orbit design and control of planetary satellite orbiters in the … · 2017-09-21 · orbit design...

186
ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE HILL 3-BODY PROBLEM by Marci Paskowitz Possner A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Aerospace Science) in The University of Michigan 2007 Doctoral Committee: Associate Professor Daniel J. Scheeres, Chair Professor Pierre T. Kabamba Professor N. Harris McClamroch Associate Professor Thomas Zurbuchen

Upload: lekhanh

Post on 09-Jun-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

ORBIT DESIGN AND CONTROL OF

PLANETARY SATELLITE ORBITERS IN THE

HILL 3-BODY PROBLEM

by

Marci Paskowitz Possner

A dissertation submitted in partial fulfillmentof the requirements for the degree of

Doctor of Philosophy(Aerospace Science)

in The University of Michigan2007

Doctoral Committee:

Associate Professor Daniel J. Scheeres, ChairProfessor Pierre T. KabambaProfessor N. Harris McClamrochAssociate Professor Thomas Zurbuchen

Page 2: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Marci Paskowitz Possnerc© 2007

All rights reserved.

Page 3: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

To my husband, Adam Possner,

and my parents, Sharon & Jerry Paskowitz

ii

Page 4: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

ACKNOWLEDGEMENTS

Completing a PhD program, like exploring the farthest reaches of space, is not a

one person endeavor. Many individuals have helped to make this accomplishment

possible.

First and foremost, I’d like to thank my advisor Professor Daniel J. Scheeres.

Professor Scheeres met with me every week during my time at Michigan except while

he was on sabbatical, and that’s only because he was halfway around the world in

Japan. Without his patient guidance and technical insights, this thesis would never

have gotten off the ground. I’d also like to thank the rest of my committee members,

Professors McClamroch, Kabamba and Zurbuchen, for taking the time to evaluate

my work.

Not many graduate students have the luxury of focusing on their research without

also worrying about where their funding is going to come from. The generous

support of the FXB Foundation truly provided the environment that allowed me to

make this endeavor everything that it could be.

I’d also like to thank the members of the Project Prometheus Jupiter Icy Moons

Orbiter Mission Design team at the Jet Propulsion Laboratory. For nearly a year

and a half, I had the pleasure of collaborating with Jon Sims, Lou D’Amario,

John Aiello, Ryan Russell and Try Lam, among others. Furthermore, I’d like to

acknowledge financial support from the Prometheus Project at JPL.

Finally, I’d like to thank my family. My husband, Adam, has been a tremendous

source of strength for me. With his constant love and encouragement, he never let

iii

Page 5: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

me forget that I could finish this dissertation. My parents, Sharon and Jerry, who

always tell me how proud they are of me, and who have supported me in everything

that I’ve done, even when it meant leaving Canada. And my siblings, Erin and

Robbie, who have always supported their big sister.

iv

Page 6: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

PREFACE

The work described in this thesis draws upon the following papers that have been

published, presented at conferences or will be submitted for publication. However,

the organization of the dissertation does not strictly follow the organization of each

paper, as this dissertation gives a full picture of our research.

Journal papers:

• [30] Marci E. Paskowitz & Daniel J. Scheeres, “Robust Capture and Transfer

Trajectories for Planetary Satellite Orbiters,” Journal of Guidance, Control

and Dynamics, 29(2), pp. 342-353, 2006.

• [29] Marci E. Paskowitz & Daniel J. Scheeres, “The Design of Science Orbits

About Planetary Satellites: Application to Europa”, Journal of Guidance,

Control and Dynamics, 29(5), pp. 1147-1158, 2006.

• [32] Marci Paskowitz Possner and Daniel J. Scheeres “Control of Science Orbits

About Planetary Satellites”, in preparation.

Conference papers:

• [26] Marci E. Paskowitz & Daniel J. Scheeres, “Orbit Mechanics About

Planetary Satellites,” paper presented at the 2004 AAS/AIAA Space Flight

Mechanics Meeting, Maui, Hawaii, February 2004, AAS 04-244.

• [25] Marci E. Paskowitz & Daniel J. Scheeres, “Identifying Safe Zones for

Planetary Satellite Orbiters,” paper presented at the 2004 AIAA/AAS

v

Page 7: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Astrodynamics Specialist Conference, Providence, Rhode Island, August 2004,

AIAA 2004-4862.

• [27] Marci E. Paskowitz & Daniel J. Scheeres, “Orbit Mechanics About

Planetary Satellites Including Higher Order Gravity Fields,” paper presented

at the 2005 AAS/AIAA Space Flight Mechanics Meeting, Copper Mountain,

Colorado, 2005, AAS 05-190.

• [28] Marci E. Paskowitz & Daniel J. Scheeres, “Transient Behavior of Planetary

Satellite Orbiters,” paper presented at the 2005 AAS/AIAA Astrodynamics

Specialist Conference, Lake Tahoe, California, August 2005, AAS 05-358.

• [31] Marci E. Paskowitz & Daniel J. Scheeres, “A Toolbox For Designing

Long-Lifetime Orbits About Planetary Satellites: Application to JIMO at

Europa,” invited paper presented at the 2006 AAS/AIAA Space Flight

Mechanics Meeting, Tampa, Florida, 2006, AAS 06-191.

vi

Page 8: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

TABLE OF CONTENTS

DEDICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . iii

PREFACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

CHAPTER

1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Overview of Results . . . . . . . . . . . . . . . . . . . . . . . 41.2 Organization of the Dissertation . . . . . . . . . . . . . . . . 7

2. OVERVIEW OF THE DYNAMICAL SYSTEM . . . . . . . . 92.1 The Three Body and Restricted Three Body Problems . . . . 102.2 The Hill 3-Body Problem . . . . . . . . . . . . . . . . . . . . 132.3 The Modified Hill 3-Body Problem . . . . . . . . . . . . . . . 18

3. FROZEN ORBITS . . . . . . . . . . . . . . . . . . . . . . . . . . 303.1 Reduction of the System by Averaging . . . . . . . . . . . . . 313.2 Frozen Orbits: Equilibrium Solutions of the 1-DOF System . 373.3 Stability of Frozen Orbit Solutions . . . . . . . . . . . . . . . 453.4 Contour Plots and Frozen Orbit Integrations . . . . . . . . . 54

4. ROBUST CAPTURE AND TRANSFER TRAJECTORIES 634.1 Non-dimensional Hill 3-Body Problem and Libration Points . 644.2 Periapsis Poincare Maps . . . . . . . . . . . . . . . . . . . . . 664.3 Symmetry Between Escape and Capture Trajectories . . . . . 744.4 Safe Zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . 784.5 Transferring to Stable Frozen Orbits . . . . . . . . . . . . . . 81

5. THE DESIGN OF LONG LIFETIME SCIENCE ORBITS . 935.1 Dynamics in the 1-DOF System . . . . . . . . . . . . . . . . 955.2 Computing Initial Conditions in the 3-DOF System . . . . . . 1005.3 Long Lifetime Orbits in the 3-DOF System . . . . . . . . . . 1095.4 A Toolbox for Computing Long Lifetime Orbits . . . . . . . . 120

6. CONTROL OF LONG LIFETIME ORBITS . . . . . . . . . . 122

vii

Page 9: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

6.1 Low-Thrust Control . . . . . . . . . . . . . . . . . . . . . . . 1246.2 Resetting a Long Lifetime Orbit . . . . . . . . . . . . . . . . 1356.3 Orbit Uncertainty . . . . . . . . . . . . . . . . . . . . . . . . 1456.4 Long Lifetime Orbits with Initial Errors . . . . . . . . . . . . 149

7. CONCLUSIONS AND FUTURE DIRECTIONS . . . . . . . 1567.1 Summary of the Results . . . . . . . . . . . . . . . . . . . . . 1577.2 Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . 159

APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

viii

Page 10: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

LIST OF FIGURES

2.1 Schematic of the Circular Restricted 3-Body Problem . . . . . . . . 122.2 Schematic of the Hill 3-Body Problem . . . . . . . . . . . . . . . . . 142.3 Trajectory integration in the Modified Hill 3-body problem (Third

degree and order gravity field) . . . . . . . . . . . . . . . . . . . . . 222.4 Trajectory integration using the full ephemeris model, and Europa’s

gravity field up to third degree and order [3]. . . . . . . . . . . . . . 232.5 Trajectory integration in the Modified Hill 3-body problem (truncated

gravity field) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252.6 Trajectory integration in the Elliptic Modified Hill 3-body problem . 29

3.1 A frozen orbit in the 1-DOF system integrated over several orbitalperiods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.2 Frozen orbit solutions for the tide-only case. . . . . . . . . . . . . . 403.3 Comparison of frozen orbit solutions with and without J2. . . . . . . 423.4 Frozen orbit solutions for J2 6= 0 and J3 > 0. . . . . . . . . . . . . . 433.5 Frozen orbit solutions with larger and smaller values of J3. . . . . . 453.6 Stability of frozen orbits about Europa for the tide-only case. . . . . 473.7 Stability of the frozen orbits for the tide plus J2 case. . . . . . . . . 483.8 Stability for frozen orbits about Europa that include the tidal term,

J2 and J3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493.9 Characteristic times for the tide-only unstable frozen orbits with 100

km altitudes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503.10 Characteristic times for unstable frozen orbits about Europa . . . . 513.11 Stability and characteristic times of frozen orbits about Europa with

J3 one order of magnitude smaller than Table 3.1 . . . . . . . . . . . 523.12 Stability and characteristic times of frozen orbits about Europa with

J3 one order of magnitude larger than Table 3.1 . . . . . . . . . . . 533.13 Contour plots for motion in the vicinity of a stable frozen orbit with

e∗ = 0.548, i∗ = 50o, J3 sin ω > 0. . . . . . . . . . . . . . . . . . . . . 553.14 Contour plots for motion in the vicinity of an unstable frozen orbit

with e∗ = 0.0.039, i∗ = 55o, J3 sin ω > 0. . . . . . . . . . . . . . . . . 563.15 Contour plots for motion in the vicinity of a stable frozen orbit with

e∗ = 0.047, i∗ = 30o, J3 sin ω < 0. . . . . . . . . . . . . . . . . . . . . 583.16 Contour plots for motion in the vicinity of an unstable frozen orbit

with e∗ = 0.021, i∗ = 80o, J3 sin ω < 0. . . . . . . . . . . . . . . . . . 59

ix

Page 11: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

3.17 Contour plots and integrations in the 3-DOF of various frozen orbits 62

4.1 Regions of allowable motion and libration points for J = −2.15. . . . 654.2 Periapsis Poincare maps for J = −2.15 (planar case). . . . . . . . . 694.3 Poincare maps for various values of J . . . . . . . . . . . . . . . . . . 714.4 First periapsis passage region for J = −2.15 (3-d case). . . . . . . . 724.5 Inclination as a function of normalized radius for 3-D capture trajec-

tories. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 734.6 Poincare maps showing symmetric capture and escape regions for J =

−2.15. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 754.7 A capture trajectory that escapes after 4 periapsis passages. . . . . . 764.8 Poincare map showing the safe zone, impact and escape regions. . . 794.9 Poincare maps of planar safe zones for various values of J . . . . . . . 794.10 Characteristics of safe capture trajectories for J = −1.70. . . . . . . 814.11 Periapsis passages of capture trajectories for J = −1.70 with altitudes

≤ 250 km. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 834.12 Potential capture trajectories with J = −1.60 for transfers to elliptic

frozen orbits and costs. . . . . . . . . . . . . . . . . . . . . . . . . . 854.13 Contour plots and the numerical results for the orbit obtained after

the transfer maneuver. . . . . . . . . . . . . . . . . . . . . . . . . . 864.14 Contour plots and numerical results for a transfer orbit to a circular

frozen orbit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 884.15 Characteristics of the near-frozen elliptic transfer orbits. . . . . . . . 894.16 Characteristics of both possible circularized orbits over one week. . . 894.17 ω as a function of eccentricity for safe capture trajectories with J =

−1.60. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

5.1 Eccentricity as a function of time for trajectories with different initialvalues. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

5.2 Frozen orbit solutions: unstable near-circular, near-polar orbits indi-cated by heavy dashed line. . . . . . . . . . . . . . . . . . . . . . . . 96

5.3 Stable and unstable manifolds for a frozen orbit, identified with an’x’, with e∗ = 0.0129, i∗ = 70o, ω∗ = −90o and a∗ = 1682.5 km. . . . 98

5.4 Expansion of the central region in 5.3(b) with Europa impact circle . 995.5 3-DOF system integration of a frozen orbit. . . . . . . . . . . . . . . 1005.6 Integration of trajectory initialized on 1-DOF manifold. . . . . . . . 1025.7 Integration of trajectories initialized at xm

0 + δxm0 in 2-DOF system

and xm0 in 1-DOF system. . . . . . . . . . . . . . . . . . . . . . . . . 106

5.8 Integration of trajectories initialized at x0+δx0 in 3-DOF system andx0 in 2-DOF system. . . . . . . . . . . . . . . . . . . . . . . . . . . 109

5.9 Integration in the 3-DOF of a trajectory initialized at x0 with the1-DOF manifolds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

