clm | << 1. Since |C20 | << 1 and |C22 | << 1, from 5.6, all higher-order gravitational
coefficients will be smaller still.
We truncate the gravitational potential series by keeping only up to the second-
order gravitational coefficients, recognizing that all higher-order gravitational coeffi-
cients can be neglected as long as C20 and C22 are sufficiently small:
 2  2 !
µ R R
U (r, φ, λ) ≈ 1 + C20 P20 (sin φ) + C22 P22 (sin φ) cos 2λ (5.11)
r r r
Note that C00 = 1. This approximation is most valid for r >> R and/or for an
ellipsoidal body with very small C20 and C22 . Note that by keeping both of the second-
order gravitational coefficients, we are assuming that |C20 | ≈ |C22 |. This assumption
is consistent throughout the model derivation, but if one of these coefficients is much
smaller than the other, the model can still be applied by eliminating the smallest of
the two coefficients in the final model equations.
By substituting the definitions of the two associated Legendre functions and sim-
plifying, we can explicitly express the potential in terms of the spherical coordinates:
 2    2 !
µ R 3 R
U (r, φ, λ) ≈ 1 + C20 1 − cos2 φ + 3C22 (cos2 φ) cos 2λ
r r 2 r
(5.12)
Using the definitions of the angles of the spherical coordinate system and r̈ = ∇U (r),
we can obtain the following approximation of the equation of motion of an orbiting
satellite, subject only to the main body’s gravitational field:

r̈ ≈ aC00 + aC20 + aC22 (5.13)

The gravitational acceleration terms aC00 , aC20 , and aC22 are functions of the relative
configuration of the satellite orbital position and the orientation of the rotating pri-
mary body, presented in Fig. 5.1 (adapted in part from Wikipedia). These terms will
be explicitly defined shortly.
58

Figure 5.1. Problem Geometry for Orbiting a Rotating Ellipsoidal Body

The right side of Fig. 5.1 shows the position of an orbiting body in a spherical
coordinate system centered on the primary body center of mass. The primary body
is an ellipsoid, and its principal axes are aligned with vectors Î, Ĵ, K̂, which are also
parallel with the axes A, B, C, respectively. The primary body rotates about its
axis of maximum inertia K̂, such that the Î and Ĵ vectors rotate in the “plane of
reference” defined in the left side of the figure. Both the K̂ vector and the reference
direction vector γ (which lies in the plane of reference) remain fixed with respect to
the stars. As the primary body rotates (clockwise in this figure), the angle ψ between
Î and γ increases at a constant rate, ψ = ψ0 + ct. The orbit parameterization also
makes use of the reference direction vector, using it to define the longitude of the
ascending node Ω. The remaining two angles are the inclination i and argument of
latitude θ = ω + f . Finally, the orbiter-fixed LVLH frame is defined in the usual way
by the orthogonal vectors êr , êt , ên .
The gravitational acceleration terms are now defined in terms of the problem
geometry:
r
aC00 = −µ (5.14)
r3
59

3µC20 R2
  2    
aC20 = 1 − 5 êr · K̂ êr + 2 êr · K̂ K̂ (5.15)
2r4
3µC22 R2
  2  2      
aC22 = −5 êr · Î − êr · Ĵ êr + 2 êr · Î Î − 2 êr · Ĵ Ĵ (5.16)
r4
Note the similarity between 5.15 and 5.16. Specifically, we can rewrite 5.16 in terms
of two components that are each structurally identical to 5.15, differing only by a
factor of two:
aC22 = aC22 ,I − aC22 ,J (5.17)
3µC22 R2
  2    
aC22 ,I = 1 − 5 êr · Î êr + 2 êr · Î Î (5.18)
r4
3µC22 R2
  2    
aC22 ,J = 1 − 5 êr · Ĵ êr + 2 êr · Ĵ Ĵ (5.19)
r4
Comparison of 5.15 and 5.17 - 5.19 may provide the reader with some insight of the
similarities between the oblateness and ellipticity-driven gravitational acceleration
terms. Note that if non-Keplerian perturbations are ignored, the orbit radius can be
expressed as r = ρa, where ρ is given in equation 4.3 in terms of the mean anomaly
M = n (t − tp ), and tp is the time of periapsis crossing. This equation remains
instantaneously valid in terms of the osculating orbit elements e(t) and M (t), which
change due to the non-Keplerian perturbation accelerations aC20 and aC22 . This
subject will be revisited later in the model derivation.
Equations 5.13 - 5.16 describe the approximate orbital dynamics for an orbiting
satellite subject only to the main body’s gravitational field. However, we are inter-
ested in the relative dynamics for two such orbiting satellites. Recall from Section
2.2 that the relative dynamical model is fundamentally obtained by subtracting the
dynamical equations of a “chief” satellite from those of a “deputy” satellite. The
resulting dynamical equations will be in terms of ∆r = rd − rc , the position of the
deputy relative to the chief:

∆r̈ ≈ ∆r̈C00 + ∆r̈C20 + ∆r̈C22 (5.20)


60

For example, ∆r̈C00 = aC00 (d) − aC00 (c) is the difference in two-body gravity felt by the
deputy and chief.
To derive the linearized dynamical model, we will retain only the terms linear in
∆r, and the equation will also be put in a dyadic form. This process was demonstrated
for the two-body acceleration term in Section 2.2. The result is repeated here, along
with the other gravitational acceleration terms in 5.20:
µ  
∆r̈C00 = − I − 3r̂r̂ · ∆r (5.21)
r3
( 
3µC20 R2 2 
   2 
∆r̈C20 = 1 − 5 K̂ · r̂ I + 2K̂K̂ + 5 7 K̂ · r̂ − 1 r̂r̂
2r5
) (5.22)
  
− 10 K̂ · r̂ K̂r̂ + r̂K̂ · ∆r

∆r̈C22 = ∆r̈C22 ,I − ∆r̈C22 ,J (5.23)


( 
3µC22 R2 2 
   2 
∆r̈C22 ,I = 1 − 5 Î · r̂ I + 2ÎÎ + 5 7 Î · r̂ − 1 r̂r̂
r5
) (5.24)
  
− 10 Î · r̂ Îr̂ + r̂Î · ∆r
( 
3µC22 R2 2 
   2 
∆r̈C22 ,J = 1 − 5 Ĵ · r̂ I + 2ĴĴ + 5 7 Ĵ · r̂ − 1 r̂r̂
r5
) (5.25)
  
− 10 Ĵ · r̂ Ĵr̂ + r̂Ĵ · ∆r

It is worth noting that even after the deputy-chief subtraction operation and subse-
quent linearization, the property described in 5.17 - 5.19 is still preserved in 5.23 -
5.25. It is also restated that each of these equations is a linear approximation in ∆r,
so they are no longer the exact differences in the acceleration terms acting on the
deputy and chief. From this point on, we will only be dealing with these linearized
equations for the acceleration differences.
We must resolve 5.21, 5.22, and 5.23 into the chief LVLH frame as ∆r = xêr +
yêt + zên . As in Chapter 4, this is done by expressing each of the necessary unit
61

vectors in their LVLH (or equivalently, R-T-N) components. The most challenging
of these are Î and Ĵ. In addition to the orbital element angles, we must account
for the rotation angle ψ of the primary body in obtaining them. If we define the
inertial frame right-hand coordinate system axes with the orthogonal unit vectors
n   o
γ̂, − γ̂ × K̂ /kγ̂ × K̂k, K̂ , then Î and Ĵ resolved into their components along
these axes will be N
Î = [cos ψ, sin ψ, 0]> and N
Ĵ = [− sin ψ, cos ψ, 0]> . To obtain
these unit vectors in the chief LVLH components, we make use of the 3-1-3 rotation
matrix defined by the chief orbit element angles Ω, i, θ:
 
cos Ω cos θ − sin Ω sin θ cos i sin Ω cos θ + cos Ω sin θ cos i sin θ sin i
RHN = − cos Ω sin θ − sin Ω cos θ cos i − sin Ω sin θ + cos Ω cos θ cos i cos θ sin i
sin Ω sin i − cos Ω sin i cos i
(5.26)
H
Finally, using Î = [RHN ] N Î, etc., and noting that H
K̂ is unchanged from Chapter
4, we obtain the following:
 
cos ψ (cos Ω cos θ − sin Ω sin θ cos i) + sin ψ (sin Ω cos θ + cos Ω sin θ cos i)
H
Î =  − cos ψ (cos Ω sin θ + sin Ω cos θ cos i) + sin ψ (− sin Ω sin θ + cos Ω cos θ cos i) 
cos ψ sin Ω sin i − sin ψ cos Ω sin i
 (5.27)

− sin ψ (cos Ω cos θ − sin Ω sin θ cos i) + cos ψ (sin Ω cos θ + cos Ω sin θ cos i)
H
Ĵ =  sin ψ (cos Ω sin θ + sin Ω cos θ cos i) + cos ψ (− sin Ω sin θ + cos Ω cos θ cos i) 
− sin ψ sin Ω sin i − cos ψ cos Ω sin i
  (5.28)
sin θ sin i
H
K̂ =  cos θ sin i  (5.29)
cos i
Note, for example, that êr · K̂ = sin θ sin i. Also, trivially, r̂ = êr .
Substituting ∆r = xêr + yêt + zên and the previous equations for the unit vectors
in terms of the angles into 5.21, 5.22, and 5.23, simplifying, and expressing the results
in matrix-vector form, we obtain the following.
 
2x
µ
∆r̈C00 = 3  −y  (5.30)
r
−z
62

  
O O12 O13 x
µC20 R  11
2
∆r̈C20 = O21 O22 O23   y  (5.31)
r5
O31 O32 O33 z
  
E E12 E13 x
3µC22 R2  11
∆r̈C22 = E21 E22 E23   y  (5.32)
r5
E31 E32 E33 z
Note that Oij = Oji and Eij = Eji , with all unique elements defined in Tables 5.1
and 5.2, respectively.