5.10 Time history of the eccentricity for a long lifetime orbit. . . . . . . . 1115.11 1-DOF manifolds with minimum eccentricity circle (dashed curve). . 1125.12 Long lifetime orbits about Europa (70o frozen orbit inclination) . . . 117

x

Page 12: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

5.13 Long lifetime orbit about Enceladus (70o frozen orbit inclination) . . 1195.14 Long lifetime orbit about Dione (70o frozen orbit inclination) . . . . 120

6.1 Time histories of orbital elements integrated in the controlled 1-DOFsystem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.2 Time histories of orbital elements integrated in the uncontrolled 1-DOF system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6.3 Integration of the 3-DOF system controlled with Fr. . . . . . . . . . 1296.4 Manifold in the 1-DOF system and integration of the 3-DOF system

controlled with Fr. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1306.5 Time histories of total thrust magnitude and components of thrust. . 1316.6 Total ∆v used to control the spacecraft using the thrust law given in

Eq.(6.13) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316.7 Integration using constant thrust in the radial direction of −9.25×10−7.1326.8 Integration using constant thrust in the radial direction of−1.045×10−6.1336.9 Eccentricity as a function of time over one orbital period. . . . . . . 1346.10 Time histories of the orbital elements when the transverse thrust is

applied. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1346.11 Diagram of the reset scheme in the 1-DOF system in (e,ω)-space. . . 1376.12 Original (black) and target (red) long lifetime trajectories. . . . . . . 1446.13 Time histories of the orbital elements of the original (black) and target

(red) long lifetime trajectories. . . . . . . . . . . . . . . . . . . . . . 1446.14 Monte Carlo simulation results for various long lifetime trajectories. 1486.15 Schematic of Hohmann transfer between two elliptic orbits. . . . . . 151

xi

Page 13: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

LIST OF TABLES

2.1 Jupiter-Europa System Parameters . . . . . . . . . . . . . . . . . . . 21

3.1 Parameters of Europa . . . . . . . . . . . . . . . . . . . . . . . . . . 41

4.1 Comparison of costs to transfer from a capture trajectory to a circularfrozen orbit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

4.2 Comparison of costs to transfer from capture trajectory to tightlybound circular orbit . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5.1 Algorithm for computing a long lifetime orbit . . . . . . . . . . . . . 1145.2 Long Lifetime Orbits for the Europa System . . . . . . . . . . . . . 1165.3 Long Lifetime Orbits about Enceladus . . . . . . . . . . . . . . . . . 1185.4 Long Lifetime Orbits about Dione . . . . . . . . . . . . . . . . . . . 118

6.1 Examples of Long Lifetime Orbit Reset . . . . . . . . . . . . . . . . 1436.2 Monte Carlo Results for Long Lifetime Orbits . . . . . . . . . . . . . 1476.3 Average cost to correct a long lifetime trajectory after an initial po-

sition error . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1546.4 Average cost to correct a long lifetime trajectory after a larger initial

position error . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

xii

Page 14: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

CHAPTER 1

INTRODUCTION

The exploration of planetary satellites by robotic spacecraft began with the Luna

1 and Pioneer 4 missions flying by the Moon in 1959[2]. Since then, spacecraft

have performed detailed studies of our moon and many other planetary satellites,

including several of the Saturnian moons by the Cassini spacecraft[33] and some

of Jupiter’s moons by the Galileo spacecraft[6]. There is a strong interest in the

scientific community to explore planetary satellites, specifically those that can teach

us more about Earth and the origins of life on Earth. To that end, planetary satellites

of primary interest are those that may have either liquid or frozen water oceans.

Specifically, moons that have ice on the surface may have liquid water underneath

if they are geologically active due to tidal heating caused by their central planet[9].

Examples of icy planetary satellites are Jupiter’s moons Europa, Ganymede and

Callisto and many of Saturn’s moons including Enceladus. Of these, there is the

most pressure from the scientific community to send a spacecraft to Europa since

evidence from the Galileo mission has shown that it has an icy surface covering a

liquid water ocean[39].

Sending a spacecraft to orbit a planetary satellite has not been performed to date,

other than our moon. Doing so is a very challenging problem due to the proximity

of the planetary satellite to the planet. The third body gravitational attraction from

the planet causes large perturbations in the system, and as a result, the two-body

model can not be used. The necessity of including third-body effects without having

1

Page 15: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

a full analytical understanding of them (as is known for the 2-body problem) can

make mission design a difficult and open problem. However, it is possible to use

these perturbations, and hence the natural dynamics of the system, as an asset for

orbit design by working with the perturbations rather than against them.

Spacecraft missions are generally divided into several stages, and missions to

planetary satellites are no exception. Of these, the phase of the mission where

scientific data is acquired is particularly challenging as the orbit is situated in a

highly perturbed environment not usually encountered for the science phases of a

planetary mission. The orbit of the spacecraft for this part of the mission is called

the science orbit. Science orbits have, in general, low altitudes and high inclinations

so that the entire surface can be mapped, and the science requirements of the

mission can be accomplished. These requirements may include imaging the surface,

determining its composition and measuring features of the planetary satellite such

as the gravity field and the radiation environment. Another important requirement

for a science orbit is that it be stable in the sense of having a lifetime long enough

such that a maneuver failure will not cause the satellite to impact with the planetary

satellite. Aside from designing the science orbit, it is also important to consider how

to transfer to that orbit from a region far away from the planetary satellite without

consuming excessive amounts of propellant. The analysis of both of these mission

phases is the focus of this dissertation, with more emphasis placed on the science

orbit design.

Although the planet-planetary satellite-spacecraft system is a three body

problem, the complexity of the system can be significantly reduced by modelling

the system with the Hill 3-body problem. The Hill 3-body problem was originally

derived by G.W. Hill [19] in 1878 to analyze the motion of the moon in the

Sun-Earth-Moon system. The Hill problem is a simplified version of the 3-body

problem in which the masses of two of the bodies are much smaller than the third,

2

Page 16: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

and where the goal is to investigate the motion of the two smaller masses relative

to each other. This is obviously applicable to a planetary satellite system, where

the masses of the planetary satellite and spacecraft are much smaller than the mass

of the planet, and where we wish to evaluate the motion of the spacecraft in the

vicinity of the planetary satellite. One of the main advantages of using the Hill

3-body problem is that it is a much simpler model in terms of analysis than the

3-body and restricted 3-body problems, but still retains the non-integrable dynamics

in the regions to be investigated. Aside from its application to the planet-planetary

satellite-spacecraft system, the Hill problem has many other applications such as

motion in the Earth-Moon-Sun system [35], motion in the vicinity of asteroids [36],

motion about comets [38], periodic orbits [17] and spacecraft maneuvers [41], in

addition to other problems in dynamical astronomy and astrodynamics.

Although the Hill 3-body problem is significantly less complex than other

representations of the 3-body problem, it is still a non-integrable system, which

makes analytic analysis difficult. Since the main goal of this research is to use the

perturbations inherent in the system to our advantage, it is necessary to have a very

good understanding of the underlying dynamics of the system. One method in which

to do this averaging. Specifically, the general Hill 3-body problem can be reduced to

an integrable system by averaging over the orbit of the spacecraft and the orbit of

the planetary satellite (or their equivalents if another system is being investigated).

The averaged system provides some insight into the natural dynamics of the full

system and directs the design of long lifetime science orbits in this dissertation. The

technique of averaging has been applied to planetary satellite systems [37, 34, 22],

to general systems [7], to the motion of asteroids [21, 23] and to stellar systems [16],

to name a few.

The Hill problem is also very useful for the investigation of the regions farther

away from the planetary satellite, in terms of planning the transfer of the spacecraft

3

Page 17: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

to the science orbit. Since for this scenario, the spacecraft is still much closer to

the planetary satellite than to the planet, the Hill approximation remains valid.

The types of trajectories studied are those that come from outside the Hill region

and circle the planetary satellite at least once. These trajectories are called capture

trajectories, and their analog, escape trajectories, have been studied in [42, 41].

Capture trajectories exist not only for spacecraft orbiting planetary satellites but also

in other systems where the Hill model applies. One example is the irregular moons

of Jupiter and Saturn which might have become captured due to this effect [5].

1.1 Overview of Results

The results in this dissertation all use the Hill 3-body problem as a model. The

primary motivation is a future Europa orbiter mission, but the results are general

enough to be applied to many different planetary satellites. The two main problems

that are investigated are the design of science orbits and the design of capture

and transfer trajectories. To accomplish both of these tasks, a thorough study of

the dynamics of the system is undertaken. This begins with averaging, where the

system, which has three degrees of freedom (DOF) is averaged twice, and reduced

to an integrable 1-DOF system. The averaging takes place over the orbit of the

spacecraft about the planetary satellite and over the orbit of the planetary satellite

about the planet. The averaging and subsequent analysis is performed for the basic

(tide-only) Hill 3-body problem as well as for a modified Hill 3-body problem which

includes terms from the gravity field of the planetary satellite.

In the 1-DOF system, equilibrium solutions exist which are called frozen orbits.

These orbits have constant values of the orbital elements on average. An analysis of

the frozen orbit solutions shows that both stable and unstable frozen orbits exist,

and both types have potential applications for the goals of this research. Adding the

planetary satellite gravity field to the model slightly changes the structure of the

4

Page 18: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

frozen orbit solutions, but the qualitative features of the near-polar frozen orbits

remain similar. Specifically, including the J3 term of the planetary satellite gravity

field causes what were circular, near-polar frozen orbits to become very slightly

elliptical. This has implications for science orbit design since the motion occurs very

close to the planetary satellite, but shows that when planning capture and transfer

trajectories, using the tide-only model is sufficient. Furthermore, it is shown that

changing the value of J3 does not affect the qualitative frozen orbit results. This

is important since an accurate value (magnitude or sign) for J3 for Europa is not

currently available. The validity of the 1-DOF system as an approximation for

motion in the 3-DOF system is verified by performing numerical simulations in the

3-DOF system and comparing the resulting motion to plots of the 1-DOF system

motion. It is found that the 1-DOF system provides a very good approximation to

the 3-DOF system, especially when the altitude of the orbit remains relatively small.

To study capture and transfer trajectories for planetary satellite orbiters, the

dynamics are reduced to a discrete map represented by periapsis passages. This

allows us to study the trajectories that enter the Hill region and orbit the planetary

satellite. Specifically, trajectories that orbit the planetary satellite for at least one

week are identified and the regions in which their periapsis passages exist are called

safe zones. These safe zones represent potential trajectories for use in a mission to

transfer the spacecraft to the vicinity of the planetary satellite. In particular, there

are trajectories that exist in the safe zones that are especially useful as gateways

to transfers to a science orbit with relatively low fuel usage. The science orbits

considered in this analysis are circular and elliptic frozen orbits.

The frozen orbits in the 1-DOF system also provide the basis for our science

orbit design. Specifically, the low-altitude, near-polar unstable frozen orbits in the

modified Hill 3-body problem have stable and unstable manifolds in the 1-DOF

system. By designing orbits such that they follow the stable to unstable manifold

5

Page 19: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

path, it is possible to extend the lifetime of trajectories significantly compared to

simply using frozen orbits. The challenge for this method is determining how to

choose initial conditions such that the trajectory in the 3-DOF system follows the

1-DOF manifolds. This is accomplished via a linearization algorithm that relates

the 1-DOF, 2-DOF and 3-DOF systems to each other in order to predict the

corrections necessary to the 1-DOF manifold solutions. The resulting trajectories

have sufficiently long lifetimes to be very practical as science orbits for planetary

satellite orbits.

Once science orbits have been designed, some aspects of controlling them must

be considered. First, the effectiveness of using constant thrust to maintain the orbit

is studied, and found to be too costly. Then, a scheme is developed for a low cost

transfer consisting of two maneuvers that restarts a science orbit trajectory before it

impacts with the planetary satellite. The effect of orbit uncertainty on science orbit

design is also investigated to determine the sensitivity of the science orbits to initial

errors. Finally, given an initial position error, a technique is developed to return the

spacecraft with minimal fuel usage to its nominal trajectory.

In summary, the main contributions of this thesis are:

• The analysis of the 1-DOF system dynamics

• The analysis of capture trajectories in safe zones

• The design method for long lifetime science orbits

• The optimal transfer strategy for restarting a long lifetime orbit

• The control strategy to correct an initial position error

6

Page 20: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

1.2 Organization of the Dissertation

This dissertation begins with a discussion of the dynamical model in Chapter II. The

Hill 3-body problem is derived from the 3-body problem, showing that the motion of

a planetary satellite orbiter can be modeled by this simpler problem. Additionally,

the planetary satellite’s gravity field is included in the model since its effect on the

motion is shown to be significant.

In Chapter III, frozen orbit solutions are derived and their structure is

investigated. Comparisons are made for frozen orbit solutions in models with and

without the planetary satellite gravity field, as well as for different values of J3.

Their applicability to the 3-DOF system is also examined.

Chapter IV deals with trajectories that come from outside the Hill region to the

vicinity of the planetary satellite. These are studied using Periapsis Poincare maps,

and sets of trajectories viable as capture and transfer trajectories for a planetary

satellite orbiter are identified.

The design of science orbits is undertaken in Chapter V, where the structure

of the 1-DOF system is used as a basis for the design. The desired trajectories

are identified in the 1-DOF system, and then an algorithm is developed whereby

these trajectories are translated into the 3-DOF system. Several examples of long

lifetime science orbits are presented for the Europa system, and for the Saturnian

moons Enceladus and Dione. Finally, a toolbox to compute the initial conditions

for long lifetime science orbits that was written and delivered to the Jet Propulsion

Laboratory for the Jupiter Icy Moons Orbiter Mission is described.