Table 5.1. Unique O Matrix Terms in 5.31, Oij = Oji

O11 = −6 + 9 sin2 i (1 − cos 2θ) O12 = −6 sin2 i sin 2θ


2
O13 = −6 sin 2i sin θ O22 = 32 − sin4 i (9 − 21 cos 2θ)
O23 = 32 sin 2i cos θ O33 = 32 + 3 cos2 i − 15
2
sin2 θ sin2 i

Table 5.2. Unique E Matrix Terms in 5.32, Eij = Eji

E11 = 3 cos (2 (Ω − ψ)) (3 + cos 2i) cos 2θ + 2 sin2 i − 12 sin (2 (Ω − ψ)) cos i sin 2θ


E12 = 8 sin (2 (Ω − ψ)) cos i cos 2θ + 2 cos (2 (Ω − ψ)) (3 + cos 2i) sin 2θ
E13 = −8 sin (2 (Ω − ψ)) sin i cos θ − 4 cos (2 (Ω − ψ)) sin 2i sin θ
E22 = − 4 cos (2 (Ω − ψ)) 7 (3 + cos 2i) cos 2θ + 6 sin2 i + 7 sin (2 (Ω − ψ)) cos i sin 2θ
1


E23 = cos (2 (Ω − ψ)) sin 2i cos θ − 2 sin (2 (Ω − ψ)) sin i sin θ


E33 = − 14 cos (2 (Ω − ψ)) 5 (3 + cos 2i) cos 2θ + 18 sin2 i + 5 sin (2 (Ω − ψ)) cos i sin 2θ


Note from Table 5.2 that every term in the E matrix in 5.32 is premultiplied
by cos (2 (Ω − ψ)) or sin (2 (Ω − ψ)), which reflects the fact that ∆r̈C22 is completely
dependent on the relative configuration of the primary body attitude and the satellite
orbits.
Now that we have linearized the right side of 5.20 and expressed the results in
LVLH components x, y, z, we must address the left side of the equation. Specifi-
cally, we must resolve the inertial acceleration of the relative position vector into the
63

Hill frame, which rotates with angular velocity ω H . Repeating the procedure from
Chapter 4, we start by applying the transport theorem twice:
H 2 H
d ∆r d∆r
∆r̈ = 2
+ ω̇ H × ∆r + 2ω H × + ω H × (ω H × ∆r) (5.33)
dt dt
The angular velocity is defined again in a hybrid coordinate system in terms of the
rates of change of the orbit element angles:
di K̂ × ên
ω H = Ω̇K̂ + + θ̇ên (5.34)
dt kK̂ × ên k
The angular velocity can broken into its LVLH components as:
di
ωr = ω H · êr = Ω̇ sin θ sin i + cos θ (5.35)
dt
ωt = ω H · êt = 0 (5.36)

ωn = ω H · ên = θ̇ + Ω̇ cos i (5.37)

Note that equation 5.36 can be shown to yield the following constraint:
di cos θ sin i
= Ω̇ (5.38)
dt sin θ
In LVLH components, equation 5.33 becomes:
∆r̈ = ẍ − ω̇n y − 2ωn ẏ − ωn2 x + ωn ωr z êr


+ ÿ + ω̇n x + 2ωn ẋ − ωn2 + ωr2 y − ω̇r z − 2ωr ż êt


 
(5.39)

+ z̈ + ωn ωr x + ω̇r y + 2ωr ẏ − ωr2 z ên




Note that, for example, ω̇n = dωn /dt. To use equation 5.39 and complete the model
derivation, we need to first find the perturbed chief orbit element rates Ω̇, di/dt,
and θ̇ to explicitly find the LVLH frame angular velocity terms. The procedure here
is analogous to what was discussed in Chapter 4, but more detailed. First, any
perturbing acceleration aP can be resolved into its chief LVLH radial, transverse,
and normal components as aP = RP êr + TP êt + NP ên . Furthermore, we repeat the
following useful definitions for the orbital angular momentum:
p
h = r × v = hên = µa (1 − e2 )ên (5.40)
64

ḣ = r × aP = rTP ên − rNP êt (5.41)

Below, we repeat from Prussing and Conway18 the equations for Ω̇ and di/dt.
The derivation of θ̇ is the most complex part of the kinematics, so that discussion is
postponed until later. Finally, note that the inertial coordinate system unit vectors
 
are named for notational convenience: êX = γ̂, êY = − γ̂ × K̂ /kγ̂ × K̂k, êZ = K̂.

hX ḣY − hY ḣX
Ω̇ = (5.42)
h2X + h2Y

di ḣhZ − hḣZ
= 1/2
(5.43)
dt h (h2 − h2Z )
In this notation, note that, for example, hX = h · êX = hên · êX .
Since we are only including up to the second-order gravitational terms, the per-
turbing acceleration is aP = aC20 + aC22 , thus RP = RC20 + RC22 , etc. Using this
definition for the perturbing acceleration, and equations 5.40 and 5.41, we expand
equations 5.42 and 5.43:

r
Ω̇ = 2 hX [(TC20 + TC22 ) (ên · êY ) − (NC20 + NC22 ) (êt · êY )]
hX + h2Y
 (5.44)
− hY [(TC20 + TC22 ) (ên · êX ) − (NC20 + NC22 ) (êt · êX )]

di r {(TC20 + TC22 ) hZ − h [(TC20 + TC22 ) (ên · êZ ) − (NC20 + NC22 ) (êt · êZ )]}
= 1/2
dt h (h2 − h2Z )
(5.45)
Note that the dot-product terms can be expressed purely as functions of the orbit
element angles. For example, ên · êX = sin Ω sin i and ên · êY = − cos Ω sin i. Now, we
need the equations for the R-T-N components of aC20 and aC22 . These are obtained
by inserting equations 5.27 - 5.29 into 5.15 and 5.16, in the same manner as 5.31 and
5.32 were obtained. The results are given below, after much simplification:
3µC20 R2
1 − 3 sin2 i sin2 θ

RC20 = 4
(5.46)
2r
3µC20 R2
2 sin2 i sin θ cos θ

TC20 = 4
(5.47)
2r
65

3µC20 R2
NC20 = (2 sin i cos i sin θ) (5.48)
2r4
3µC22 R2 
RC22 = 3 sin (2 (Ω − ψ)) cos i sin 2θ
r4 (5.49)
3 
− cos (2 (Ω − ψ)) 1 + 3 cos 2θ − 2 cos 2i sin2 θ
4
3µC22 R2 
TC22 = − 2 sin (2 (Ω − ψ)) cos i cos 2θ
r4 (5.50)
1 
− cos (2 (Ω − ψ)) (3 + cos 2i) sin 2θ
2
2
3µC22 R 
NC22 = 2 sin (2 (Ω − ψ)) sin i cos θ + cos (2 (Ω − ψ)) sin 2i sin θ (5.51)
r4
Substituting these definitions into 5.44 and 5.45 and simplifying significantly, we
obtain the following:

3µR2 2

Ω̇ = C 22 sin (2 (Ω − ψ)) sin 2θ + [C 20 + 2C 22 cos (2 (Ω − ψ))] cos i sin θ
hr3
(5.52) 
3µR2

di 1
= 2C22 sin (2 (Ω − ψ)) cos2 θ sin i + [C20 + 2C22 cos (2 (Ω − ψ))] sin 2θ sin 2i
dt hr3 4
(5.53)
It can be shown that these equations satisfy the constraint 5.38. Also note that, as
we would expect, if we let C22 = 0, we recover Ω̇ and di/dt due to C20 = −J2 alone,
which were given in equation 4.17.
It is possible to average equations 5.46 - 5.51 over the course of one orbit, and the
averaged radial components are necessary for the θ̇ derivation. The orbit-averaging
procedure for a general perturbing acceleration is:
Z 2π
1
aP = (RP êr + TP êt + NP ên ) dθ = RP êr + T P êt + N P ên (5.54)
2π 0

Averaging of 5.46 - 5.48 is unchanged from Chapter 4 and the results are repeated
below:
3µC20 R2
RC20 = (1 + 3 cos 2i) (5.55)
8r4
T C20 = 0 (5.56)
66

N C20 = 0 (5.57)

To average equations 5.49 - 5.51 over an orbit, we make use of the stable rotation
assumption and the circular orbit assumption to express both ψ and θ as linear
functions of time:
ψ = ψ0 + ct (5.58)

θ = θ0 + nt (5.59)

This allows for the linear transformation which will enable re-expression of equations
5.49 - 5.51 (functions of ψ and θ) as explicit functions of θ alone:
c
ψ = ψ0 + (θ − θ0 ) (5.60)
n
The averaged forms of 5.49 - 5.51 can thus be obtained after using this transformation,
and substituting Γ = c/n:
9µC22 R2 2 (Γ2 + Γ cos i) − sin2 i
 

RC22 = − sin (2πΓ) cos 2 (Ω − ψ 0 + Γ (θ0 − π))
4r4 Γ(Γ − 1)(Γ + 1)π
(5.61)
2
 
3µC22 R 4Γ cos i + cos 2i + 3 
T C22 = − sin (2πΓ) sin 2 (Ω − ψ 0 + Γ (θ0 − π))
4r4 (Γ − 1)(Γ + 1)π
(5.62)
2
 
6µC22 R (2Γ + cos i) sin i 
N C22 = − sin (2πΓ) sin 2 (Ω − ψ 0 + Γ (θ0 − π))
r4 (1 − 4Γ2 ) π
(5.63)
The angular velocity ratio Γ = c/n defines the ratio of the primary body rotation
rate to the orbital mean motion. It is clear that it is an important parameter. Note
that the 0/0 singularities that occur in RC22 and T C22 for Γ = 1 can be cleared by
applying L’Hospital’s rule.
At this point, we have obtained many terms (both dynamic and kinematic) that
are functions of C20 and/or C22 , but they are all linear in these terms, since they
were analytically obtained from the truncated gravitational potential. The remaining
kinematic terms are all obtained in ways that can introduce terms that are higher-
order in C20 or C22 , but such higher-order terms must be neglected. For example,
67

2
inclusion of O (C20 ) terms would only be valid if the truncated gravitational potential
included the terms associated with the fourth-order gravitational coefficients C40 , C42 ,
and C44 .
Now, we will discuss the derivation of θ̇ due to the included gravitational effects.
As in Chapter 4, θ̇ can be broken into an “unperturbed” component and a “perturbed”
component:
θ̇ = θ̇u + θ̇p (5.64)

Furthermore, θ̇p = −Ω̇ cos i, thus:

θ̇ = θ̇u − Ω̇ cos i (5.65)

There are two methods for obtaining θ̇u . The first is nearly identical to the method
used in Chapter 4. We start by balancing the centripetal acceleration with the sum
of the two-body and average C20 radial acceleration, but now include the average C22
radial acceleration term:
s  
ρ3
 
2
1 1  − 23

θ̇u = n − RC20 + RC22 ≈ nρ 1− RC20 + RC22 (5.66)
ρ3 r 2rn2
Note that the method of balancing the centripetal acceleration assumes that the chief
orbit is very nearly circular.
By substituting equations 5.55 and 5.61 and simplifying, we obtain:
  2
− 23 3 R h
θ̇u =nρ 1− C20 (1 + 3 cos 2i)
16 r
2 (Γ2 + Γ cos i) − sin2 i
  i
− 6C22 sin (2πΓ) cos (2 (Ω − ψ0 + Γ (θ0 − π)))
Γ(Γ − 1)(Γ + 1)π
(5.67)
Note that n2 = µ/a3 and equation 4.3 are used to obtain this result.
An alternative approach is potentially beneficial by allowing us to account for the
variation of the radial acceleration (and thus θ̇u ) with θ by modifying equation 5.66:
s  
ρ3
 