Chapter VI looks at the control of long lifetime science orbits by both resetting

ideal trajectories before they impact and examining the effect of orbit uncertainty

on the orbit design. Monte Carlo simulations are used to determine how science

orbits are impacted by initial position errors, and a scheme is developed to return

the spacecraft to its nominal trajectory after an initial position error.

7

Page 21: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Finally, a summary and possible future research directions are presented in

Chapter VII.

8

Page 22: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

CHAPTER 2

OVERVIEW OF THE DYNAMICAL SYSTEM

The three body problem is the most general way to describe the motion of

three masses under their mutual gravitation attraction. It can be used to analyze

the motion of any system of three bodies such as the Sun, Earth and Moon, a

spacecraft orbiting a binary asteroid system, or our application to a spacecraft in

orbit about a planetary satellite. There are special cases of the three body problem

in which certain assumptions can be made that simplify the analysis of the system.

We will first briefly discuss the Restricted Three Body Problem which deals with

the case of a small mass attracted by 2 point masses revolving around each other.

This simplification can be used for the analysis of the motion of a spacecraft in the

vicinity of two bodies such as planets, a planet and the sun, two asteroids, or a

planet and a planetary satellite.

A further simplification can be made for the case where two of the three masses

are close together and small compared to the third one. This is called the Hill

3-body problem. For the general Hill 3-body problem, it is not necessary to assume

that one of the two smaller masses is extremely small, such as a spacecraft in the

Restricted 3-body problem [18]. However, since our application is to the motion of

a spacecraft in orbit about a planetary satellite, we consider what is formally called

the Restricted Hill 3-body problem. For simplicity, we refer to it as the Hill 3-body

problem.

All of our analysis in this dissertation is performed using the Hill 3-body problem

9

Page 23: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

as the model. However, to account for the shape of the planetary satellite, we

develop a modified form of Hill’s 3-body problem that includes some of the gravity

field coefficients of the planetary satellite. The question then arises of which gravity

field terms to include. This is a difficult question since, for our main focus of Europa,

accurate values of most of the gravity field coefficients are not known. When gravity

field coefficients are included in this research, those included are J2, C22 and J3.

These are chosen since they have a large effect of the motion of the spacecraft. This

choice is justified by the demonstration that if the full third degree and order gravity

field is included, the differences in the motion are not significant. We also justify the

use of the Hill model by comparing it to results generated using the full ephemeris

model. The Hill approximation provides results that are close to those using the

full ephemeris model, but it allows for analysis that would not be possible in the

ephemeris model.

The Hill model used in this research assumes a circular orbit for the planetary

satellite. The validity of this assumption is verified for Europa by simulating the

elliptic Hill 3-body problem. It is found that the results from the two models are

not substantially different. The conclusion is therefore that using a modified form

of the Hill 3-body problem where the J2, C22 and J3 gravity field coefficients are

included gives strong results that give a good picture of the motion and can be used

for mission design.

2.1 The Three Body and Restricted Three Body

Problems

The general 3-body problem is used to describe the motion of three bodies under

their mutual gravitational attraction. The equation of motion for each body is given

by[12]:

Miri = −GMi

3∑j=1,j 6=i

Mj(ri − rj)

|ri − rj|3, (2.1)

10

Page 24: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

where G = 6.673×10−11m3/kgs2 is the universal gravitational constant and i = 0−2.

It is convenient to express the equations of motion in force potential notation, where

the force potential U is[12]:

U =1

2

3∑i=1

3∑j=1,j 6=i

GMiMj

|rij|, (2.2)

and rij = rj − ri. Then, the equations of motion for k = 0− 2 are

Mkrk =∂U

∂rk

. (2.3)

The restricted three body problem describes the dynamics of a small mass

such as a spacecraft attracted by two point masses rotating about each other. We

concentrate on the circular restricted three body problem in which the two larger

masses are in a circular orbit about each other. Let P0 and P1 be the larger masses

and P2 the smaller mass (spacecraft), so that M2 = 0. Let the origin of this system

be the center of mass of the P0, P1 system (see Figure 2.1). Since we are focussing

on a planet/planetary satellite system in this dissertation, let P0 be the planetary

satellite and P1 the planet. Then, R is the vector from the planetary satellite to the

planet and r is the position of the spacecraft. Since the planetary satellite is in a

circular orbit about the planet, |R| is constant. We focus only on the motion of the

spacecraft since that is our primary interest here. Then,

r =∂U

∂r, (2.4)

where

U =GM0

|r + RM1/(M0 + M1)|+

GM1

|r−RM0/(M0 + M1). (2.5)

Let N =√

G(M0 + M1)/R3 be the mean motion of the planetary satellite,

and note that it is also a constant. Now, express the equations of motion of the

spacecraft in the frame rotating about the planetary satellite. Let rI be the position

of the spacecraft in the inertial frame and rr be the position of the spacecraft in

11

Page 25: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Figure 2.1: Schematic of the Circular Restricted 3-Body Problem

the rotating frame. The rotation of the frame is expressed as Ω = N z. Since both

frames have the same origin, rI = rr. The position and velocity in the rotating frame

are [14]:

rI = rr + Ω× rr , (2.6)

rI = rr + 2Ω× rr + Ω×Ω× rr . (2.7)

Since rI = rr, the subscripts on r can be removed. Then, substituting for Ω and

equating Eqs.(2.4) and (2.7), we obtain:

r + 2N z + N2z× z× r =∂U

∂r. (2.8)

This can be broken down into equations for the x, y and z positions of the spacecraft:

x− 2Ny −N2x =∂U

∂x, (2.9)

y + 2Nx−N2y =∂U

∂y, (2.10)

z =∂U

∂z, (2.11)

12

Page 26: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

where U from Eq.(2.5) expressed in coordinate form is:

U =GM0√

(x + M1/(M0 + M1)R)2 + y2 + z2

+GM1√

(x−M0/(M0 + M1)R)2 + y2 + z2. (2.12)

To simplify the equations of motion, let

V =1

2N2(x2 + y2) + U . (2.13)

Then, Eqs.(2.9)-(2.11) can be simplified to:

x−Ny =∂V

∂x, (2.14)

y + Nx =∂V

∂y, (2.15)

z =∂V

∂z. (2.16)

Eqs.(2.14)-(2.16) are the equations of motion for a spacecraft in the Circular

Restricted 3-Body Problem.

2.2 The Hill 3-Body Problem

The conditions of the Hill 3-body problem are that the mass of the planet, M0,

dominates the system and that the planetary satellite and spacecraft are close to

each other and their center of mass is far away from the planet[35]. As previously

mentioned, for simplicity, we consider only the restricted Hill 3-body problem, in

which the mass of the spacecraft tends to zero since that is the specific problem

being addressed in this dissertation. In other words, we focus on the situation where

the motion of the spacecraft is close to the planetary satellite and the mass of the

planetary satellite is much smaller than the mass of the planet (see Figure 2.2).

First, let

µ =M1

M0 + M1

. (2.17)

13

Page 27: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Figure 2.2: Schematic of the Hill 3-Body Problem

The assumption is then that µ 1. Next, non-dimensionalize the system using the

length scale R and the time scale 1/N . Then, x = x/R, y = y/R, z = z/R and

t = Nt. Substituting these in Eqs.(2.14)-(2.16), the non-dimensional equations of

motion for the Circular Restricted 3-Body Problem are:

¨x− 2y =∂V

∂x, (2.18)

¨y + 2x =∂V

∂y, (2.19)

¨z =∂V

∂z, (2.20)

where

V =1

2(x2 + y2) +

1− µ√(x + µ)2 + y2 + z2

+µ√

(x− 1 + µ)2 + y2 + z2. (2.21)

Next, consider motion in the vicinity of the planetary satellite, and shift the

origin to the planetary satellite. This entails setting x = 1 − µ − x where x is the

14

Page 28: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

new x-coordinate. Then, the equations of motion are:

¨x− 2y =∂V

∂x, (2.22)

¨y + 2x =∂V

∂y, (2.23)

¨z =∂V

∂z, (2.24)

where

V =1

2((1− µ− x)2 + y2) +

1− µ√(1− x)2 + y2 + z2

+µ√

x2 + y2 + z2. (2.25)

The condition that µ 1 is made explicit by the following change of coordinates[35]:

x = µ1/3x , (2.26)

y = µ1/3y , (2.27)

z = µ1/3z . (2.28)

Note that x, y and z are now the coordinates of the Hill 3-Body Problem, with

the following equations of motion in the rotating frame, centered on the planetary

satellite:

x− 2y =1

µ2/3

∂V

∂x, (2.29)

y + 2x =1

µ2/3

∂V

∂y, (2.30)

z =1

µ2/3

∂V

∂z, (2.31)

where we have dropped the ∼ on V. The remaining step is to expand the equations of

motion in powers of µ1/3, ignoring orders of µ2/3 and higher. This entails expanding

15

Page 29: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

1/µ2/3V as follows:

1

µ2/3V =

1

2µ2/3

[(1− µ + µ1/3x)2 + µ2/3y2

]+

1− µ

µ2/3√

µ2/3r2 + 1 + 2µ1/3x+

1

r

=1

2µ2/3− µ1/3 +

µ4/3

2+

1

2(x2 + y2) +

1

r+

x

µ2/3− µ2/3x (2.32)

+1− µ

µ2/3√

1 + 2µ1/3x + µ2/3r2,

where r =√

x2 + y2 + z2. The term in the above expression to be expanded is:

1

µ2/3√

1 + 2µ1/3x + µ2/3r2. (2.33)

This is accomplished using the following Taylor expansion:

1√1 + ε

= 1− 1

2ε +

3

8ε2 + · · · (2.34)

Therefore,

1

µ2/3√

1 + 2µ1/3x + µ2/3r2=

1

µ2/3− x

µ1/3− 1

2r2 + 3x2

+3

2µ1/3xr2 + O(µ2/3) . (2.35)

Then, substituting into Eq.(2.32):

V = const +1

r+

3

2x2 − 1

2z2 + O(µ2/3) . (2.36)

Finally, since µ1/3 1, the orders of µ2/3 are ignored and since the equations

of motion use the derivatives of V the constant terms can also be ignored. The

potential for the Hill problem is then

VH =1

r+

3

2x2 − 1

2z2 . (2.37)

It is convenient to divide the potential for the Hill problem into two components: the

part due to the planetary satellite and the part due to the planet, which is called the

perturbing potential and denoted R. In addition, for the majority of this dissertation

16

Page 30: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

the dimensional form of the potential is used, and so we re-dimensionalize here:

VH =µ

r+

1

2N2(x2 + y2) + R . (2.38)

with

R = −1

2N2r2 +

3

2N2x2 . (2.39)

Observe that substituting R from Eq.(2.39) into Eq.(2.38) yields the simpler form of

VH that combines the planetary satellite and the planet components:

VH =µ

r+

1

2N2(x2 + y2)− 1

2N2r2 +

3

2N2x2

r+

1

2N2(x2 + y2)− 1

2N2(x2 + y2 + z2) +

3

2N2x2

r+

3

2N2x2 − 1

2N2r2 .

Then, the general equations of motion for the Hill 3-body problem are:

x− 2Ny = − µ

r3x + N2x +

∂R

∂x, (2.40)

y + 2Nx = − µ

r3y + N2y +

∂R

∂y, (2.41)

z = − µ

r3z +

∂R

∂z. (2.42)

The Hill 3-Body Problem has the Jacobi Integral, which can be derived by

multiplying Eq.(2.40)-(2.42) by x, y and z respectively:

xx + yy + zz − 2Nxy + 2Nxy = − µ

r3(xx + yy + zz) + N2(xx− yy)

+∂R

∂xx +

∂R

∂yy +

∂R

∂zz . (2.43)

This is equivalent to

d

dt

[1

2(x2 + y2 + z2)

]=

d

dt

[1

r+ N2(x2 + y2) + R

], (2.44)

17

Page 31: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

which can be expressed as:

d

dt

[1

2(x2 + y2 + z2)− 1

r−N2(x2 + y2)−R

]= 0 . (2.45)

Therefore, the Jacobi Integral, which is constant, is:

J =1

2(x2 + y2 + z2)− 1

r−N2(x2 + y2)−R . (2.46)

2.3 The Modified Hill 3-Body Problem

The modified Hill 3-body problem is the general Hill 3-body problem with some

additional system parameters included. The first logical additions are the higher

order terms of the gravity field of the planetary satellite. It is especially interesting

to consider the effect of including these terms when considering motion very close

to the planetary satellite. In general, for a particular planetary satellite under

consideration, only some of the gravity field terms will be known. In addition, even

if the values of many of the gravity field terms are available, including them in our

analysis may not change the results, or may make the analysis difficult. Another

possible modification to the Hill 3-body problem is removing the assumption that

the planetary satellite is in a circular orbit about the planet by constructing a model

that includes the ellipticity in the planetary satellite’s orbit.

2.3.1 Planetary Satellite Gravity Field

Including gravity field terms into the model simply entails adding them to the

perturbing potential as follows[24]:

R = −N2

2r2 +

3

2N2x2 +

µ

r

∞∑n=2

n∑m=0

Rns

rnPnm(sin φ)(Cnm cos mλ + Snm sin mλ) , (2.47)

18

Page 32: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

where

x = r cos φ cos λ ,

y = r cos φ sin λ ,

z = r sin φ .