2
1 1 − 32
θ̇u = n − (RC20 (θ) + RC22 (θ)) ≈ nρ 1− (RC20 (θ) + RC22 (θ))
ρ3 r 2rn2
(5.68)
68

By substituting equations 5.46 and 5.49 and simplifying, we obtain:


  2
− 23 3 R h 
C20 1 − 3 sin2 i sin2 θ + 6C22 sin (2 (Ω − ψ)) cos i sin 2θ

θ̇u =nρ 1−
4 r

1 2
 i
− cos (2 (Ω − ψ)) 1 + 3 cos 2θ − 2 cos 2i sin θ
4
(5.69)

It can be shown that applying the linear transformation 5.60 and averaging equation
5.69 over θ actually yields equation 5.67. Using equations 5.65, 5.52, and 5.67 or 5.69,
we can define θ̇.
Now that the orbit element rates Ω̇, di/dt, and θ̇ have been obtained, we can
define the nonzero angular velocity terms appearing in equation 5.39. Note that
using equations 5.38 and 5.65, equations 5.35 and 5.37 can be expressed in simpler
forms:
sin i
ωr = Ω̇ (5.70)
sin θ
ωn = θ̇u (5.71)

Before obtaining ω̇r and ω̇n , note that the time derivative of any general parameter
S (a, e, Ω, i, θ) can be expressed as:

d d d d d
Ṡ (a, e, Ω, i, θ) = (S) ȧ + (S) ė + (S) Ω̇ + (S) i̇ + (S) θ̇ (5.72)
da de dΩ di dθ

This analysis is restricted to orbits in which the semimajor axis a and eccentricity
e remain nearly constant such that we can assume ȧ ≈ 0 and ė ≈ 0. Thus, we
differentiate ωr and ωn , noting that not all derivative terms appear due to their order
in C20 and C22 (for example, Ω̇i̇ ≈ 0) or our restriction ȧ ≈ ė ≈ 0:
 
sin i sin i cos θ sin i d   cos θ
ω̇r = Ω̈ − Ω̇θ̇ = Ω̇ − Ω̇ θ̇ (5.73)
sin θ sin2 θ sin θ dθ sin θ

d   d  
ω̇n = θ̇u = θ̇u θ̇ (5.74)
dt dθ
69

Furthermore, n is the only zeroth-order term in θ̇. Thus, to first order in C20 and
C22 , these angular accelerations are:
 
sin i d   cos θ
ω̇r = n Ω̇ − Ω̇ (5.75)
sin θ dθ sin θ

d  
ω̇n = n θ̇u = 0 (5.76)

where the linear transformation 5.60 must be used before taking the derivative with
respect to θ. To first order, ωn ωr = nρ−3/2 Ω̇ sin
sin i
θ
, and the only other nonzero angular
velocity squared term is ωn2 , whose value will vary depending on which θ̇u is used. The
first equation uses the averaged θ̇u , while the second equation uses the unaveraged
θ̇u :
  2
3 R h
ω 2n 2 3
=n ρ 1 − C20 (1 + 3 cos 2i)
8 r
2 (Γ2 + Γ cos i) − sin2 i
  i
− 6C22 sin 2πΓ cos (2 (Ω − ψ0 + Γ (θ0 − π)))
Γ(Γ − 1)(Γ + 1)π
(5.77)
 2 h

R 3 
ωn2 2 3 2 2

=n ρ 1−C20 1 − 3 sin i sin θ + 6C22 sin (2 (Ω − ψ)) cos i sin 2θ
r 2

1 2
 i
− cos (2 (Ω − ψ)) 1 + 3 cos 2θ − 2 cos 2i sin θ
4
(5.78)
By using equations 5.30 - 5.32 and 5.39 with the necessary kinematic equations,
we have everything we need to fully define the linearized form of equation 5.20 and
complete the model. Only one issue remains. The finalized model will vary depending
on the treatment of the eccentricity. In the following discussion, we assume that the
eccentricity remains small enough to ignore, and present the finalized model for that
case explicitly. Then, we will discuss a more rigorous treatment allowing for slightly
eccentric chief orbits, while the eccentricity is assumed to remain nearly constant and
 2 
O (ep ) = O C20 Ra , where p is an integer.
70

5.1.1 Case of Vanishing Chief Eccentricity

There will be many instances when the eccentricity will remain small enough that it
can be treated as zero. There are analytical and numerical techniques for finding such

orbits. For such cases, we substitute ρ = 1, e = 0, r = a, and h = µa wherever they
appear in the model equations. After applying these substitutions, the kinematic
angular velocity terms are presented below in their final forms, retaining all terms
linear in C20 and C22 . Note that the normal component of the angular velocity can be
obtained using either the averaged or the unaveraged θ̇u . Both forms are presented,
with the terms using the averaged θ̇u marked with a bar:
 2
R    
ωr = 3n sin i 2C22 sin (2 (Ω − ψ)) cos θ+ C20 +2C22 cos (2 (Ω − ψ)) cos i sin θ
a
(5.79)
 2
R 
ω̇r = 3n2
 
sin i C20 cos i + 2C22 (cos i − 2Γ) cos (2 (Ω − ψ)) cos θ
a (5.80)
  
+ 2C22 (2Γ cos i − 1) sin (2 (Ω − ψ)) sin θ
 2
2 R    
ωn ωr = 3n sin i 2C22 sin (2 (Ω − ψ)) cos θ+ C20 +2C22 cos (2 (Ω − ψ)) cos i sin θ
a
(5.81)
  2 h
3 R
ωn = n 1 − C20 (1 + 3 cos 2i)
16 a
2 (Γ2 + Γ cos i) − sin2 i
  i
− 6C22 sin (2πΓ) cos (2 (Ω − ψ0 + Γ (θ0 − π)))
Γ(Γ − 1)(Γ + 1)π
(5.82)
  2 h
3 R 
C20 1 − 3 sin2 i sin2 θ + 6C22 sin (2 (Ω − ψ)) cos i sin 2θ

ωn = n 1 −
4 a

1 2
 i
− cos (2 (Ω − ψ)) 1 + 3 cos 2θ − 2 cos 2i sin θ
4
(5.83)
  2 h
3 R
ω 2n = n2 1 − C20 (1 + 3 cos 2i)
8 a
2 (Γ2 + Γ cos i) − sin2 i
  i
− 6C22 sin (2πΓ) cos (2 (Ω − ψ0 + Γ (θ0 − π)))
Γ(Γ − 1)(Γ + 1)π
(5.84)
71

  2
3 R h 
ωn2 2
C20 1 − 3 sin2 i sin2 θ + 6C22 sin (2 (Ω − ψ)) cos i sin 2θ

= n 1−
2 a

1 2
 i
− cos (2 (Ω − ψ)) 1 + 3 cos 2θ − 2 cos 2i sin θ
4
(5.85)
Note that the terms ωr and ω̇r disappear for an orbit with zero inclination. This
matches our expectation that, absent unmodeled perturbations, a spacecraft orbiting
purely in the plane of reference can be expected to remain in that plane.
Substituting the kinematic equations obtained using averaging into 5.39 and set-
ting the result equal to the sum of equations 5.30 - 5.32, and simplifying, we can
express the first linearized relative motion model in a convenient matrix-vector form
as follows:

       
ẍ P11 P12 P13 x 0 2ω n 0 ẋ
ÿ  = n2 P21 P22 P23  y  + −2ω n 0 2ωr  ẏ  (5.86)
z̈ P31 P32 P33 z 0 −2ωr 0 ż

By instead using the kinematic equations obtained without averaging, we obtain


an alternate model:

       
ẍ Q11 Q12 Q13 x 0 2ωn 0 ẋ
ÿ  = n2 Q21 Q22 Q23  y  + −2ωn 0 2ωr   ẏ  (5.87)
z̈ Q31 Q32 Q33 z 0 −2ωr 0 ż

The terms in the P matrix (for the averaged model) are defined below, where
ξ0 = 2 (Ω − ψ0 + Γ (θ0 − π)). These are followed by the terms in the Q matrix for
the unaveraged model. Note that the differences between the P and Q matrix are
restricted to the upper left 2 × 2 submatrix.
72

 2  
R 2 2
 3
P11 = 3 + C20 6 3 sin i sin θ − 1 − (1 + 3 cos 2i)
a 8
 2 
R
3 cos (2 (Ω − ψ)) (3 + cos 2i) cos 2θ + 2 sin2 i

+3C22
a
  
3 (−1 + 4Γ (Γ + cos i) + cos 2i)
− 12 sin (2 (Ω − ψ)) cos i sin 2θ + sin 2πΓ cos ξ0
8Γ (Γ2 − 1) π
  2  2 h
R 2 R
P12 = − 6 C20 sin i sin 2θ − C22 cos (2 (Ω − ψ)) (3 + cos 2i) sin 2θ
a a
i
+4 sin (2 (Ω − ψ)) cos i cos 2θ
  2  2 h
R R
P13 = − 15 sin i C20 cos i sin θ + 2C22 cos (2 (Ω − ψ)) cos i sin θ
a a
i
+ sin (2 (Ω − ψ)) cos θ

P21 = P12
 2
21 R
P22 = C20 sin2 i cos 2θ
4 a
 2   
3 R 3 (−1 + 4Γ (Γ + cos i) + cos 2i)
+ C22 sin 2πΓ cos ξ0
8 a Γ (Γ2 − 1) π

2

−2 cos (2 (Ω − ψ)) 7 (3 + cos 2i) cos 2θ + 6 sin i + 56 sin (2 (Ω − ψ)) cos i sin 2θ
  2 h
R
P23 = 6 sin i 2C22 cos (2 (Ω − ψ)) (cos i − Γ) cos θ
a
 2 
i R
+ sin (2 (Γ − ψ)) (Γ cos i − 1) sin θ + C20 cos i cos θ
a
P31 =P13
 2
R
P32 =12C22 Γ sin i (cos (2 (Ω − ψ)) cos θ − sin (2 (Ω − ψ)) cos i sin θ)
a
 2 h
1 R i
P33 = − 1 + C20 6 + 12 cos2 i − 30 sin2 i sin2 θ
4 a
 2 h
1 R
cos (2 (Ω − ψ)) −15 (3 + cos 2i) cos 2θ − 54 sin2 i

+ C22
4 a
i
+60 sin (2 (Ω − ψ)) cos i sin 2θ
(5.88)
73

 2  
3 R 2 2

Q11 = 3 − C20 20 1 − 3 sin i sin θ
8 a
 2 
3 R
30 cos (2 (Ω − ψ)) (3 + cos 2i) cos 2θ + 2 sin2 i

+ C22
8 a

−120 sin (2 (Ω − ψ)) cos i sin 2θ
 2  2 
R 2 3 R
Q12 = − 6C20 sin i sin 2θ + C22 8 cos (2 (Ω − ψ)) (3 + cos 2i) sin 2θ
a 4 a

+32 sin (2 (Ω − ψ)) cos i cos 2θ

Q21 = Q12
 2  2 
R 2 3 R
Q22 = 3C20 sin i cos 2θ − C22 4 cos (2 (Ω − ψ)) (3 + cos 2i) cos 2θ
a 4 a

−16 sin (2 (Ω − ψ)) cos i sin 2θ

Q13 = Q31 = P13 , Q23 = P23 , Q32 = P32 , Q33 = P33


(5.89)

R 2

By setting C22 = 0 in the first model and collecting α = −C20 a
, we recover
the model derived in Chapter 4, which modeled relative motion in near-circular orbits
perturbed by the Earth’s dominant C20 gravitational term. Applying the similar
procedure to the second model would result in an unaveraged alternate form of the
model in Chapter 4. Note that the model obtained using averaging has terms that
will result in 0/0 singularities for Γ = 1, but the model remains analytic, since these
singularities can be removed by applying L’Hospital’s rule.
These models are stand-alone and do not require any integration of the chief orbit.
Furthermore, applying the transformation ψ = ψ0 + Γ (θ − θ0 ) and using the initial
values for Ω and i, the model is expressed entirely in terms of the argument of latitude
θ. The substitution θ = nt + θ0 can be used to render this model as an LTV system.
However, due to the altered orbit period not captured by this substitution, the model
accuracy will degrade over time. The argument of latitude can be obtained for longer
74

time spans by directly integrating equation 5.65 with either choice of θ̇u .