Rs is the radius of the planetary satellite, n is the degree and m is the order.

Pnm(sin φ) are the Legendre polynomials and the Cnm and Snm are coefficients that

describe the dependence on the internal mass distribution of the planetary satellite.

In addition, the coefficients Cn0 are generally denoted as −Jn. The equations of

motion for the model that includes the planetary satellite gravity field are:

x + 2Ny = − µ

r3x + N2x +

∂R

∂x, (2.48)

y − 2Nx = − µ

r3y + N2y +

∂R

∂y, (2.49)

z = − µ

r3z +

∂R

∂z. (2.50)

Obviously, it is not possible to include all of the gravity field terms for the

planetary satellite under consideration. First of all, the Cnm and Snm coefficients are

determined by analyzing data returned from missions to these planetary satellites.

This means that only a few gravity field coefficient values are available. Secondly,

the more gravity field terms included, the more difficult is the analysis. When it

comes to integrating the equations of motion to determine the path of a trajectory,

any number of gravity field terms can be included. However, for analysis such as

that performed in this dissertation, it is necessary to restrict the inclusion of gravity

field terms to those that have the largest effect on the motion.

An important consideration in this research is what the effect of using the Hill

model has on the motion of the spacecraft, as compared to using a full ephemeris

model which includes all of the physics in the system. It is necessary to verify

19

Page 33: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

that the results obtained for the Hill model with fewer of the gravity field terms

are reasonable approximations. For these comparisons, we use the Jupiter-Europa

system. It is important to note that the Europa gravity field is only known for

certain up to the second degree and order [4]. The third degree values were obtained

from [3] and are very rough estimates. All of the values used in this analysis are

in Table 2.1. For this comparison, a trajectory is integrated in the modified Hill

3-Body Problem, with Europa’s third degree and order gravity field included. It is

compared to the integration in a full ephemeris model with the same Europa gravity

field of a trajectory with the same initial conditions. The time histories of the

orbital elements for both cases are shown in Figures 2.3 and 2.4, respectively. The

results in Figure 2.4 were obtained from John Aiello [3], who generated the plots

using DPTRAJ at the Jet Propulsion Laboratory. It is clear that the qualitative

features of the plots are a very good match. The periods of all of the oscillations in

the plots are identical, and the motion of all of the orbital elements agree very well.

Therefore, using a modified form of the Hill 3-body problem provides good results

and is justified for use in our analysis.

Although the modified Hill 3-body problem with the Europa third degree and

order gravity field adequately approximates motion in the vicinity of Europa, it is

not practical for analysis to include that many gravity field coefficients. For Europa,

which is the primarily application of our results in this dissertation, the J2, C22 and

J3 contributions have been shown to have a strong effect on the motion [27], and are

included in the majority of our analysis. The components of the perturbing potential

corresponding to J2, C22 and J3 are:

RJ2 = −µJ2

2

(3z2

r5− 1

r3

), (2.51)

RC22 =3µC22

r5(x2 − y2) , (2.52)

RJ3 = −µJ3

2

(5z3

r7− 3z

r5

). (2.53)

20

Page 34: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 2.1: Jupiter-Europa System Parameters

Parameter Symbol Value

Jupiter gravitational parameter µp 1.267× 108 km3/s2

Europa gravitational parameter µ 3201 km3/s2

Europa orbit rate N 2.05× 10−5 rad/sEuropa orbital period TE 3.55 daysEuropa eccentricity es 0.01Europa semi-major axis as 671100 kmEuropa radius Rs 1560.8 kmEuropa J2 J2 1041.39 km2

Europa C21 C21 −0.32 km2

Europa S21 S21 −20.47 km2

Europa C22 C22 312.97 km2

Europa S22 S22 −4.71 km2

Europa J3 J3 524117.32 km3

Europa C31 C31 120230.75 km3

Europa S31 S31 81414.49 km3

Europa C32 C32 −11938.14 km3

Europa S32 S32 −3617.07 km3

Europa C33 C33 −1765.93 km3

Europa S33 S33 3669.41 km3

21

Page 35: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 20 40 60 80 100 1201550

1600

1650

1700

1750

1800ra

dius

(km

)

time (days)

0 20 40 60 80 100 1201660

1661

1662

1663

1664

1665

1666

1667

1668

sem

i−m

ajor

axi

s (k

m)

time (days)

0 20 40 60 80 100 1200

0.01

0.02

0.03

0.04

0.05

0.06

0.07

ecce

ntric

ity

time (days) 0 20 40 60 80 100 120

109.5

110

110.5

111

111.5

112

incl

inat

ion

(deg

)

time (days)

0 20 40 60 80 100 120−120

−100

−80

−60

−40

−20

0

20

40

ω (

deg)

time (days)

0 20 40 60 80 100 120−200

−150

−100

−50

0

50

100

150

200

Ω (

deg)

time (days)

Figure 2.3: Trajectory integration in the Modified Hill 3-body problem (Third degree and ordergravity field)

22

Page 36: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Figure 2.4: Trajectory integration using the full ephemeris model, and Europa’s gravity field up tothird degree and order [3].

23

Page 37: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

where the values of J2, C22 and J3 have been dimensionalized by their multiplications

by R2s and R3

s, respectively. Therefore, we verify that including only the J2, C22

and J3 coefficients does not drastically affect the results. In a real mission design

situation, analysis would be performed using this reduced gravity field and a

trajectory valid in the Hill problem would be computed. This trajectory could then

be used as a starting point for designing a trajectory in the full ephemeris model.

In examining results obtained using a truncated Europa gravity field, it is not

necessary to find the exact same behavior of the trajectory, especially in terms of

its lifetime. The environment around Europa in particular is highly perturbed, and

a slight change in the initial conditions can result in a drastic change in the overall

trajectory. Therefore, we are only interested in verifying that using a model with a

truncated gravity field captures the important features of the motion. Comparing

Figure 2.5 which shows the integration of a trajectory in the modified Hill 3-body

problem with the gravity field truncated to J2, C22 and J3 to Figure 2.3 which

shows an integration using the third degree and order gravity field, we see that the

overall features are very similar. Specifically, the amplitudes of the oscillations in

inclination and semi-major axis are the same in both cases, as are the trends in the

motion of all of the orbital elements. The main difference between the two cases

is the amplitude of the eccentricity and argument of periapsis oscillations. In the

third degree and order gravity field case, the size of the oscillations are larger for

both orbital elements. However, since the average motion remains the same between

the full gravity field and truncated gravity field case, the results obtained from the

truncated gravity field case are sufficient for our purposes.

2.3.2 Planetary Satellite Eccentricity

In this section the assumption that the planetary satellite is in a circular orbit about

the planet is verified. To do so, an elliptic planetary satellite orbit is added to the

24

Page 38: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 20 40 60 80 100 1201550

1600

1650

1700

1750

1800ra

dius

(km

)

time (days)0 20 40 60 80 100 120

1661

1662

1663

1664

1665

1666

1667

1668

sem

i−m

ajor

axi

s (k

m)

time (days)

0 20 40 60 80 100 1200

0.01

0.02

0.03

0.04

0.05

0.06

0.07

ecce

ntric

ity

time (days)0 20 40 60 80 100 120

109.8

110

110.2

110.4

110.6

110.8

111

111.2

111.4

111.6

111.8

incl

inat

ion

(deg

)

time (days)

0 20 40 60 80 100 120−100

−80

−60

−40

−20

0

20

40

ω (

deg)

time (days)0 20 40 60 80 100 120

−200

−150

−100

−50

0

50

100

150

200

Ω (

deg)

time (days)

Figure 2.5: Trajectory integration in the Hill 3-body problem, with the Europa gravity fieldtruncated to include only the J2, C22 and J3 contributions.

25

Page 39: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

model. In the circular model, the assumption that the rotation rate of the planetary

satellite is synchronous with its orbit rate was made. Therefore, when an elliptic

planetary satellite orbit is introduced, it is important to ensure that the rotation

rate of the planetary satellite is still constant, and does not vary as the speed of

the planetary satellite on its orbit varies. For these comparisons, the modified

Hill 3-body problem includes the J2, C22 and J3 planetary satellite gravity field

components.

The equations of motion for the elliptic Hill 3-body problem are most easily

expressed in non-dimensional coordinates. The potential of the non-dimensional

circular Hill problem is:

VC =1

r+

1

2(x2 + y2) + RC , (2.54)

where

RC = −1

2r2+

3

2x2− J2

2ε2d2

(3z2

r5− 1

r3

)+

3C22

ε2d2r5(x2−y2)− J3

2ε2d2

(5z3

r7− 3z

r5

). (2.55)

The factors used to remove the dimensions of the system are ε and the radius of the

planetary satellite in orbit about the planet, d:

ε = (µ/µp)1/3 , (2.56)

d =as(1− e2

s)

1 + es cos fs

, (2.57)

where µp is the gravitational parameter of the planet, as is the semi-major axis

of the planetary satellite, es is the eccentricity of the planetary satellite and fs is

the true anomaly of the planetary satellite. Then, the potential of the elliptic Hill

3-body problem is[13]:

VE =1

1 + es cos fs

VC . (2.58)

One final correction to the potential must be made, to ensure the constant rotation

rate of the planetary satellite. This correction only applies to the C22 component of

26

Page 40: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

the potential since it is the only non-symmetric component. Specifically, the x and

y values in the C22 component in Eq.(2.55) are replaced by x′ and y′, where x′

y′

=

cos (Ms − fs) sin (Ms − fs)

sin (Ms − fs) cos (Ms − fs)

x

y

, (2.59)

and Ms is the mean anomaly of the planetary satellite. Then, the equations of

motion for the elliptic modified Hill 3-body problem are:

x = 2y +∂VEH

∂x, (2.60)

y = −2x +∂VEH

∂y, (2.61)

z =∂VEH

∂z, (2.62)

where

VEH =1

1 + es cos fs

[1

r+

1

2(x2 + y2) + REH

], (2.63)

and

REH = −1

2r2 +

3

2x2 − J2

2ε2d2

(3z2

r5− 1

r3

)+

3C22

ε2d2r5(x′

2 − y′2)

− J3

2ε2d2

(5z3

r7− 3z

r5

). (2.64)

Figure 2.6 shows the results of the integration of a trajectory in the elliptic

modified Hill 3-body problem. The parameters used are those for a spacecraft in

orbit about Europa, and are given in Table 2.1. Compared to Figure 2.5, it is clear

that the main features of the motion are the same. The orbital elements show the

same trends, and the sizes of the oscillations are the same. The major difference

is that in the elliptic case, the semi-major axis shows a slow decrease. This is

not a large concern, since the overall decrease is only about 2 km over 130 days.

Ignoring this effect in the circular case will not strongly affect the overall analysis.

Therefore, for ease of analysis, the Circular Modified Hill 3-body problem is used for

27

Page 41: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

the remainder of this dissertation.

28

Page 42: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 20 40 60 80 100 120 1401560

1580

1600

1620

1640

1660

1680

1700

1720

1740

1760ra

dius

(km

)

time (days)0 20 40 60 80 100 120 140

1656

1658

1660

1662

1664

1666

1668

sem

i−m

ajor

axi

s (k

m)

time (days)

0 20 40 60 80 100 120 1400

0.01

0.02

0.03

0.04

0.05

0.06

0.07

ecce

ntric

ity

time (days)0 20 40 60 80 100 120 140

109

109.5

110

110.5

111

111.5

incl

inat

ion

(deg

)

time (days)

0 20 40 60 80 100 120 140−120

−100

−80

−60

−40

−20

0

20

40

ω (

deg)

time (days)0 20 40 60 80 100 120 140

−60

−40

−20

0

20

40

60

80

100

120

140

Ω (

deg)

time (days)

Figure 2.6: Trajectory integration in the Elliptic Modified Hill 3-body problem, with the Europagravity field truncated to include only the J2, C22 and J3 contributions.

29

Page 43: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

CHAPTER 3

FROZEN ORBITS

The Hill 3-body problem is very complex in that it is a 3 degree of freedom (DOF)

non-integrable system. Detailed analysis of it in that form is only possible through

numerical simulations and it is difficult to obtain insight into the system except for

some special situations such as equilibrium points. Given that the ultimate goal of

the study of planetary satellite orbiters is the design of useful orbits for a mission,

it is important to learn more than what can be gained from numerical simulations.

This is accomplished with the analysis of a reduced system. By averaging over the

motion of the spacecraft about the planetary satellite and over the motion of the

planetary satellite about the planet, the 3-DOF system can be reduced to a 1-DOF

integrable system. This 1-DOF system has equilibrium solutions, denoted as frozen

orbits, which have constant values of the eccentricity, inclination and argument of

periapsis, on average. We compute frozen orbit solutions for three models: tide-only,

tide + J2 and tide + J2 + J3 and characterize the differences between them.