5.1.2 Extension to Eccentric Chief Orbits

For small eccentricity, the main parts of the derivation that need to be modified are
the final treatment of the eccentricity-dependent terms h and ρ. We must use r = aρ,
p
h = a (1 − e2 ), and ρ as defined in equation 4.3 wherever they appear in the model
equations (including the equation for θ̇u ). Since the smallest terms retained in the
 
R 2
model equations are O C20 a (or their equivalent C22 terms), we want to avoid
including any eccentricity terms that are significantly smaller than this. Thus, we
 2 
will assume O (ep ) = O C20 Ra , where p is an integer. For example, if p = 2, we
will retain terms up to O (e2 ). Similarly, if p = 1, terms that are O (e) will be treated
 2 
equivalently to terms that are O C20 Ra . For example, consider the application
of such a procedure to equation 5.30 for p = 1:
     
2x 2x 2x
µ µ
∆r̈C00 = 3  −y  = 3 ρ−3  −y  = n2 (1 − 3e cos M )  −y  (5.90)
r a
−z −z −z

Recall M = n (t − tp ) and note that a binomial expansion has been used to retain
only the linear terms in e. The extension of this procedure to all other model terms
should be relatively straightforward for the reader at this point. Only some of the
kinematic terms will retain eccentricity-dependent terms for p = 1, and the final
∆r̈C20 and ∆r̈C22 will be unchanged from Section 5.1.1.
Furthermore, note that for increasingly eccentric orbits, the linear transformation
equation 5.60 becomes invalid, along with the current centripetal balance method used
to obtain θ̇u . To address such cases, it is best to completely avoid use of equation
5.60, and to keep ψ = ψ0 + ct. The only instance where this transformation was truly
unavoidable was in averaging the R-T-N components of aC22 . A modified centripetal
balance equation is now derived, which will bypass the need for averaging and may
also be more accurate for non-circular orbits, due to the inclusion of a previously
75

neglected term.
The centripetal balance method for obtaining θ̇u finds the normal component of
the angular velocity by first equating the acceleration in LVLH frame components
with the gravitational acceleration terms, noting (˙) = H d ( ) here: dt

r̈ + 2ω O × ṙ + ω̇ O × r + ω O × (ω O × r) = aC00 + aC20 + aC22 (5.91)

Substituting ω O = (ω H · ên ) ên = θ̇u ên and dotting both sides by êr :
µ
r̈ + 2θ̇u (ên × ṙ) · êr + r (ω̇ O × êr ) · êr − θ̇u2 r = − + RC20 + RC22 (5.92)
r2
H d
Noting that ṙ = dt
(r) = ṙr êr and expressing ω̇ O = ω̇r êr + ω̇t êt + θ̈u ên , where
ω̇t = 0, we obtain:
µ
r̈ − θ̇u2 r = − + RC20 + RC22 (5.93)
r2
Note that for a circular orbit (r = a ∀t), r̈ = 0, but this is no longer the case. Since
r = aρ and n2 = µ/a3 , we obtain the following expression for the component of the
angular velocity normal to the (osculating) orbital plane:
s  
1 1
θ̇u = n2 − (RC20 + RC22 − r̈) (5.94)
ρ3 r

Before, this equation was used with the orbit-averaged radial components of the
gravitational acceleration terms (assuming r̈ = 0) to obtain a constant θ̇u = n∗
corresponding to a corrected orbital period T ∗ = 2π/n∗ . Recall, however, that using
the instantaneous RC20 (θ) and RC22 (θ) results in θ̇u = f (θ) whose average over θ is
still n∗ . Thus, the centripetal balance method results in the same period modification
whether the instantaneous or time-averaged radial accelerations are used. Use of
the instantaneous radial acceleration terms is beneficial because it allows the model
kinematics to account for both the modified orbital period and the time-varying
nature of the orbital speed.
For a near-frozen orbit, r(t) ≈ r(t+T ) and the orbit radius r can thus be expressed
in terms of a series of periodic functions of a continuously increasing angular variable.
76

If the semimajor axis does not significantly change over time, r(t) = aρ(t) and r̈(t) =
aρ̈(t) may be used in equation 5.94, with ρ given in equation 4.3. Thus, we can obtain
a more accurate expression for θ̇u , after this final substitution:
s  
1 1
θ̇u = n2 3
− (RC20 (θ) + RC22 (θ) − aρ̈) (5.95)
ρ aρ

Equation 5.95 may be treated as the noncircular orbit equivalent of equation 5.66.
Assuming that ȧ ≈ 0 and ė ≈ 0 and tp does not change over time, we can determine
ρ̈:
d2  2 d 
ρ̈ = ρ(e, M ) Ṁ + ρ(e, M ) M̈ (5.96)
dM 2 dM
1/2
where Ṁ = n + ṅ (t − tp ) ≈ n, since ṅ = d/dt (µ/a3 ) = −3nȧ/2a ≈ 0. Thus M̈ = 0
and we obtain:
d2
ρ̈ = n2

2
ρ(e, M ) (5.97)
dM
To O (e3 ), with M = n (t − tp ):
 
2 2 3 3
ρ̈ = n e cos M + 2e cos 2M + e (− cos M + 9 cos 3M ) + ... (5.98)
8
d
Finally, recall that the convenient derivative substitution Ṡ = dθ
(S) θ̇ was previ-
ously used for obtaining the time derivative of any generic parameter S(ψ, θ) after
applying the linear transformation 5.60 to make S a function of θ alone. (See equation
5.76 for example). For cases where use of equation 5.60 is to be avoided, the time
derivative of any S(ψ, θ) must be explicitly calculated.
Note that if all of the corrections suggested in this section were applied, the
resulting model will still only be valid for situations satisfying the following criteria:

1. Small separation between chief and deputy spacecraft ( ∆r


a
<< 1)

2. Nearly constant chief semimajor axis and eccentricity (ȧ ≈ ė ≈ 0)

3. Slow precession of the orbital plane, to justify the use of initial values of Ω0 and
i0 in the final model equations
77

4. Nearly constant time of periapsis crossing tp

R 2 R 2
 
5. Sufficiently small C20 a
and C22 a
to justify neglecting higher-order terms

Application of the model to situations violating these criteria will result in in-
creased error. Generally, these restrictions describe a very desirable subset of near-
frozen orbits that would be practical choices for mission design. Thus, the assump-
tions made in deriving the model should not unreasonably reduce its utility.

5.2 Model Analysis

In this section, we will focus on several interesting details of the two models obtained
for the case of vanishing chief eccentricity. The analysis in this section serves to
provide additional confirmation of the models derived, and to obtain potentially useful
extensions and conclusions.

5.2.1 The Case of Equatorial Orbits

First, note that both models predict a decoupling of the z dynamics for orbits in the
plane of reference (i = 0). The following useful identities are introduced:

cos (2 (Ω − ψ)) cos 2θ − sin (2 (Ω − ψ)) sin 2θ = cos (2 (Ω − ψ + θ)) (5.99)

cos (2 (Ω − ψ)) sin 2θ + sin (2 (Ω − ψ)) cos 2θ = sin (2 (Ω − ψ + θ)) (5.100)

Using the equations for the model obtained without averaging, letting i → 0 and
using the identities 5.99 and 5.100 along with the linear transformation ψ = ψ0 +
Γ (θ − θ0 )), we obtain the following simple equations for relative motion dynamics for
78

near-equatorial orbits:
   
2 5 2 3 9
ẍ = 3n 1 + α + 15β cos Θ x + 24n β sin Θy + 2n 1 + α − β cos Θ ẏ
2 4 2
 
3 9
ÿ = 24n2 β sin Θx − 12n2 β cos Θy − 2n 1 + α − β cos Θ ẋ
4 2
 
9
z̈ = − n2 1 − α + 15β cos Θ z
2
(5.101)

R 2 R 2
 
where α = −C20 a
and β = C22 (positive quantities for a stably rotating
a

ellipsoidal body), and Θ = 2 (1 − Γ) θ + (Ω + Γθ0 − ψ0 ) . Note that for i = 0,
Ω is undefined, so θ would be measured from the reference direction γ, and thus

Θ = 2 (1 − Γ) θ + (Γθ0 − ψ0 ) . For Γ = 1, the θ-dependent terms completely vanish,
and we obtain an LTI system similar to the HCW equations, but with additional
in-plane coupling terms and rescaling of the coefficients in the HCW ODEs. All
additional terms are simply functions of the gravitational coefficients and the initial
conditions. If the chief remains fixed in the rotating body frame (the case of Γ = 1),
we would expect the relative dynamics equations to lose all time-varying terms, and
the model satisfies this condition.

5.2.2 A Special Case LTI Model

For an equatorial orbit and Γ = 1, we can express the model in a useful LTI form
ẋ = Ax, where x = [x, y, z, ẋ, ẏ, ż]> and A is a 6 × 6 matrix given below:
 
03×3 I3×3
A= (5.102)
Ap Av

3n 1 + 52 α + 15β cos 2λ
 2  
24n2 β sin 2λ 0
Ap =  24n2 β sin 2λ −12n2 β cos 2λ 0 
9
2

0 0 −n 1 − 2 α + 15β cos 2λ
(5.103)
3 9
  
3
0
9
 2n 1 + 4 α − 2 β cos 2λ 0
Av =  −2n 1 + 4
α − 2
β cos 2λ 0 0 (5.104)
0 0 0
79

where λ = θ0 − ψ0 = Θ/2 for the special case of an equatorial orbit with Γ = 1.


There are two cases of λ in particular that we are interested in analyzing, cor-
responding to relative motion about the two distinct types of libration points that
exist in the body frame of a rotating ellipsoidal body. The first case is for λ = 0, in
which the chief spacecraft position vector would be aligned with Î. In the second case,
λ = π/2, and the chief position vector is aligned with Ĵ. These libration points each
share a dynamically identical case for λ + π. All four libration points are depicted in
Fig. 5.2.

Figure 5.2. Libration Points for Rotating Ellipsoidal Body

Substituting λ = 0 into A, we obtain the characteristic equation for the eigenvalues


by computing det (sI − A) = 0:
 
6 1 2
s + n 8 − 24α + 9α − 108 (2 + α) β + 324β s4
2 2
4

1 4
+ n − 81α3 + 18α2 (4 + 69β) − 12α (4 + 57β (−2 + 9β))
8 (5.105)

+8 (1 + 15β) (1 + 3β (−35 + 27β)) s2

−9n6 β (2 − 9α + 30β) (2 + 5α + 30β) = 0


80

We define the useful constants c1 and c2 to aid in the upcoming analysis:

c1 = 4 − 6α + 9α2 − 12 (23 + 9α) β + 324β 2 (5.106)

c2 = 1152β (2 + 5α + 30β) (5.107)

The six roots for this equation are the eigenvalues of the system:
√ r
2
q
s1,2 = ± ni c1 + c2 + c21
4 r

2
q
s3,4 = ± n −c1 + c2 + c21 (5.108)
4r
9
s5,6 = ± ni 1 − α + 15β
2
It is clear that in this case, only eigenvalues with zero real part will allow us to avoid
instability of the linearized dynamics. For positive values of α and β, we have four
stability conditions that must be satisfied:
q
1. c1 + c2 + c21 > 0
q
2. − c1 + c2 + c21 < 0
(5.109)
9
3. 1 − α + 15β > 0
2
4. c2 + c21 ≥ 0
p
However, we know c2 + c21 > c1 since c2 > 0, so the second stability condition
cannot be satisfied, and thus the special LTI case ẋ = Ax is unstable for λ = 0 (and
similarly, λ = π). Motion about this libration point was shown to be unstable by
Kaula.31 Work by Abalakin32 in 1957 and Zhuravlev33 in 1972 similarly predicted
instability of the libration points along the long-axis of a triaxial ellipsoid. This new
model successfully predicts this instability property!
Substituting λ = π/2 into A, we obtain the characteristic equation for the eigen-
81

values by computing det (sI − A) = 0:


 
6 1 2
s + n 8 − 24α + 9α + 108 (2 + α) β + 324β s4
2 2
4

1 4
− n 81α3 + 18α2 (−4 + 69β) + 12α (4 + 57β (2 + 9β))
8 (5.110)

+8 (−1 + 15β) (1 + 3β (35 + 27β)) s2

−9n6 β (−2 + 9α + 30β) (2 + 5α − 30β) = 0


The six roots for this equation are the eigenvalues of the system:
√ r
2
q
s1,2 = ± ni c1 + c21 − c2
4 r

2
q
s3,4 = ± ni c1 − c21 − c2 (5.111)
4r
9
s5,6 = ± ni 1 − α − 15β
2
Again, we seek conditions that give the eigenvalues zero real part. For positive values
of α and β, we have four new stability conditions:
q
1. c1 + c21 − c2 > 0
q
2. c1 − c21 − c2 > 0
(5.112)
9
3. 1 − α − 15β > 0
2
2
4. c1 − c2 ≥ 0
Unlike in the previous case, it is clear that there is no contradiction in the stability
conditions. It can be shown that since c2 > 0 for positive α and β, stability is ensured
by the following conditions:
1. c21 ≥ c2 > 0
9 (5.113)
2. 1 − α − 15β > 0
2
There are positive choices of α and β that will satisfy both of these stability conditions.
Thus, for the libration points at λ = π/2 and λ = 3π/2, we have the possibility for
marginally stable linearized dynamics, which again agrees with the work of Kaula,31
Abalakin,32 and Zhuravlev.33
82

5.2.3 Analysis of Unstable and Stable Eigenspaces

From equations 5.102 - 5.104, it is clear that the x-y motion and z motion are decou-
pled. The in-plane and out-of-plane dynamics are thus expressed separately:

∆r̈(x,y) + Av(x,y) ∆ṙ(x,y) + Ap(x,y) ∆r(x,y) = 02×1 (5.114)


 
2 9
z̈ + n 1 − α + 15β cos 2λ z = 0 (5.115)
2
where Ap(x,y) and Av(x,y) are the upper left 2×2 submatrices of Ap and Av in equations
5.102 - 5.104. The first four eigenvalues in equations 5.108 and 5.111 are each asso-
ciated with the in-plane modes, and s5,6 are associated with the out-of-plane mode.
The solution for the in-plane and out-of-plane motion is given in equations 5.116 and
5.117, where the Ci are six arbitrary constants determined by the initial conditions.
4
X
∆r(x,y) (t) = C i ν i e si t (5.116)
i=1

6
X
z(t) = C i e si t (5.117)
i=5

Focusing on the in-plane dynamics, we solve the following for the eigenvector ν i
associated with each eigenvalue si :

s2i I2×2 + si Av(x,y) + Ap(x,y) ν i = 02×1



(5.118)

The solution for the eigenvectors is of the form given in equation 5.119, where di is a
constant. The constant di is given by equation 5.120 for the unstable libration points
(λ = 0, π) and equation 5.121 for the stable libration points (λ = π/2, 3π/2).
 
1 di
νi = p 2 (5.119)
di + 1 1

(s2i − 12n2 β)
di(u) = (5.120)
2n 1 + 43 α − 92 β si


(s2i + 12n2 β)
di(s) = (5.121)
2n 1 + 43 α + 92 β si

83

For a triaxial ellipsoid with semi-axes (A, B, C) of 6, 3, and 2.5 km, density of
2.6 g/cm3 , and rotation period of 77.2 hours, we obtain a synchronous circular orbit
radius of 40 km. Thus α = 0.002, β = 8.4375 × 10−4 , and the body and orbit angular
rates are c = n = 2.2607 × 10−5 rad/s. The angular rate-normalized eigenvalues
sni = si /n are given for motion about the unstable and stable libration points in
Table 5.3.

Table 5.3. Unstable and Stable Eigenvalues for Example Case


Unstable Stable
sn1,2 = ±0.9852i sn1,2 = ±0.9511i
sn3,4 = ±0.1785 sn3,4 = ±0.1849i
sn5,6 = ±1.0018i sn5,6 = ±0.9890i

Motion about the unstable libration point is oscillatory in the z-direction (with
a period of 77.06 hours, nearly matching the period of asteroid rotation). The os-
cillatory in-plane mode period for this system is 78.36 hours. The in-plane motion
is confirmed to be unstable from the positive real eigenvalue sn3 = 0.1785. The
Lyapunov time for this system is thus only 68.84 hours.
Motion about the stable libration point is confirmed to be stable from the com-
puted eigenvalues. As expected, all six eigenvalues are purely imaginary. The out-of-
plane mode is oscillatory with nearly the same period as the asteroid rotation. The
in-plane motion is composed of two modes: the short-period mode (T = 81.17 hrs)
for sn1,2 = ±0.9511i and the long-period mode (T = 417.54 hrs) for sn3,4 = ±0.1849i.
A superposition of the short and long-period in-plane modes is simulated by setting
the constants from the equation for the general solution: C1 = 1, C2 = −0.9, C3 =
0.2, C4 = 0.1. The resulting motion is plotted in Fig. 5.3. The amplitude of the x
and y motion is equal, but the axes are unlabeled since in the linear regime the scale
of the orbit is unimportant. In Fig. 5.4, we see that actual in-plane motion (from
simulation results - see Section 5.3.5) can correspond very closely with this example.
This striking similarity helps to confirm the validity of this analysis.
84

Figure 5.3. Superposition of In-Plane Modes at Stable Libration Point

0.1

0.05

-0.05

-0.1
0.2 0.15 0.1 0.05 0 -0.05 -0.1 -0.15 -0.2

Figure 5.4. Simulated In-Plane Motion at Stable Libration Point

Further interesting results can be obtained by returning to the discussion of the


behavior for the whole system as defined in equations 5.102 - 5.104. Let qi and pi be
right and left eigenvectors, respectively, of matrix A. Thus, Aqi = si qi , p> >
i A = pi si ,

and q>
i pj = δij . The solution of ẋ = Ax can be expressed as:
6
X
At
x (t) = e x (0) = esi t qi p>
i x (0) (5.122)
i=1

For the solution to be in the i th mode, the initial conditions must satisfy x(0) = Cqi ,
p>
j x(0) = 0, j 6= i.
34
Furthermore, the stable, center, and unstable subspaces E s ,
E c , and E u are the subspaces of R6 spanned by the real and imaginary parts of
the eigenvectors qi corresponding to eigenvalues si with negative, zero, and positive
real parts respectively. For motion in the body-fixed frame of a uniformly rotating
body with a second degree and order gravity field, stable (W s ) and unstable (W u )
manifolds of the unstable libration point exist, and are tangent to E s and E u at the
85

libration point.35 If the stable and unstable subspaces obtained from the special case
LTI model can be used to obtain these manifolds, it will provide additional analytical
confirmation of the validity of the unaveraged relative orbital motion model from
which the special case LTI model was derived.
The stable and unstable manifolds can be found by numerically integrating the
true nonlinear dynamics with an initial condition on the linear E s and E u subspaces
a short distance from the libration point. The stable manifold must be obtained
by integrating the equations backwards in time. The nonlinear dynamical equations
describing motion in a rotating body-fixed frame are given below.36 This x, y, z
coordinate system is centered on the primary body center of mass, and aligned with
its principal axes, parameterized by the orthogonal unit vectors Î, Ĵ, K̂ defined in
Fig. 5.1. Recall c is the angular velocity of the body rotation about K̂.