Previous work using averaging methods has studied the dynamics of orbits

about planetary satellites and frozen orbits. Broucke [7] studied orbit dynamics in

a model that included the tidal force. He found elliptic and circular frozen orbits

and investigated their stability. Scheeres, et al. [37] studied near-circular frozen

orbits in a model that included both the tidal force and J2. They investigated the

stability of these orbits and developed an analytic theory to postpone impact with

the surface for significant periods of time. In addition San-Juan et al. [34] and Lara

30

Page 44: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

et al. [22] studied orbit dynamics about planetary satellites using a more rigorous

averaging method that involved Lie transforms to develop averaged equations to a

higher order. They included the tidal term and J2 and the tidal term, J2 and C22,

respectively.

Aside from computing the frozen orbit solutions, we also examine their stability

since some frozen orbits are stable and others are unstable. The concept of the

characteristic time is also introduced. Just as the distribution of the frozen orbit

solutions in orbit element space is affected by the inclusion of the gravity field

coefficients, so too is their stability. The stability of the frozen orbit solutions is

therefore computed for the three different cases.

Contour plots are a way to study the qualitative secular motion in the vicinity

of frozen orbits. The motion in the vicinity of the different types of frozen orbits

is shown using contour plots in the 1-DOF system and integrations in the 3-DOF

system. The integration of frozen orbit solutions in the 3-DOF system is used to

determine whether the stable frozen orbits are viable as trajectories in the 3-DOF

system and to show that unstable frozen orbits in the 1-DOF system are also

unstable in the 3-DOF system.

3.1 Reduction of the System by Averaging

Recall the perturbing potential of the 3-DOF system:

R =1

2N2(3x2 − r2

)− µJ2

2

(3z2

r5− 1

r3

)+

3µC22

r5(x2−y2)− µJ3

2

(5z3

r7− 3z

r5

). (3.1)

Before presenting the results of the reduction of the system, it is necessary to

introduce three regimes other than Cartesian coordinates in which the motion can

be analyzed. Orbital elements are a set of six values used frequently in celestial

mechanics and astrodynamics to describe the size, shape and orientation of an orbit.

Semi-equinoctial elements are useful for dealing with near-circular orbits. Delaunay

31

Page 45: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

variables are a canonical set of six elements of which three are angles and three are

momenta. The equations of motion in terms of Delaunay variables are quite simple

and lead directly to the idea of the reduction of the system by averaging.

The orbital elements uniquely define a Keplerian orbit, given the Cartesian

position and velocity. For the case of the motion of a point mass in a central

gravitational field (which is integrable) they describe the specific conic section that

the path of the particle follows. In our case, that of a spacecraft in a non-central

field, the orbital elements vary as functions of time and represent a convenient way

to describe the motion. The orbits that we are primarily interested in are elliptical.

The six orbital elements most commonly used are:

• a: semi-major axis

• e: eccentricity

• i: inclination

• ω: argument of periapsis

• Ω: longitude of the ascending node

• M : mean anomaly

Some other orbital elements that we use are:

• ra = a(1 + e): radius of apoapsis

• rp = a(1− e): radius of periapsis

• E: eccentric anomaly (M = E − e sin E)

• f : true anomaly(tan f/2 =

√1+e1−e

tan E/2)

32

Page 46: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

All of the above orbital elements are in the inertial frame. Since the motion that we

consider takes place in the rotating frame, the longitude of the ascending node in

the rotating frame is also used:

Ω = Ω−N(t− t0) , (3.2)

where t0 is an initial epoch.

The equations of motion in terms of the orbital elements are obtained from

the Lagrange Planetary Equations, where R is the perturbing potential under

consideration[11]:

da

dt=

2

na

∂R

∂M, (3.3)

de

dt=

1− e2

na2e

∂R

∂M−√

1− e2

na2e

∂R

∂e, (3.4)

di

dt=

1

na2√

1− e2

[cot i

∂R

∂ω− csc i

∂R

∂Ω

], (3.5)

dt=

√1− e2

na2e

∂R

∂e− cot i

na2√

1− e2

∂R

∂i, (3.6)

dt=

1

na2√

1− e2 sin i

∂R

∂i, (3.7)

dM

dt= n− 1− e2

na2e

∂R

∂e− 2

na

∂R

∂a. (3.8)

Semi-equinoctial elements, denoted as h and k, are defined as [24]:

h = e sin ω ,

k = e cos ω .

This set of variables is particularly useful for near-circular orbits since all of the

motion is concentrated in a small region of possible (h, k) space, making for easier

analysis and plotting.

The Delaunay variables consist of the three angles M , ω and Ω complemented by

33

Page 47: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

three corresponding momenta, denoted as L, G and H and defined as follows:

L =√

µa ,

G = L√

1− e2 , (3.9)

H = G cos i .

The equations of motion of the 3-DOF system in terms of Delaunay variables are[8]:

dL

dt=

∂R

∂M,

dM

dt= n− ∂R

∂L,

dG

dt=

∂R

∂ω,

dt= −∂R

∂G, (3.10)

dH

dt=

∂R

∂Ω,

dt= − ∂R

∂H.

The fact that the full system has three degrees of freedom can be seen from the

Delaunay variable representation. The three degrees of freedom correspond to the

three free angular coordinates, M , ω and Ω. Note that H is the z-component of

the angular momentum. The Jacobi integral, defined in Eq.(2.46) in Cartesian

coordinates can be expressed in osculating orbital elements as:

J = − µ

2a−NH −R(a, e, i, ω, Ω, M) . (3.11)

Averaging is an approximation technique that can be used to remove degrees of

freedom from this system. In general, averaging occurs over an angle of the system

and causes its corresponding momentum to be conserved. For the first averaging, we

assume that the spacecraft’s mean motion about the planetary satellite, n =√

µ/a3,

is much greater than N , the mean motion of the planetary satellite about the

planet[37]. This means that over the time it takes for one revolution of the spacecraft

about the planetary satellite, the planetary satellite will only have moved by a small

amount around the planet. For a low-altitude orbit about Europa, the ratio of N/n

is ∼ 0.02− 0.03 which means that Europa will only have moved around Jupiter by

34

Page 48: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

about 7-10 degrees. This condition is also satisfied by many other planetary satellites

in our solar system. Two in particular to be discussed later are Saturn’s moons

Enceladus and Dione. They have N/n ratios of ∼ 0.09 and ∼ 0.04, respecively [37].

The first averaging, to reduce the system to 2 degrees of freedom is over the

mean anomaly of the spacecraft about Europa; define

R =1

∫ 2π

0

RdM . (3.12)

The computation of R requires the definition of the Cartesian coordinates x, y and

z as:

x = r[cos (ω + ν) cos Ω− sin (ω + ν) sin Ω cos i] , (3.13)

y = r[cos (ω + ν) sin Ω + sin (ω + ν) cos Ω cos i] , (3.14)

z = r sin i sin (ω + ν) , (3.15)

where t is the time. Note that for this computation, Ω is assumed to be constant.

The singly-averaged potential is found to be:

R =N2a2

4

(1− 3

2sin2 i +

3

2cos 2Ω sin2 i

)(1 +

3

2e2

)+

15

4e2 cos 2ω

[sin2 i + cos 2Ω

(1 + cos2 i

)]− 15

2e2 sin 2ω sin 2Ω cos i

+

µJ2

2a3(1− e2)3/2

(1− 3

2sin2 i

)+

3µC22

2a3(1− e2)3/2cos 2Ω sin2 i

+3µJ3

2a4(1− e2)5/2e sin ω sin i

(1− 5

4sin2 i

). (3.16)

Since we averaged over the mean anomaly, it is no longer present in the potential

defined by Eq.(3.16). Additionally, recall the equation for the time-derivative of

the semi-major axis in Eq.(3.3). Since M does not appear in R, a = 0 and so the

semi-major axis is constant and acts as another integral of motion. Since L =√

µa,

L is also constant. We have reduced the system to a 2-DOF system consisting of the

pairs (H, Ω) and (G, ω).

35

Page 49: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

The 2-DOF system is a time-periodic function of Ω. This makes initial analysis

of the qualitative behavior of the system difficult and so we reduce it further by

performing another averaging, over the orbit of the planetary satellite about the

planet. This averaging is justified for Europa since there are order of magnitude

differences between the period of a Europa orbiter (∼ 2 hours) and the period of

Europa about Jupiter (3.5 days) and between the period of Europa about Jupiter

and the time period of the instability of the spacecraft (∼ 1 month)[37]. It is

similarly justified for Enceladus and Dione.

Since Ω is uniformly decreasing with time due to its definition with respect to

the rotating coordinate frame, it can be used as an independent variable in place of

the time. Then, the dynamics are averaged over this new independent variable to

find their simplest form; define

R =1

∫ 2π

0

RdΩ . (3.17)

This defines the doubly averaged potential:

R =N2a2

4

[(1− 3

2sin2 i

)(1 +

3

2e2

)+

15

4e2 cos 2ω sin2 i

]+

µJ2

2a3(1− e2)3/2

(1− 3

2sin2 i

)(3.18)

+3µJ3

2a4(1− e2)5/2e sin ω sin i

(1− 5

4sin2 i

).

This potential has only tidal, J2, and J3 components since the C22 component

vanishes under the averaging. Although this means that the effect of C22 is not

included during the analysis of this system, its effect is considered in later chapters

when the averaging is removed. Since Ω is now eliminated, an additional integral of

motion exists for this system, the z-component of the angular momentum defined

by Eq.(3.9). Since a was constant in the 2 degree of freedom system, it is also

constant in this system. This leads to the definition of a simpler functional form of

36

Page 50: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

the integral, θ = H2/µa, or [7]:

θ = (1− e2) cos2 i . (3.19)

Therefore, the system governed by the doubly averaged potential in Eq.(3.18) is a

1-DOF system in (G,ω). To simplify later analysis, recall the Jacobi Integral which,

since it was an integral of motion in the full 3-DOF system, is also in integral of

motion here and has the same value as before the averaging since it is a constant.

Therefore, Eq.(3.11) implies that

J = − µ

2a−NH − R(e, i, ω) . (3.20)

Since a and H are integrals of motion in this system, the potential of the 1-DOF

system, R is also an integral of motion for the reduced 1-DOF system. Note that

this is not true for the potentials of the 2 and 3-DOF systems.

3.2 Frozen Orbits: Equilibrium Solutions of the

1-DOF System

The 1-DOF system is completely described by the variables (G,ω). The equations

of motion of these two variables can be obtained from the 1-DOF potential R as

follows:

dG

dt=

∂R

∂ω, (3.21)

dt= −∂R

∂G. (3.22)

Note that since R is an integral of motion in this system, it can be formally reduced

to quadratures. However, for this analysis it is more useful to consider the time

derivatives of the osculating orbital elements e, i and ω, noting that the equations for

e and i can be combined into the equation for G. These equations are obtained by

substituting the 1-DOF potential, Eq.(3.18) into the Lagrange Planetary Equations

37

Page 51: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

given in Eqs.(3.3)-(3.8). Before performing a full-blown analysis of the 6 equations,

some general observations can be made to simplify the analysis. First, note that

the 1-DOF potential is independent of M and Ω. This means that the other orbital

elements don’t depend on M and Ω in the 1-DOF system and so the equations for M

and Ω can be excluded from this analysis. Also, recall that the semi-major axis, a, is

constant in the 1-DOF system, and so its equation of motion can also be excluded.

The three remaining equations of motion, those for e, i and ω, are then the only

ones considered in this section:

de

dt=

15N2

8ne√

1− e2 sin2 i sin 2ω − 3J3n

2a3(1− e2)3sin i

(1− 5

4sin2 i

)cos ω , (3.23)

di

dt=− 15N2

16n

e2

√1− e2

sin 2i sin 2ω +3J3n

2a3(1− e2)3e cos i (1

−5

4sin2 i

)cos ω , (3.24)

dt=

3N2

8n√

1− e2

[5 cos2 i− 1 + 5 sin2 i cos 2ω + e2(1− 5 cos 2ω)

]+

3J2n

4a2(1− e2)2

(1− 5

4sin2 i

)+

3J3n

2a3(1− e2)3

sin ω sin i

e(3.25)

·[(

1− 5

4sin2 i

)(1 + 4e2

)− e2

sin2 i

(1− 19

4sin2 i +

15

4sin4 i

)].

Denote the equilibrium solutions of Eqs.(3.23)-(3.25) as frozen orbits. These

orbits are not frozen in space, but have constant values of e, i and ω on average.

Figure 3.1 shows an example of a frozen orbit integrated in the 1-DOF system over

a few orbital periods.

3.2.1 Tide-only Case

Consider first the tide-only case with J2 = J3 = 0 in Eqs.(3.23)-(3.25). First of all,

for circular orbits where e = 0, de/dt = di/dt = 0 automatically and ω is undefined.

Therefore, circular frozen orbits exist in the tide-only case for all inclination values.

Then with e 6= 0, de/dt = 0 and di/dt = 0 for ω = ±π/2, ω = 0 and ω = π. However,

for ω = 0,π the condition necessary to satisfy dω/dt = 0 is e = 1 which corresponds

38

Page 52: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−3000−2000

−10000

10002000

3000

−3000

−2000

−1000

0

1000

2000

3000−3000

−2500

−2000

−1500

−1000

−500

0

500

1000

1500

x (km)y (km)z

(km

)

Figure 3.1: A frozen orbit in the 1-DOF system integrated over several orbital periods.

to a rectilinear orbit. Since we are only interested in elliptic and circular orbits, we

consider only ω = ±π/2. The e and i solutions for the frozen orbits are then given

by:

5

3cos2 i− 1 + e2 = 0 (3.26)

which gives the following relationship between e and i:

e =

√1− 5

3cos2 i (3.27)

Figure 3.2 shows the frozen orbit solutions for the tide-only case. Note that the

elliptic frozen orbits bifurcate from the circular frozen orbits at i ∼ 39.23o and

i ∼ 140.77o.