µx ∂U2
ẍ − 2cẏ = c2 x − + (5.123)
r3 ∂x
µy ∂U2
ÿ + 2cẋ = c2 y − + (5.124)
r3 ∂y
µz ∂U2
z̈ = − + (5.125)
r3 ∂z
U2 is the second degree and order force potential in the body-fixed frame, resolved in
Cartesian coordinates:

µC20 (x2 + y 2 − 2z 2 ) 3µC22 (x2 − y 2 )


U2 = − + (5.126)
2r5 r5

The four previously discussed equilibrium points for this system can be represented
as E±x = (±xeq , 0, 0) and E±y = (0, ±yeq , 0). The distances xeq and yeq can be found
to first order in C20 and C22 via a perturbative approach:36
 
∗ 1 3
xeq = r 1 − ∗2 C20 + ∗2 C22 (5.127)
2r r
 
∗ 1 3
yeq = r 1 − ∗2 C20 − ∗2 C22 (5.128)
2r r
86

µ 1/3
where r∗ =

c 2 is the circular orbit radius for an orbital mean motion commensu-
rate with the angular rate of rotation of a spherical body. In general, since |C20 | << 1
and |C22 | << 1, xeq ≈ yeq ≈ r∗ .
Using the A matrix for the special case LTI system and the eigenvalues given
in Table 5.3 for the previously defined hypothetical asteroid, the eigenvectors can
be numerically obtained by solving (si I6×6 − A) qi = 06×1 . Doing this with s3 and
s4 for motion about the unstable libration point, we obtain the stable and unstable
subspaces (in units of km and km/s):
   
−0.1171 0.1171

 −0.9931 


 −0.9931 

 0.0   0.0 
Es =   4.7228 × 10−7
u
, E =   (5.129)




 4.7228 × 10−7 

 4.0068 × 10−6   −4.0068 × 10−6 
0.0 0.0
Using the stable and unstable subspaces given above and the nonlinear dynamics
in equations 5.123 - 5.125, we are able to obtain the unstable and stable manifolds
using the following initial and final conditions, respectively:
 
xeq
x (0) = ± E u (5.130)
05×1
 
xeq
x (tf ) = ± E s (5.131)
05×1
where  = 10−3 provides a satisfactory small deviation from the libration point. The
sign of this small deviation will simply determine whether the y < 0 or y > 0 branch
of the given manifold is obtained. The unstable and stable manifolds are shown in
Fig. 5.5. The time for either manifold to bring the satellite from the vicinity of one
unstable libration point to the other is around 110 rotation periods (nearly 354 days).
The LTI model (given by equations 5.102 - 5.104) was obtained from the unaver-
aged relative motion model (equation 5.87) for the special case of equatorial orbits
with Γ = 1. Since this LTI model has been shown to successfully predict the stable
and unstable subspaces associated with motion about the unstable libration point, we
87

have significant additional analytical evidence of the validity of the new unaveraged
relative motion model.

40

30

20

10

-10

-20

-30

-40

-50 -40 -30 -20 -10 0 10 20 30 40 50

Figure 5.5. Stable and Unstable Manifolds (Unstable Libration Point)

5.3 Model Validation via Simulation

In this section, we will focus on validating the two near-circular orbit models obtained
in 5.1.1. A number of parameter studies were performed to evaluate the performance
of the new models, and their performance was compared to the HCW model and
the model obtained in Chapter 4 (which considers only one of the two second-order
gravity terms). The error in all models was estimated by comparing their results to
the results of a truth model, which separately simulated the motion of both a chief
and deputy spacecraft using equation 5.13, then subtracted the chief state from the
deputy, and rotated the result into the moving LVLH frame.
88

5.3.1 Hypothetical Asteroid for Case Studies

To validate the models obtained in 5.1.1, we will simulate and study the relative
motion around a hypothetical asteroid described by the parameters in Table 5.4. The
spin rate and initial configuration angle ψ0 are treated as variables in the upcoming
study. The maximum diameter of 12 km is realistic, with thousands of cataloged
asteroids of this size or greater. The density is fairly typical for an S-type asteroid.

Table 5.4. Simulated Asteroid Data


Semi-axes (A, B, C) (6.0, 3.0, 2.5) km
Density 2.6 g/cm3
Mass 4.901 × 1014 kg
Gravitational Parameter µ = 3.2709 × 10−5 km3 s-2
Gravitational Coefficients C20 = −0.0903, C22 = 0.0375

The true orbital dynamics are assumed to be sufficiently approximated by equa-


tion 5.13. Inclusion of higher-order gravity terms and other perturbations would be
inappropriate in this context of validating the new linearized models, but such effects
should be included for higher-fidelity simulations.

5.3.2 Variable Chief Inclination with Constant Γ

The notation for orbit elements and differences is unchanged from Chapters 1 and
2. For this study, we consider the initial chief orbit elements œc and deputy orbit
element differences δœ given in Table 5.5, noting œd = œc + δœ.
In this study, the chief inclination i is varied (for many simulations) from 0 to
90 degrees, with the body initial configuration angle ψ0 = π/8. All other initial

Table 5.5. Orbital Elements, Variable Chief Inclination


a (km) e i (◦ ) Ω (◦ ) ω (◦ ) f (◦ )
œc 40.0 0.0 i 0.0 0.0 0.0
δœ 0.0 0.002 0.2 0.0 0.0 0.0
89

orbit elements (and orbit element differences) are constant. The unperturbed relative
orbit resulting from these initial conditions would be non-drifting, with an average
separation of 153 meters. The elements were chosen such that the x, y, and z average
separation are similar for the unperturbed relative orbit. Although the separation
may seem small, with a semimajor axis of only 40 km, 153 m would be equivalent to
a 27 km average separation for a low-Earth orbit!
Three values for the angular rate ratio Γ are considered. For the first set of plots,
we consider an asteroid rotation period of Tb = 32 hours, so Γ = 2.4. To otain the
second set of plots, the asteroid rotation period is set to Tb = 64 hours, resulting in
Γ = 1.2. Finally, setting the asteroid rotation period to 96.5 hours, our third case is
for Γ = 0.8. Note that the angular rate ratio Γ and asteroid rotation period Tb are
related by Γ = 2π/ (3600nTb ). The data in the plots that follow are all obtained after
simulating one full orbit.
For each case of Γ, two plots are provided: (1) the average modeling error norm
and (2) the max deviation of relative orbit angular momentum. The average modeling
error norm is obtained as before: subtracting the true position ∆rT (t) from the model-
predicted position ∆rM (t), computing the norm of this “error vector” at every time
step, then averaging the norms over one full orbit (cf. Section 2.2.3). The max
deviation of relative orbit angular momentum is defined as max(k∆h (t) − ∆h (t0 ) k),
where ∆h = ∆r × ∆v, an important angular momentum quantity that was discussed
in Section 2.3.3. This quantity is particularly useful for representing large changes to a
relative orbit that would otherwise be non-drifting and periodic if perturbations were
absent. The models are evaluated by these two very different criteria for good reason:
no one plot can fully describe the degree to which a model is successfully simulating
the relative orbit dynamics. If a model shows superior performance regardless of the
evaluation criteria or presentation of the data, it is most likely correctly simulating
the dynamics.
90

i. Variable Inclination, Γ = 2.4

For this case, the average modeling error and max deviation of relative orbit angular
momentum are provided in Figs. 5.6 and 5.7. The two models developed in Section
5.1.1 are consistently labeled as the averaged and unaveraged “second-order models,”
since they are models that account for both terms in the second degree and order
gravity field: C20 and C22 . The model developed in Chapter 4 for J2 -perturbed
relative orbits is referred to as “C20 only,” since C20 = −J2 . From Fig. 5.6, it is clear
that the two second-order models tend to offer much greater performance than the
HCW and C20 -only model for Γ = 2.4 and the orbital parameters given in equation
5.5. The unaveraged model has an average modeling error norm of roughly 1-3 meters
over the whole inclination range of prograde chief orbits tested.

14

12

10

0
0 10 20 30 40 50 60 70 80 90

Figure 5.6. Average Modeling Error vs. Inclination, Γ = 2.4

Interestingly, the average error performance of the averaged and unaveraged mod-
els converges and flips at i ≈ 65◦ , with the averaged model error being the lowest
under this value, and the unaveraged model having the lowest error for higher incli-
91

nations. Although we expect the unaveraged solution to have the best performance
for many if not all cases, it is possible that the angular rate ratio Γ is high enough
in this case that averaging the time-varying effect of the C20 and C22 gravitational
perturbations on θ̇ does not result in appreciable degradation of the accuracy of the
unaveraged model. The C20 -only model doesn’t offer much improvement over the
HCW model, since it only accounts for one of the two gravitational terms. This
demonstrates that it is so important to account for all dynamical terms of the same
order, otherwise the resulting model could easily have worse error.
Fig. 5.7 shows that the two second-order models are much better at predicting the
maximum deviation of the relative orbit angular momentum than the HCW model
or the C20 -only model. Recall from Section 2.3.3 that the maximum deviation of the
relative orbit angular momentum is useful in these parameter studies for visualizing
how changes in a parameter (in this case i) effect the scale of perturbations on the
relative orbital motion. Thus from Fig. 5.7, it is clear that for Γ = 2.4 and the given
orbital parameters, the relative motion is generally most perturbed for lower values
of the chief inclination. Note that the HCW model predicts the same value for all
inclinations, because it does not account for the nonsphericity of the Earth at all.
The model only accounting for C20 does only a little better since it fails to account
for C22 .
Comparing the results for the second-order models in Figs. 5.6 and 5.7, it is
evident that there is some correlation between the relative closeness of the predicted
max(k∆h (t) − ∆h (t0 ) k) and the relative size of the average modeling error norm.
For low inclinations, the averaged second-order model has a lower error norm and
also the closest agreement with the true value for the maximum relative orbit angular
momentum deviation. For higher inclination angles, the unaveraged second-order
model’s prediction is notably close to the true angular momentum deviation, and its
error is also similarly much lower than for the averaged model. While this correlative
trend generally holds for both second-order models (which will become clear as more
92

10-8
8

0
0 10 20 30 40 50 60 70 80 90

Figure 5.7. Maximum Deviation of Relative Orbit Angular Momentum, Γ = 2.4

results are presented), it does not hold for the HCW and C20 -only models. This
reinforces the fact that the average error norm does not give the full picture of the
relative performance of the models.

ii. Variable Inclination, Γ = 1.2

For this case, the average modeling error and max deviation of relative orbit angular
momentum are provided in Figs. 5.8 and 5.9. The data is presented in an identical
manner as in the previous case, with all parameters identical to the previous case,
except for Γ. The performance of the two second-order models is different for this
case, with both models having generally worse performance for lower values of the
chief inclination. Fig. 5.8 shows that the averaged second-order model notably has
the worst error when the inclination is less than 38◦ . The unaveraged second-order
model has better performance, and generally has the lowest error of all four models,
except for low inclinations and inclinations above 75◦ , where its averaged counterpart
has slightly lower error.
Inspection of Fig. 5.9 indicates that the unaveraged second-order model is very
93

15

10

0
0 10 20 30 40 50 60 70 80 90

Figure 5.8. Average Modeling Error vs. Inclination, Γ = 1.2

10-7
2

1.8

1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 10 20 30 40 50 60 70 80 90

Figure 5.9. Maximum Deviation of Relative Orbit Angular Momentum, Γ = 1.2


94

good at estimating the maximum deviation of the relative orbit angular momentum.
Again, for the second-order models, there is strong correlation between the relative
closeness of the predicted max(k∆h (t)−∆h (t0 ) k) and the relative size of the average
modeling error norm. The drastically different model performance for this case of Γ
emphasizes the importance of this parameter.

iii. Variable Inclination, Γ = 0.8

For this case, the average modeling error and max deviation of relative orbit angular
momentum are provided in Figs. 5.10 and 5.11. The data is presented in the same
two types of plots as in the previous two cases. The averaged and unaveraged second-
order models tend to have the lowest error, but this varies with inclination. Overall,
we have seen that the modeling error of the unaveraged model seems to reduce as the
inclination is raised.

15

10

0
0 10 20 30 40 50 60 70 80 90

Figure 5.10. Average Modeling Error vs. Inclination, Γ = 0.8

Overall, this parameter study clearly shows that the new models outperform the
HCW and C20 -only solution. In over 90% of the cases tested in this parameter
95

10-7
1.5

0.5

0
0 10 20 30 40 50 60 70 80 90

Figure 5.11. Maximum Deviation of Relative Orbit Angular Momentum, Γ = 0.8

study, either the unaveraged or averaged second-order model gave the lowest error.
Specifically, in over 50% of the cases tested, the unaveraged second-order model has
the lowest error, while around 40% of the cases resulted in the averaged model giving
the lowest error. The averaged model seems to perform best in high-Γ, low-inclination
situations. This makes sense, because these are instances in which orbit-averaging the
effects of a rotating gravity field would seem somewhat reasonable. In cases with a
high inclination or a small value of Γ, the completely unaveraged model is the logical
choice.

5.3.3 Variable Γ with Constant Chief Inclination

In this study, we vary the angular rate ratio Γ from 0 to 5 for the constant set of orbit
elements and differences given in Table 5.5. This will be done for 3 different values
of inclination: i = 10◦ , 45◦ , 75◦ . It should be clear to the reader that changes in Γ
will not affect the HCW or C20 -only solution results at all, since these models have
no means of accounting for the rotation of the primary body. We will use the same
type of plots as in the previous study, except the x-axis now represents changes in Γ
96

with all other simulation parameters held constant.

i. Variable Γ, i = 10◦

The average modeling error and max deviation of relative orbit angular momentum
are provided in Figs. 5.12 and 5.13. From Fig. 5.12, it is clear that the two second-
order models tend to have lower error than the HCW and C20 -only models. There
is a range of Γ for which both of the new models have worse error. In general, the
averaged second-order model has worse performance for Γ < 0.8 and slightly better
performance for Γ > 1.5. The performance of the unaveraged second-order model is
more consistent vs. Γ. Fig. 5.13 shows that the two second-order models are mostly
successful in predicting the max deviation of the relative orbit angular momentum,
with the unaveraged model typically offering better performance.

35

30

25

20

15

10

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Figure 5.12. Average Modeling Error vs. Γ, i = 10◦

ii. Variable Γ, i = 45◦

The average modeling error and max deviation of relative orbit angular momentum
are provided in Figs. 5.14 and 5.15. The performance in this case for i = 45◦ is very
97

10-7
5

4.5

3.5

2.5

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Figure 5.13. Maximum Deviation of Relative Orbit Angular Momentum, i = 10◦

similar to the previous case of i = 10◦ .

25

20

15

10

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Figure 5.14. Average Modeling Error vs. Γ, i = 45◦

iii. Variable Γ, i = 75◦

The average modeling error and max deviation of relative orbit angular momentum
are provided in Figs. 5.16 and 5.17. In this case, the unaveraged second-order model
98

10-7
3.5

2.5

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Figure 5.15. Maximum Deviation of Relative Orbit Angular Momentum, i = 45◦

clearly offers consistently lower error than the HCW and C20 -only models, and often
outperforms the averaged second-order model as well. This again shows that the un-
averaged model is typically the best of the two new models to use for high-inclination
cases. From Fig. 5.17, we see that the two second-order models again predict the
max deviation of relative orbit angular momentum with reasonable accuracy.

12

10

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Figure 5.16. Average Modeling Error vs. Γ, i = 75◦


99

10-7
1.8

1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Figure 5.17. Maximum Deviation of Relative Orbit Angular Momentum, i = 75◦

Table 5.6. Orbital Elements, Variable Chief Semimajor Axis


a (km) e i (◦ ) Ω (◦ ) ω (◦ ) f (◦ )
œc a 0.0 75.0 0.0 0.0 0.0
δœ 0.0 0.002ϑ 0.2ϑ 0.0 0.0 0.0

5.3.4 Variable Chief Semimajor Axis with Constant Body Rotation


Rate

In this scenario, we set the body rotation period to be 36 hours and i = 75◦ , and vary
the chief semimajor axis from 25 to 60 km. The orbit element differences are linearly
scaled with the semimajor axis (as given in Table 5.6) so that the unperturbed relative
orbit dimensions are preserved and the unperturbed average separation remains 153
meters. The linear scaling parameter is ϑ = 40/a. Thus, we are isolating the effects
of distance from the asteroid on our model accuracy. It should be clear that raising
the chief semimajor axis will raise the value of Γ, since the orbital period increases
but the body rotation rate is held constant in this parameter study. The value of Γ
increases from Γ = 1.1 at 25 km to Γ = 3.9 at 60 km.
Fig. 5.18 shows that the second-order models offer consistently better performance
than the HCW and C20 -only models. Again, the unaveraged second-order model’s
100

25

20

15

10

0
25 30 35 40 45 50 55 60

Figure 5.18. Average Modeling Error vs. Semimajor Axis, i = 75◦

10-7
8

0
25 30 35 40 45 50 55 60

Figure 5.19. Max Deviation of Relative Angular Momentum vs. Semimajor Axis

performance is more consistent. It also offers the lowest error for much of the range
of chief semimajor axes tested. The performance of all models begins to converge for
higher values of semimajor axis, and the max deviation of relative orbit angular mo-
mentum also decays asymptotically (Fig. 5.19). This is showing that the perturbing
effects of the C20 and C22 gravity terms become less significant as the distance from
the primary body is increased, with the exception of the local maxima in the max
101

relative orbit angular momentum deviation at a = 26 and a = 33.


Note that any local maximum of max(k∆h(t) − ∆h(t0 )k) in these parameter stud-
ies represents “relative orbit resonance”, in which the oscillatory effects of the rotat-
ing gravity field apply periodic forcing to the relative orbit. It is obvious from these
parameter studies that the second-order models are very good at predicting these res-
onance regions in the parameter space. A practical application is that these models
could be used to perform quick numerical studies in the parameter space to identify
ranges of parameters which would result in highly destabilized relative orbital mo-
tion. While the full nonlinear dynamics can also be used to perform such studies, it
would be far more numerically efficient to use these linearized models instead. Thus,
these models could be used to quickly identify the situations which would obviously
be impractical for formation flying or proximity operations.

5.3.5 Interesting Cases

The previous parameter studies were very useful for demonstrating that the new
linearized relative motion models accounting for C20 and C22 are successful. In this
section, we will focus on several individual cases that demonstrate the utility of these
models or connect back to previously discussed topics.

i. A Case of Relative Orbit Resonance

First, we will revisit the relative orbit scenario first introduced in the end of Chapter
2. For that simulation, the orbit elements and differences are given in Table 2.1 and
the asteroid data is the same as what was given in Table 5.4. We set Γ = 1.6, so the
asteroid rotation period is just over 48 hours. The initial configuration angle ψ0 = π/8
is also the same. Note that we can extrapolate from Fig. 5.13 that Γ = 1.6 will yield
highly resonant behavior for this set of orbit elements and differences. Using the
same initial conditions, we show that the second-order models predict the resulting
102

dynamical behavior very well for two of the four orbits. Although they were derived
for relative motion with circular chief orbits, the small initial chief eccentricity has
very little effect on their accuracy.
First, new relative distance and relative orbit angular momentum plots are pro-
vided in Figs. 5.20 and 5.21. These are similar to Figs. 2.4 and 2.5, except they are
only shown for two orbits instead of four. These figures show how well each model
is able to track the true distance and true ∆h(t). It is clear that the second-order
models are able to track the changes in these quantities, while the HCW model and
C20 -only models completely fail to track the behavior. The HCW model predicts
that ∆h is constant, while the C20 -only model predicts small variations that are not
similar to the true behavior of ∆h(t). The HCW and C20 -only models also cannot
predict the relative orbit destabilization that occurs due to C22 .
The model error norm is provided in Fig. 5.22. After two orbits, the second-
order models have predicted the position of the deputy with an accuracy within tens
of meters, while the HCW and C20 -only model error has increased to nearly half a
kilometer. Two orbits may seem short, but since the orbital period is 3.2 days, these
new linearized models have successfully predicted the relative motion behavior for
nearly a week! Finally, Fig. 5.23 is a plot of the relative motion (predicted and
actual) in the LVLH frame. Only the second-order models are able to track the large
along-track motion of the deputy. The HCW and C20 -only models do not capture
this large destabilization of the relative motion at all.
This example clearly demonstrates that the second-order models have successfully
predicted the relative orbit resonance behavior with a high degree of accuracy. While
the averaged and unaveraged second-order models have similar performance for this
example, we would expect the averaged model performance to degrade with a higher
chief inclination. The unaveraged model does not suffer from such weaknesses because
of its fully rigorous treatment of the kinematics.
103

0.5

0.45

0.4

0.35

0.3

0.25

0.2

0.15

0.1

0.05
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 5.20. Relative Distance with Relative Orbit Resonance

10-7
14

12

10

2
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 5.21. ∆h(t) with Relative Orbit Resonance


104

0.45

0.4

0.35

0.3

0.25

0.2

0.15

0.1

0.05

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 5.22. Model Error (Relative Orbit Resonance Case)

0.15

0.1

0.05

-0.05

-0.1

-0.15
0.1
0
0.1 0 -0.1 -0.1
-0.2 -0.3 -0.4 -0.5

Figure 5.23. Relative Orbit (Resonance Case)


105

ii. A Case of Marginal Stability

While it has been shown that the second-order models successfully predict destabilized
relative motion, an equally interesting test would be to investigate how well the
second-order models predict motion in cases of true marginal stability of the relative
motion. In Section 5.2.2, we considered motion about the unstable and stable libration
points in the body-fixed frame of a rotating ellipsoidal asteroid. It was shown that
the reduced second-order model predicts (correctly) that marginally stable motion is
only possible about the libration points collinear with the minor axis. It will now be
shown via simulation that for the asteroid given in Table 5.4, this marginally stable
motion is realized, and successfully predicted by the unaveraged second-order model.
To do this, we consider a chief 40 km from the center of mass, in a line collinear
with the minor axis (so ψ0 = −π/2). We set Γ = 1 so that the asteroid rotation rate
matches the orbit period at this distance: 77.2 hours.
The resulting motion of the deputy about the chief is simulated for 8 rotations
of the asteroid. Fig. 5.24 shows the relative motion of the deputy about the chief,
which was placed as close as possible to the marginally stable libration point. It is
clear that the relative motion remains bounded in this case. While there is some
deformation of the relative motion, there is no drift. The model error is shown in
Fig. 5.25. The HCW and C20 -only model errors grow over time with similar error.
The averaged second-order model performs well initially but its error grows with
time. This instance of a body-fixed chief orbit is a case when any averaging in the
kinematics would be best avoided. The unaveraged second-order model successfully
tracks the relative motion without considerable accumulation of error.
106