An important observation of the tide-only frozen orbit solutions is that they

don’t have any dependence on the system parameters or the altitude of the desired

frozen orbit. Therefore, these frozen orbit solutions are valid for any planetary

satellite/planet system where only the tidal force is considered and for any altitude

frozen orbit.

39

Page 53: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

Figure 3.2: Frozen orbit solutions for the tide-only case (ω = ±π/2 for elliptic orbits).

3.2.2 Tide and J2 Case

Next, we remove the J2 = 0 condition and evaluate the frozen orbit solutions. Once

again, circular frozen orbits exist for all values of the inclination and elliptic frozen

orbits have the ω = ±π/2 condition. The equation to be solved to obtain the

relationship between eccentricity and inclination for the frozen orbit solutions is

Eq.(3.25) set equal to zero with J3 = 0 and ω = ±π/2:

3N2

3n√

1− e2[10 cos2 i− 6(1− e2)

3nJ2

4a2(1− e2)2

(1− 5

4sin2 i

)= 0 . (3.28)

This yields the following analytic relationship between e and i for the frozen orbit

solutions:

sin2 i =2N2(4 + 6e2)a2(1− e2)3/2 + 2n2J2

20N2a2(1− e2)3/2 + 5n2J2

. (3.29)

One difference between this result and the tide-only result is that the semi-major

axis appears in this solution. In the tide-only case, the frozen orbit solutions were

independent of the semi-major axis, meaning that given either an inclination or

eccentricity, all sizes of orbits would have the same frozen orbit characteristics. In

this case, however, the frozen orbit characteristics will vary as the size of the orbit

40

Page 54: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 3.1: Parameters of Europa

Parameter Symbol Value

Europa radius RE 1560.8 km

Europa orbital period T 3.55 days

Europa orbit rate N 2.05× 10−5 rad/s

Europa gravitational parameter µ 3.201× 103 km3/s2

Non-dimensional Europa J2a J2 4.2749× 10−4

Non-dimensional Europa C22a C22 1.2847× 10−4

Non-dimensional Europa J3a J3 1.3784× 10−4

a Values obtained from John Aiello, personal communication, August2004.

varies. Since we are primarily interested in low-altitude orbits, when solving for the

frozen orbit solutions we specify a radius of periapsis and scale the semi-major axis

correspondingly, where:

a =rp

1− e. (3.30)

A periapsis altitude of 100 km is chosen and the values of the other parameters used

are given in Table 3.1.

Figure 3.3 shows the frozen orbit solutions for the tide plus J2 case. The dashed

lines represent the frozen orbit solutions for the tide plus J2 case that differ from the

tide-only case. For eccentricities larger than about 0.4 and circular orbits, the frozen

orbit solutions are not affected by the inclusion of the J2 gravity field coefficient.

However, for elliptic orbits with eccentricities smaller than about 0.4, the frozen

orbit solutions for the tide plus J2 case have larger inclinations than the solutions

for the tide-only case. This occurs because the elliptic frozen orbits bifurcate from

the circular frozen orbits at a larger inclination for the tide plus J2 case than for

the tide-only case. The amount of the inclination difference depends on the value

41

Page 55: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

tide−onlytide+J

2

Figure 3.3: Comparison of frozen orbit solutions with and without J2 (ω = ±π/2 for elliptic orbits).

of J2 and the desired radius of periapsis. The results shown here are obviously for

particular values of J2 and rp. The overall importance of this result is that for

near-circular frozen orbits, including J2 leads to orbits with higher inclinations which

could be useful for the design of a science orbit about a planetary satellite since in

general, near-polar orbits are desirable for science orbits.

3.2.3 Tide, J2 and J3 Case

Consider now Eqs.(3.23)-(3.25) with J2 6= 0 and J3 6= 0. Again, Eqs.(3.23) and

(3.24) are in equilibrium for ω = ±π/2. Substituting this into Eq.(3.25) and setting

it equal to zero, we obtain:

0 =3

8

N2

n

1√1− e2

[4− 10 sin2 i + 6e2] +3nJ2

4a2(1− e2)2

(1− 5

4sin2 i

)± 3J3n

2a3(1− e2)3

sin i

e

[(1− 5

4sin2 i

)(1 + 4e2

)− e2

sin2 i

(1− 19

4sin2 i +

15

4sin4 i

)], (3.31)

where the ± on the J3 terms corresponds to ω = ±π/2. This implies that the

inclusion of the J3 gravity field coefficient breaks the symmetry between the

ω = ±π/2 frozen orbit solutions. In its place, a symmetry between the signs of J3

42

Page 56: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

Figure 3.4: Frozen orbit solutions for J2 6= 0, J3 > 0 and ω = ±π/2.

and ω is created, where changing the sign of J3 is equivalent to shifting ω by π. This

is useful in terms of mission design since at this time only the degree two terms of

the Europa gravity field are known [4]. An accurate value of J3 for Europa is not

currently available, including its sign. Due to the existence of the symmetry between

the signs of J3 and ω any desired frozen orbit can be attained, regardless of the sign

of J3 by adjusting ω appropriately.

Eq.(3.31) cannot be solved analytically for the relationship between e and i for

frozen orbit solutions in this case and must be solved numerically. One possible

solution method is to use the θ integral and eliminate either e or i from Eq.(3.31).

However, since the range of θ is not immediately apparent, it is easier to choose

values of either e or i and solve Eq.(3.31) for the other orbital element.

Figure 3.4 shows the frozen orbit solutions for J2 6= 0 and J3 > 0. It is

immediately apparent that including the effect of J3 significantly increases the

complexity of the structure of the frozen orbit solutions. First of all, it is not

surprising that the structure of the frozen orbit solutions changes depending on the

sign of J3 sin ω. However, the symmetry returns for large values of the eccentricity

where the solutions are the same as in the tide-only case. This occurs since as the

43

Page 57: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

eccentricity increases, the apoapsis of the orbit also increases and the influences of J2

and J3 on the dynamics decrease. Therefore, we conclude that the high-eccentricity

frozen orbits are not significantly changed by adding J2 and J3.

One of the new features apparent in Figure 3.4 is near-equatorial frozen orbits

that exist for all values of eccentricity for J3 sin ω > 0. We also observe that

circular frozen orbits exist only at four isolated points, which in this case have

inclinations of 0o, 180o, 63.43o and 116.56o. In contrast, in the tide-only and tide

plus J2 cases, circular frozen orbits existed over all inclinations. The exact values

of the non-equatorial inclinations for the four circular frozen orbits depend on the

parameters of the system. Those values will always be close to but not exactly equal

to i = sin−1(√

4/5). Their deviation from that value is due to the existence of the

tidal term. In the case where only J2 and J3 are included, these circular frozen

orbits exist independently of the parameter values [37]. Including the tidal term as

we do here breaks this relationship and the frozen orbits with i = sin−1(√

4/5) have

non-zero eccentricities.

The frozen orbits with eccentricities smaller than about 0.2 have very different

characteristics depending on whether J3 sin ω > 0 or J3 sin ω < 0. In terms of

Figure 3.4 where J3 > 0, frozen orbits with ω = π/2 have inclinations in the middle

range, between about 45o − 65o and 105o − 135o. Conversely, for frozen orbits with

ω = −π/2, near-circular frozen orbits exist for inclinations between about 65o− 105o

(i.e. near-polar). In addition, for eccentricities up to about 0.1 there exist frozen

orbits with inclinations ranging from circular to about 40o for direct orbits and in

the 140o − 180o range for retrograde orbits.

3.2.4 Variations in J3

The results discussed above were generated for specific values of J2 and J3. Since an

accurate value for J3 for Europa is not currently known, it is important to verify that

44

Page 58: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

(a) Order of magnitude larger J3

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

(b) Order of magnitude smaller J3

Figure 3.5: Frozen orbit solutions with larger and smaller values of J3 (compared to Table 3.1).

the qualitative features of the results stay the same for different values of J3. This

would demonstrate that these results are applicable not only to Europa, whatever

its gravity coefficients are, but also to other planetary satellites. Figure 3.5 shows

the frozen orbits solutions using values of J3 an order of magnitude larger and an

order of magnitude smaller than the value given in Table 3.1. Although the shape of

the curves is slightly different from the curves in Figure 3.4, in both cases the same

classes of frozen orbits still exist. Therefore, even though the exact value of J3 for

Europa is not known at this time, the analysis of the system shown thus far should

be valid no matter what the value is and these results can be used in subsequent

analysis.

3.3 Stability of Frozen Orbit Solutions

To determine the stability of the frozen orbit solutions, linearize Eqs.(3.23)-(3.25)

about the frozen orbit solutions. First, express Eqs.(3.23)-(3.25) as:

de

dt= fe(e, i, ω) , (3.32)

di

dt= fi(e, i, ω) , (3.33)

45

Page 59: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

dt= fω(e, i, ω) . (3.34)

Define i = i∗ + δi, e = e∗ + δe and ω = ω∗ + δω where e∗, i∗ and ω∗ are the frozen

orbit values of these orbital elements. Then,

d(e∗ + δe)

dt= fe(e

∗ + δe, i∗ + δi, ω∗ + δω) , (3.35)

d(i∗ + δi)

dt= fi(e

∗ + δe, i∗ + δi, ω∗ + δω) , (3.36)

d(ω∗ + δω)

dt= fω(e∗ + δe, i∗ + δi, ω∗ + δω) . (3.37)

Expanding up to first order, we obtain:

e∗ + δe = fe(e∗, i∗, ω∗) +

∂fe

∂e

∣∣∣∣∗δe +

∂fe

∂i

∣∣∣∣∗δi +

∂fe

∂ω

∣∣∣∣∗δω , (3.38)

i∗ + δi = fi(e∗, i∗, ω∗) +

∂fi

∂e

∣∣∣∣∗δe +

∂fi

∂i

∣∣∣∣∗δi +

∂fi

∂ω

∣∣∣∣∗δω , (3.39)

ω∗ + δω = fω(e∗, i∗, ω∗) +∂fω

∂e

∣∣∣∣∗δe +

∂fω

∂i

∣∣∣∣∗δi +

∂fω

∂ω

∣∣∣∣∗δω . (3.40)

First, since the frozen orbit solutions are equilibrium solutions to Eqs.(3.32)-(3.34)

e∗ = fe(e∗, i∗, ω∗) = 0, i∗ = fi(e

∗, i∗, ω∗) = 0 and ω∗ = fω(e∗, i∗, ω∗) = 0. In addition,

since ω∗ = ±π/2:

∂fe

∂e

∣∣∣∣∗

=∂fe

∂i

∣∣∣∣∗

=∂fi

∂e

∣∣∣∣∗

=∂fi

∂i

∣∣∣∣∗

=∂fω

∂ω

∣∣∣∣∗

= 0 .

Therefore, Eqs.(3.38)-(3.40) simplify to the following:

δe =∂fe

∂ω

∣∣∣∣∗δω , (3.41)

δi =∂fi

∂ω

∣∣∣∣∗δω , (3.42)

δω =∂fω

∂e

∣∣∣∣∗δe +

∂fω

∂i

∣∣∣∣∗δi . (3.43)

46

Page 60: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

unstablestable

Figure 3.6: Stability of frozen orbits about Europa for the tide-only case.

Next, taking the second derivative of Eq.(3.43) yields:

δω =

(∂fω

∂e

∣∣∣∣∗

∂fe

∂ω

∣∣∣∣∗+

∂fω

∂i

∣∣∣∣∗

∂fi

∂ω

∣∣∣∣∗

)δω . (3.44)

For each frozen orbit the partial derivatives evaluated at that frozen orbit are

constants. Therefore, δω is just a constant multiplied by δω. This second order

linear differential equation has the simple form:

δω = Λ2(e∗, i∗, ω∗)δω , (3.45)

where Λ2 is a constant for each frozen orbit.

The evaluation of Λ2(e∗, i∗, ω∗) for a particular frozen orbit determines its

stability. If Λ2 > 0 the frozen orbit is unstable and if Λ2 < 0 the frozen orbit is

stable. Figure 3.6 shows the stability of the frozen orbits for the tide-only case. Note

that all of the elliptic frozen orbits are stable and the circular frozen orbits are stable

for inclinations less than about 39.23o and greater than about 140.77o. The circular

frozen orbits with inclinations between ∼ 39.27o − 140.77o are unstable.

Figure 3.7 shows the stability of the frozen orbits for the tide plus J2 case. The

similarities between this and the tide-only case are that the elliptic frozen orbits

47

Page 61: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 10

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

unstablestable

Figure 3.7: Stability of the frozen orbits for the tide plus J2 case.

with eccentricities larger than about 0.2 are stable. As well, the near-equatorial

circular frozen orbits are also stable. The major difference between this case and

the tide-only case is that the elliptic frozen orbits with eccentricities smaller than

about 0.2 are unstable as are the circular frozen orbits with inclinations ranging

from about 47o − 133o. The circular frozen orbits change from stable to unstable at

the bifurcation points (i.e., the locations where the elliptic frozen orbits bifurcate

from the circular frozen orbits).