0.15

0.1

0.05

-0.05

-0.1
0.2
-0.15
0.1
-0.1 0
0 -0.1
-0.2
0.1

Figure 5.24. Relative Orbit (Marginal Stability Case)

0.25

0.2

0.15

0.1

0.05

0
0 1 2 3 4 5 6 7 8

Figure 5.25. Model Error (Marginal Stability Case)


107

6. Conclusions and Suggestions for Future


Work

In this thesis, a linearized relative motion model accounting for J2 = −C20 was
presented and tested, and a new linearized model accounting for the effects of C20
and C22 was derived and tested. The J2 model was numerically tested and validated
with case studies, and the “second-order” model (accounting for all nonzero second
degree and order terms, C20 and C22 ) was studied analytically and tested extensively
numerically. The new second-order model was derived in two forms. The first made
use of averaging in the kinematics and the second avoided all use of averaging. Both
treated the kinematics in a formal manner, describing the motion of the LVLH frame
in terms of the perturbed orbit element angular rates. The averaged second-order
model was most accurate for low-inclination orbits with a high angular rate ratio
Γ = c/n, but its performance otherwise was not always consistent. The unaveraged
second-order model performed well overall in parameter studies, and a special-case
LTI model was used to successfully analyze stability and simulate motion about the
libration points in the body frame of a rotating, gravitating triaxial ellipsoid.
Setting C22 = 0 in the averaged second-order model would result in a model equiv-
alent to the J2 model also discussed in this thesis. Setting C22 = 0 in the unaveraged
second-order model yields an alternate J2 model that completely avoids use of aver-
aging in the kinematics. This model will often have improved performance over the
existing J2 model, especially for high-inclination orbits in which orbit-averaging the
gravitational effects of the equatorial bulge would yield the greatest inaccuracy.
There is a lot of potential for future work with these models, especially the newly
derived second-order model. First, derivation of perturbation solutions for the new
model will enable simulation of the linearized dynamics without numerical integra-
tion. While numerical integration of linearized models is already much faster than
108

numerical integration of the full nonlinear dynamics, a time-explicit solution will


be computationally optimal. Such a solution will be especially useful for terminal
guidance and relative navigation in the highly perturbed orbital environment about
asteroids.
Further analysis of the second-order models may yield new tools for efficiently
designing more stable formations or planning other proximity operations around as-
teroids. It is clear from the simulation results that these models successfully capture
the interesting dynamical phenomenon of relative orbit destabilization due to reso-
nance with periodic gravitational disturbance forces. The simulation results also show
that the effect of this resonance is a considerable and often rapid distortion of relative
motion that would otherwise be non-drifting in the absence of these perturbations.
It is obvious that future multi-spacecraft missions for these environments will have to
actively manage these forces to some degree, but perhaps these new models can aid
in finding formation orbital configurations that minimize the control effort.
Finally, the influence of SRP forces begins to dominate the primary body grav-
itational forces for cases where the primary body is very small or when orbiting
spacecraft have a high area-to-mass ratio. Much more work can be done in deriving
models to develop formations that can attenuate the effects of SRP forces or even
make use of them.
109

References

[1] W. H. Clohessy and R. S. Wiltshire, “Terminal Guidance System for Satellite


Rendezvous,” J. Aerospace Sci. Vol. 27, pp. 653-658, 1960.
[2] K. Johnson, S. R. Vadali, and K. T. Alfriend, “Comparison of Orbit Element
Sets for Modeling Perturbed Satellite Relative Motion,” AAS 16-357, AIAA/AAS
Spaceflight Mechanics Meeting, Napa, CA, February 14-18, 2016.
[3] D. W. Gim and K. T. Alfriend, “State Transition Matrix of Relative Motion for
the Perturbed Noncircular Reference Orbit,” Journal of Guidance, Control, and
Dynamics, Vol. 26(6), pp. 956-971, 2003.
[4] D. Brouwer, “Solution of the Problem of Artificial Satellite Theory Without
Drag,” The Astronomical Journal, Vol. 64, pp. 378-396, 1959.
[5] R. H. Lyddane, “Small Eccentricities or Inclinations in the Brouwer Theory of the
Artificial Satellite,” The Astronomical Journal, Vol. 68, pp. 555-558, 1963.
[6] G. Gaias, J. S. Ardaens, and O. Montebruck, “Model of J2 Perturbed Satellite
Relative Motion with Time-Varying Differential Drag,” Celestial Mechanics and
Dynamical Astronomy, 2015.
[7] B. Mahajan, S. R. Vadali, and K. T. Alfriend, “Analytic Solution for Satellite
Relative Motion: The Complete Zonal Gravitational Problem,” AAS 16-262,
AIAA/AAS Spaceflight Mechanics Meeting, Napa, CA, February 14-18, 2016.
[8] L. Riggi and S. D’Amico, “Optimal Impulsive Closed-Form Control for Spacecraft
Formation Flying and Rendezvous,” American Control Conference, Boston, MA,
July 6-8, 2016.
[9] A. H. Nayfeh, Problems in Perturbation, Wiley, NY, 1985.
[10] A. Biria and R. Russell, “A Satellite Relative Motion Model Including J2 and J3
via Vinti’s Intermediary,” AAS 16-537, AIAA/AAS Spaceflight Mechanics Meet-
ing, Napa, CA, February 14-18, 2016.
[11] H. S. London, “Second Approximation to the Solution of the Rendezvous Equa-
tions,” AIAA Journal, Vol. 1(7), pp. 1691-1693, 1963.
[12] J. P. de Vries, “Elliptic Elements in Terms of Small Increments of Position and
Velocity Components,” AIAA Journal, Vol. 1, pp. 2626-2629, 1963.
[13] M. L. Anthony and F. T. Sasaki, “Rendezvous Problem for Nearly Circular
Orbits,” AIAA Journal, Vol. 3(9), pp. 1666-1673, 1965.
110

[14] E. A. Butcher, T. A. Lovell, and A. Harris, “Third-Order Cartesian Relative


Motion Perturbation Solutions for Slightly Eccentric Chief Orbits,” AAS 16-496,
AIAA/AAS Spaceflight Mechanics Meeting, Napa, CA, February 14-18, 2016.
[15] E. A. Butcher, E. Burnett, and T. A. Lovell, “Comparison of Relative Orbital
Motion Perturbation Solutions in Cartesian and Spherical Coordinates,” AAS 17-
202, AIAA/AAS Spaceflight Mechanics Meeting, San Antonio, TX, February 5-9,
2017.
[16] S. A. Schweighart and R. J. Sedwick, “High-Fidelity Linearized J2 Model for
Satellite Formation Flight,” Journal of Guidance, Control, and Dynamics, Vol.
25(6), pp. 1073-1080, 2002.
[17] S. A. Schweighart and R. J. Sedwick, “Cross-Track Motion of Satellite Forma-
tions in the Presence of J2 Disturbances’,” Journal of Guidance, Control, and
Dynamics, Vol. 28(4), pp. 824-826, 2005.
[18] J. E. Prussing and B. A. Conway, Orbital Mechanics, 2nd ed., Oxford, NY, 2013.
[19] S. Casotto, “The Equations of Relative Motion in the Orbital Reference Frame,”
Celestial Mechanics and Dynamical Astronomy, Vol. 124(3), pp. 215-234, 2016.
[20] E. A. Butcher, E. Burnett, J. Wang, and T. A. Lovell, “A New Time-Explicit
J2-Perturbed Nonlinear Relative Orbit Model with Perturbation Solutions’,” AAS
17-758, AAS/AIAA Astrodynamics Specialist Conference, Columbia River Gorge,
Stevenson, WA, Aug. 20-24, 2017.
[21] D. A. Vallado, Fundamentals of Astrodynamics and Applications, Microcosm
Press, Hawthorne, California, 4th ed., 2013.
[22] R. H. Battin, An Introduction to the Mathematics and Methods of Astrodynamics,
AIAA, Reston, VA, 1999.
[23] H. Schaub and J. L. Junkins, Analytical Mechanics of Space Systems, 3rd ed.,
AIAA, Reston, VA, 2014.
[24] D. J. Scheeres, Orbital Motion in Strongly Perturbed Environments: Applications
to Asteroid, Comet and Planetary Science Orbiters, Springer-Praxis, London, UK,
2012.
[25] Y. Takahashi and D. J. Scheeres, “Small Body Surface Gravity Fields Via Spher-
ical Harmonic Expansions,” Celestial Mechanics and Dynamical Astronomy, Vol.
119(2), pp. 169-206, 2014.
[26] D. J. Scheeres, “Orbital Mechanics About Asteroids and Comets,” Journal of
Guidance, Control, and Dynamics, Vol. 35(3), pp. 987-997, 2012.
111

[27] D. A. Danielson, Vectors and Tensors in Engineering and Physics, 2nd ed, West-
view Press, Cambridge, MA, 2003.

[28] K. T. Alfriend, S. R. Vadali, P. Gurfil, J. P. How, and L. S. Breger, Spacecraft


Formation Flying, Elsevier, Oxford, 2010.

[29] G. Balmino, “Gravitational Potential Harmonics from the Shape of an Homoge-


neous Body”, Celestial Mechanics and Dynamical Astronomy, Vol. 60, pp. 331-
364, 1994.

[30] W. Boyce, “Comment on a Formula for the Gravitational Harmonic Coeffi-


cients of a Triaxial Ellipsoid,” Celestial Mechanics and Dynamical Astronomy,
Vol. 67(2), pp. 107-110, 1997.

[31] W. M. Kaula, Theory of Satellite Geodesy: Applications of Satellites to Geodesy,


Blaisdell Publishing, Waltham, MA, 1966.

[32] V. K. Abalakin, “On Stability of Equilibrium Points of a Triaxial Rotating El-


lipsoid,” Byull. Inst. Teor. Astron. (USSR), Vol. 6(81), p. 543, 1957.

[33] S. G. Zhuralev, “Stability of the Libration Points of a Rotating Triaxial Ellip-


soid,” Celestial Mechanics, Vol. 6(3), pp. 255-267, 1972.

[34] B. Wie, Space Vehicle Dynamics and Control, 2nd ed., AIAA, Reston, VA, 2008.

[35] F. Verhulst, Nonlinear Differential Equations and Dynamical Systems, Springer,


New York, NY, 1996.

[36] W. D. Hu and D. J. Scheeres, “Periodic Orbits in Rotating Second Degree and


Order Gravity Fields,” Chinese Journal of Astronomy and Astrophysics, Vol. 8(1),
pp. 108-118, 2008.

You might also like