The stability of the frozen orbits for the case including the tidal term, J2 and

J3 is similar to the tide plus J2 case in that the higher eccentricity frozen orbits are

stable. However, for the lower-eccentricity frozen orbits, the stability characteristics

of this case depend on the sign of J3 sin ω. For J3 sin ω > 0, the frozen orbits with

eccentricities less than about 0.2 are unstable. These are the mid-inclination frozen

orbits. In addition, the near-equatorial frozen orbits that exist for all inclinations are

also stable. On the other hand, for J3 sin ω < 0, the low-eccentricity, near-equatorial

frozen orbits are stable while the near-circular, near-polar frozen orbits are unstable.

The strength of the instability of an unstable frozen orbit can be measured by the

characteristic time. The characteristic time is defined as τ = 1/Λ, and is the time

48

Page 62: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

unstablestable

Figure 3.8: Stability for frozen orbits about Europa that include the tidal term, J2 and J3.

interval it takes the eccentricity to increase by a factor of exp (1) ∼ 2.718[37]. In

general, the eccentricity will increase by an order of magnitude after exp (1) ∼ 2.718

characteristic times. Even though the tide-only frozen orbit solutions themselves do

not depend on the system parameters, the characteristic times of the unstable frozen

orbits are parameter-dependent. Consider Figure 3.9 which shows the characteristic

times for the 100 km altitude unstable circular frozen orbits for the tide-only cases of

Europa, Dione and Enceladus. Observe that the characteristic times for Europa are

the largest, followed by Dione and then Enceladus. This means that the frozen orbits

about Enceladus are the most unstable. In particular, the Enceladus characteristic

times are about an order of magnitude smaller than the Europa case. This has

strong implications for orbit design since a smaller characteristic time implies a more

highly unstable system in the sense that it will grow along its unstable manifold

more rapidly. All three of these planetary satellites are highly perturbed systems

in which science orbit design is very difficult, and the goals for the orbit design

will have to take the characteristic times into consideration. Observe in Figure 3.9

that the more polar the frozen orbit, the more unstable it is since it has a smaller

characteristic time.

49

Page 63: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

40 50 60 70 80 90 100 110 120 130 1400

10

20

30

40

50

60

70

80

90

inclination (degrees)

time

(day

s)

(a) Europa

40 50 60 70 80 90 100 110 120 130 1400

5

10

15

20

25

30

inclination (degrees)

time

(day

s)

(b) Dione

40 50 60 70 80 90 100 110 120 130 1400

1

2

3

4

5

6

7

inclination (degrees)

char

acte

ristic

tim

e (d

ays)

(c) Enceladus

Figure 3.9: Characteristic times for the tide-only unstable frozen orbits with 100 km altitudes.

50

Page 64: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

60 70 80 90 100 110 12013

13.5

14

14.5

15

15.5

inclination (degrees)

char

acte

ristic

tim

e (d

ays)

(a) Circular, near-polar

40 50 60 70 80 90 100 110 120 130 14010

20

30

40

50

60

70

80

90

inclination (degrees)

char

acte

ristic

tim

e (d

ays)

(b) Elliptic mid-inclination

Figure 3.10: Characteristic times for unstable frozen orbits about Europa

For the case including the tidal term, J2 and J3, there are two classes of unstable

frozen orbits, namely the near circular near polar frozen orbits with J3 sin ω < 0

and the mid-inclination frozen orbits with J3 sin ω > 0. Figure 3.10 shows the

characteristic times as functions of the inclination for both of these classes of

unstable frozen orbits about Europa. The characteristic times for the near-circular,

near-polar frozen orbits have the same structure as the circular frozen orbits for

the tide-only case, with the minimum characteristic time corresponding to a polar

orbit. The same trend is also apparent for the mid-inclination frozen orbits where

the closest to polar inclination also corresponds to the minimum characteristic time.

Just as we verified the frozen orbit solutions for different possible values of J3,

we also look at the stability and characteristic times for values of J3 one order of

magnitude smaller and one order of magnitude larger than the value in Table 3.1.

Figure 3.11 shows the stability and characteristic times for frozen orbits computed

with a J3 value one order of magnitude smaller and Figure 3.12 shows the same

plots computed with a J3 value one order of magnitude larger. We see that when

compared the the results for our nominal J3 case, the stability properties follow the

same trends. For the larger J3 value, the characteristic times are smaller, meaning

51

Page 65: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

unstablestable

(a) Stability

65 70 75 80 85 90 95 100 105 110 11513.5

14

14.5

15

inclination (degrees)

char

acte

ristic

tim

e (d

ays)

(b) Characteristic times for circular, near-polar orbits

40 50 60 70 80 90 100 110 120 130 14010

20

30

40

50

60

70

80

90

100

inclination (degrees)

char

acte

ristic

tim

e (d

ays)

(c) Characteristic times for elliptic, mid incli-nation orbits

Figure 3.11: Stability and characteristic times of frozen orbits about Europa with J3 one order ofmagnitude smaller than Table 3.1

that the orbits are more unstable. This is expected since the J3 perturbation is

acting more strongly. For the smaller J3 value, the characteristic times are larger

than for the nominal J3 value case meaning that the orbits are more stable.

52

Page 66: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

unstablestable

(a) Stability

65 70 75 80 85 90 95 100 105 110 1157

8

9

10

11

12

13

inclination (degrees)

char

acte

ristic

tim

e (d

ays)

(b) Characteristic times for circular, near-polar orbits

40 50 60 70 80 90 100 110 120 130 14010

12

14

16

18

20

22

24

26

28

inclination (degrees)

char

acte

ristic

tim

e (d

ays)

(c) Characteristic times for elliptic, mid-inclination orbits

Figure 3.12: Stability and characteristic times of frozen orbits about Europa with J3 one order ofmagnitude larger than Table 3.1

.

53

Page 67: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

3.4 Contour Plots and Frozen Orbit Integrations

In this section, the secular motion in the vicinity of frozen orbits will be examined

using contour plots. Then, results of integrations in the 3-DOF system of frozen

orbit solutions will be compared with the contour plots to determine if the stable

frozen orbits are viable orbits in the full 3-DOF system and how unstable frozen

orbits behave in the 3-DOF system. All of the results in this section are computed

for the Europa system, with the tidal component, J2, C22 and J3 gravity field terms,

but are applicable to other planetary satellite systems.

3.4.1 Contour Plots

Contour plots are a tool that can be used to visualize the secular motion of an

orbiter. They are plotted in the 1-DOF system which is integrable and a function

of only two variables, (G,ω) and so a plot in terms of those two variables can

completely describe the motion. It is useful, however, to produce contour plots for

other sets of variables in order to get a better picture of the motion. Recall that

G =√

µa(1− e2). Therefore, a plot in terms of (e,ω) can also be used to describe

the motion. The final contour plot that we use is in terms of the semi-equinoctial

variables (h,k). Although we initially compute these plots for a range of frozen

orbits, they are most useful to describe near-circular motion, as will be shown later.

Note that contour plots of the set (i,ω) are also possible. However, since the motion

is integrable, the inclination can be determined from the integral of motion θ and

the eccentricity. We are, in general, more interested in the motion of the eccentricity

than the inclination since both remain relatively constant for stable orbits, and

for unstable orbits it is the growing eccentricity that causes an impact with the

planetary satellite.

The contour plots are made up of curves with constant values of the 1-DOF

potential, R. The curves define the path that the motion will follow on average.

54

Page 68: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 250 300 3500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

ω (degrees)

ecce

ntric

ity

−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

k=ecosω

h=es

inω

−3000 −2000 −1000 0 1000 2000 3000

−3000

−2000

−1000

0

1000

2000

3000

Gcosω

Gsi

Figure 3.13: Contour plots for motion in the vicinity of a stable frozen orbit with e∗ = 0.548,i∗ = 50o, J3 sinω > 0.

55

Page 69: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 250 300 3500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

ω (degrees)

ecce

ntric

ity

−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

k=ecosω

h=es

inω

−2000 −1500 −1000 −500 0 500 1000 1500 2000

−2000

−1500

−1000

−500

0

500

1000

1500

2000

Gcosω

Gsi

Figure 3.14: Contour plots for motion in the vicinity of an unstable frozen orbit with e∗ = 0.0.039,i∗ = 55o, J3 sinω > 0.

56

Page 70: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Figures 3.13-3.16 show contour plots in all three regimes for motion in the vicinity

of four different types of frozen orbits. Figure 3.13 is an example of motion in the

vicinity of a stable frozen orbit with an inclination of 50o, an eccentricity of about

0.548 and J3 sin ω > 0. It is, therefore, a mid-inclination, mid-eccentricity frozen

orbit. We see that for this case the motion is symmetric for ω = ±π/2, showing

that the tidal effect dominates over the J2 and J3 effects. Figure 3.14 shows motion

in the vicinity of an unstable frozen orbit with an inclination of 55o, an eccentricity

of about 0.039 and J3 sin ω > 0. In this case, the motion is not symmetric about

ω = ±π/2 since the eccentricity is small and the J2 and J3 effects are important.

Note that there appears to be two stable regions in the plot, other than the unstable

frozen orbits. These regions correspond to equilibrium points in the 1-DOF system

that do not have the altitude of periapsis of 100 km that we consider. Their locations

correspond to motion below the surface of the planetary satellite since the contour

plots have a constant value of the semi-major axis. For this case, the semi-major

axis is a = rp/(1 − e) = 1660.8/(1 − 0.039) = 1728 km and so motion below the

surface of the planetary satellite occurs for eccentricities greater than 0.097. The

consequence of this result is that the only motion that we are interested in from a

contour plot is that in the vicinity of the frozen orbit.

Now, consider contour plots for frozen orbits with J3 sin ω < 0. First, Figure 3.15

shows motion in the vicinity of a stable frozen orbit with an inclination of 30o and

an eccentricity of about 0.047. Again, the motion is not symmetric about ω = ±π/2

since this is a low-altitude orbit where the effects of J2 and J3 are significant. The

last example, in Figure 3.16, is an unstable frozen orbit with an inclination of 80o

and an eccentricity of 0.021. Note that the plot for (e,ω) is not over the entire range

of eccentricity but just in the region of the frozen orbit. This gives a better picture

of the motion in the vicinity of the frozen orbit and as previously noted, motion

for larger values of the eccentricity takes place under the surface of the planetary

57

Page 71: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 250 300 3500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

ω (degrees)

ecce

ntric

ity

−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

k=ecosω

h=es

inω

−2000 −1500 −1000 −500 0 500 1000 1500 2000

−2000

−1500

−1000

−500

0

500

1000

1500

2000

Gcosω

Gsi

Figure 3.15: Contour plots for motion in the vicinity of a stable frozen orbit with e∗ = 0.047,i∗ = 30o, J3 sinω < 0.

58

Page 72: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 250 300 3500

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

ω (degrees)

ecce

ntric

ity

−2000 −1500 −1000 −500 0 500 1000 1500 2000

−2000

−1500

−1000

−500

0

500

1000

1500

2000

Gcosω

Gsi

−2000 −1500 −1000 −500 0 500 1000 1500 2000

−2000

−1500

−1000

−500

0

500

1000

1500

2000

Gcosω

Gsi

Figure 3.16: Contour plots for motion in the vicinity of an unstable frozen orbit with e∗ = 0.021,i∗ = 80o, J3 sinω < 0.

59

Page 73: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

satellite. The frozen orbit in this example is near-circular and near-polar, and it is

the type that will be considered later as a possible science orbit for a mission to a

planetary satellite.

3.4.2 Frozen Orbit Integrations

Results of integrations of frozen orbit solutions are presented overlaying contour

plots in (e,ω) space for both stable and unstable frozen orbits. For stable frozen

orbits, the goal is to determine whether they are viable long-term stable orbits in

the 3-DOF system. For unstable frozen orbits, we are interested in comparing the

motion in the 3-DOF system with the contour plots to see if the motion follows the

contours well.

Figure 3.17(a) shows a contour plot in (e,ω) space along with the results of the

integration in the 3-DOF system of a stable frozen orbit with an inclination of 134o,

an eccentricity of 0.387 and J3 sin ω > 0. The integration takes place over a period

of 150 days and it is clear that the orbit is stable since it stays in the libration

region. Note that the center of the circle representing the integrated trajectory is

not exactly at the frozen orbit location. This is due to the perturbations that are

averaged out in the 1-DOF system. Even though the integrated trajectory does not

follow the curves in the contour plot exactly, the frozen orbit conditions still yield a

stable orbit in the 3-DOF system.

The next example, shown in Figure 3.17(b), shows the integration of a stable

frozen orbit with an inclination of 62o, an eccentricity of 0.795 and J3 sin ω > 0.

The integrated trajectory, shown by the red curve, does not follow the curves in the

contour plot at all and reaches a very high eccentricity which leads to an impact with

the planetary satellite. This occurs since although the orbit is stable in the 1-DOF

system, it has a large eccentricity which means that it has a large apoapsis radius.

Since the motion takes place farther away from the planetary satellite, perturbations

60

Page 74: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

not taken into account in the derivation of the 1-DOF arise and cause the 3-DOF

trajectory to be unstable.

Finally, Figure 3.17(c) is the integration of an unstable frozen orbit with an

inclination of 80o, an eccentricity of 0.021 and J3 sin ω < 0. Since this orbit is

unstable, all that we expect is the trajectory to follow the contour curve starting at

the frozen orbit location. This is observed, allowing us to conclude that for small

eccentricities the contour plots give a very good estimate of what motion to expect in

the 3-DOF system. Therefore, the motion in the 1-DOF system can be used to plan

orbits for a mission to a planetary satellite, as will be seen in subsequent chapters.

61

Page 75: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 250 300 3500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

ω (degrees)

ecce

ntric

ity

(a) e∗ = 0.387, i∗ = 134o, J3 sinω > 0

0 50 100 150 200 250 300 3500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

ω (degrees)

ecce

ntric

ity

(b) e∗ = 0.795, i∗ = 62o, J3 sinω > 0

0 50 100 150 200 250 300 3500

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.1

ω (degrees)

ecce

ntric

ity

(c) e∗ = 0.021, i∗ = 80o, J3 sinω < 0

Figure 3.17: Contour plots and integrations in the 3-DOF of various frozen orbits

62

Page 76: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

CHAPTER 4

ROBUST CAPTURE AND TRANSFER

TRAJECTORIES

In this chapter, we study capture trajectories in the Hill 3-body problem.

Capture trajectories begin in the exterior region of Hill’s problem, enter the interior

region and orbit the planetary satellite at least once. One particular feature of

capture trajectories that we investigate is their lifetime. Uncontrolled, these orbits

can impact with the planetary satellite or exit the interior Hill region after very

short time spans. Koon et al. [20] investigated orbits that travel between the interior

and exterior Hill regions and showed that the amount of time a trajectory spends

orbiting the planetary satellite is determined by chaotic dynamics.

A method is developed which identifies sets of capture trajectories that do

not impact or escape the planetary satellite for extended time periods. These are

called ‘safe trajectories’ and the regions in which they lie ‘safe zones’. These safe

low-energy trajectories may be useful for a future mission to a planetary satellite

such as Europa. With the science goals of that type of mission being such that a

low-altitude, high inclination stable orbit about Europa is desirable, we examine

low-cost methods to transfer from a safe capture trajectory to a long-term stable

orbit such as the elliptic frozen orbits developed in the last chapter.

Criteria on the safe capture trajectories that result in the lowest cost transfers

are determined and we develop schemes to transfer to elliptic frozen orbits and

circular frozen orbits. These schemes are analyzed based on their costs and specific

63

Page 77: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

examples are shown. The end result is a method that identifies trajectories that

enter into orbit about Europa on low-energy capture trajectories and which do not

impact or escape for at least one week, and from which it is possible to transfer to a

more stable orbit with relatively low cost.

4.1 Non-dimensional Hill 3-Body Problem and

Libration Points

In this chapter, only the tidal term is included in the perturbing potential R, given

in Eq.(3.1), yielding the following potential:

R =1

2N2(3x2 − r2

). (4.1)

This is an acceptable simplification since we are primarily dealing with motion

farther away from Europa than the low-altitude frozen orbits in which the gravity

coefficients are important. Although the capture trajectories to be presented in this

chapter are in the vicinity of Europa, if they do pass very closely to Europa, it is

only for a short period of time. In addition, when we consider transfers from capture

trajectories to frozen orbits, we consider either the elliptic frozen orbits or higher

altitude circular orbits, both of which are not affected very strongly by J2, C22 or J3.

The equations of motion for this system are:

x = 2Ny − µ

r3x + 3N2x , (4.2)

y = −2Nx− µ

r3y , (4.3)

z = − µ

r3z −N2z , (4.4)

and they are in dimensional form. However, to simplify the computations in this

chapter we make the model non-dimensional. To do this, take l = (µ/N2)1/3

as the

unit length and τ = 1/N as the unit time. The equations then take on the following

64

Page 78: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8−0.5

−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

0.4

0.5

x

y

Forbidden Region

Forbidden Region

L2 L

1

Figure 4.1: Regions of allowable motion and libration points for J = −2.15. The dotted linerepresents a planar capture trajectory that originates near L2.

form

x = 2y − x

r3+ 3x, , (4.5)

y = −2x− y

r3, (4.6)

z = − z

r3− z , (4.7)

and have no free parameters.

Equilibrium solutions exist in the Hill 3-body problem, analogous to the L1 and

L2 libration points in the CR3BP. There are two libration points that are symmetric

about the origin with coordinates x = ± (µ/3N2)1/3

, y = z = x = y = z = 0

where the x-coordinate in nondimensional form is x = ± (1/3)1/3 = ±0.693 · · · (see

Figure 4.1).

Recall that Eqs. (4.5)-(4.7) have the Jacobi integral, which in non-dimensional

form for the tide-only case is:

J =1

2v2 − 1

r− 1

2(3x2 − z2) , (4.8)

where v =√

x2 + y2 + z2 is the speed of the spacecraft in the rotating frame. The

condition v ≥ 0 in Eq.(4.8) places a restriction on the position of the particle for

a given value of J . Setting v = 0 defines the zero-velocity surface, which sets a

65

Page 79: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

physical boundary for the motion (see Figure 4.1).

For values of the Jacobi constant above a critical value, it is possible for both

escaping and capture trajectories to exist in this problem. This critical value of J

defines the energy at which the zero-velocity surfaces open at L1 and L2 and is equal

to

Jcrit = −1

292/3 = −2.16337 . (4.9)

Since our goal is to characterize capture trajectories, we restrict ourselves to

trajectories with a Jacobi constant greater than Jcrit.

4.2 Periapsis Poincare Maps

A Poincare map associates a continuous time dynamical system to a discrete time

system. The use of a Poincare map reduces the dimensionality of a system by at

least one, and by two if there exists a first integral in the system, as is the case in the

Hill 3-body problem. In general, Poincare maps are defined such that the surface of

section is a plane in position space, such as x = 0. However, since all that is needed

to define a Poincare map are two surfaces of section that are transversal to the flow,

they are defined differently here and periapsis passages are used as the surface of

section, following [42].

4.2.1 Definition of the Map

Our Poincare map relates capture trajectories that start near the libration point

L2 to the periapsis passages of these trajectories. Following [41], the initial surface

of section is defined as the surface of a sphere with radius (13)1/3 bounded by the

zero-velocity surface. This surface passes through the libration point region and

extends as far as the zero-velocity surface of the Hill problem. The image surface is

defined by the periapsis condition r = 0 and r > 0. For the 3-D Hill problem, the

phase space is six-dimensional. The use of the Poincare map restricts the problem

66

Page 80: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

to the surface of section, and hence reduces the problem by one dimension. The

Poincare map is then computed at a given value of the Jacobi integral J , (and so

the map is a function of J), reducing the problem by one further dimension to four

dimensions.

Due to the symmetry between the L1 and L2 libration points, analysis is only

performed on capture trajectories that originate near L2; however, the results

can be directly related to trajectories that originate near L1 by the following

transformation[42]

(x, y, z, x, y, z, t)Gµ−→ (−x,−y,−z,−x,−y,−z, t) (4.10)

since the equations of motion remain unchanged. By a capture trajectory, we mean a

trajectory that remains in the Hill region of the planetary satellite for a finite period

of time. These trajectories enter the region of the planetary satellite in the vicinity

of the libration point and orbit the planetary satellite at least once. Figure 4.1 shows

an example of a capture trajectory.

In the 3-D case, the periapsis Poincare map is four dimensional and four

parameters are needed to characterize it. For the initial surface of section, we use

the (x, z) coordinates and two angles (δ, φ) for the direction of the velocity vector.

The (x, z) coordinates are chosen randomly on the section of the surface of the

sphere that falls within the allowable region. Since the surface of section is a sphere,

the y coordinate can be computed from the relationship r =√

x2 + y2 + z2, where

r = (1/3)1/3 and the sign for y is chosen randomly. The initial conditions are then:

x0 = x x0 = v cos φ cos δ

y0 = ±√

r2 − x2 − z2 y0 = v cos φ sin δ (4.11)

z0 = z z0 = v sin φ

67

Page 81: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

where (φ, δ) ∈ [π/2, 3π/2], and v is computed from the Jacobi integral:

v =

√2

(J +

1

r

)+ (3x2 − z2) . (4.12)

The parameters used to represent the image map are the periapsis position vector

(x, y, z) and the inclination.

With the initial conditions given by Eq.(4.11), an 8(7) order Runge-Kutta-

Fehlberg integration routine is used to integrate the trajectories. They are integrated

for multiple passages through the image map, meaning that they have multiple

periapsis passages. The phase space is studied in terms of these multiple periapsis

passages for various values of J .

4.2.2 Poincare Map Results

As an initial study, consider the planar case of the Hill 3-body problem. In this

model, motion is restricted to lie in the x-y plane, and the periapsis Poincare

map reduces to two dimensions. The initial surface, which consists of the surface

of a sphere in the three-dimensional case reduces here to an arc of circle. It is

parameterized with the x coordinate and an angle δ which defines the direction of

the velocity vector. This is equivalent to the initial conditions defined by Eq.(4.11)

with z = φ = 0. On the image map, the coordinates (x, y) of the periapsis position

vector are used.

Although the periapsis Poincare map is computed in non-dimensional coordinates

to allow for its application to many different physical systems, we primarily consider

Europa. On many of the periapsis Poincare maps, the surface of Europa is indicated

to give perspective to the plot. Although some of the periapsis passages of these

trajectories lie beneath the surface of Europa, indicating that they impact with it, for

the time being they are included in the analysis since they are not necessarily impact

trajectories when a different planetary satellite is considered. In a later section of

68

Page 82: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

(a) First periapsis passage region

(b) First 4 periapsis passage regions: Black - 1st, blue- 2nd, red - 3rd, green - 4th

Figure 4.2: Periapsis Poincare maps for J=-2.15 (planar case) where the circle represents thesurface of Europa.

this chapter, we distinguish between impacting and non-impacting trajectories.

We first consider the periapsis Poincare map for the first periapsis passage of

trajectories with J = −2.15, as was done in [42]. As shown in Figure 4.2(a) the first

periapsis passages are divided into two disjoint regions. The first, located near the

planetary satellite, corresponds to trajectories that immediately enter the region in

the vicinity of the planetary satellite and have their first periapsis passage there.

The second region, located near the L2 libration point, corresponds to trajectories

that do not immediately leave the vicinity of L2 and have their first periapsis passage

in that region. These can be associated with the periodic orbit that exists about L2,

as shown in [41]. Some of these trajectories will subsequently enter the region of the

planetary satellite and some will escape from the system entirely. Note that by our

current definition of capture trajectory, there is no guarantee that the trajectory

actually comes from outside the Hill region and not from inside the region itself.

To remedy this, an additional integration is performed whereby the trajectory is

integrated backwards in time for four time units. This is a sufficient amount of

69

Page 83: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

time for a true capture trajectory to return to its starting location outside of the

Hill region. Therefore, the set of capture trajectories is restricted to those that lie

beyond the boundary of the Hill region after a backwards-time integration of four

time units. If after the backwards-time integration a trajectory was located inside

the Hill region, we could conclude that it is not a true capture trajectory, but just

a trajectory that happened to have an inward velocity at the opening of the Hill

region at the initial time.

We now extend the periapsis Poincare map to four periapsis passages by plotting

the first four periapsis passages of each trajectory for J = −2.15. Each periapsis

passage is assigned a different color, as shown in Figure 4.2(b). Observe that

each subsequent periapsis passage region becomes more spread out. This occurs

since symplectic coordinates are not used on our map. In addition, some of the

colors overlap (i.e, there is a small blue region inside the black region). These

overlaps occur because, as previously mentioned, some of the trajectories do not

immediately have a periapsis passage in the vicinity of the planetary satellite but

instead have their first periapsis passage near the L2 region. Therefore it is not until

the second periapsis passage that these trajectories are near the planetary satellite.

The dynamics of these trajectories, once they enter the region near the planetary

satellite, are the same as the dynamics of the trajectories that enter that region

immediately. It is also important to note that the gaps in some of the regions are not

because trajectories cannot have periapsis passages there, but because these regions

were generated numerically and, as such, only a finite number of trajectories can be

considered. From this point forward, the region of first periapsis passages that occur

in the vicinity of the L2 libration point will be ignored. The first periapsis passage

of a particular trajectory is defined to be its first periapsis passage that occurs in

the vicinity of the planetary satellite.

To illustrate how the periapsis Poincare maps depend on J , Figure 4.3 shows

70

Page 84: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Figure 4.3: Poincare maps for various values of J . The circle in each plot denotes the surface ofEuropa.

71

Page 85: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.12−0.11

−0.1−0.09

−0.08−0.07

−0.06−0.05

0

0.01

0.02

0.03

0.04

0.05−0.015

−0.01

−0.005

0

0.005

0.01

0.015

xy

Figure 4.4: First periapsis passage region for J=-2.15 (3-d case). The position is indicated by thebase of the arrow, the velocity direction by the direction of the arrow, and the magnitude of thevelocity by the length of the arrow.

a series of Poincare maps for increasing values of J . The regions become larger

as J increases, which is expected since the measure of the initial condition region

increases with J . Only the first periapsis passage regions are shown, since the trends

shown by these regions also apply to subsequent periapsis passages. Note that the

plots in Figure 4.3 are mirror images of the corresponding plots shown in [42].

It is much more difficult to visualize the periapsis Poincare maps for the 3-D

problem since the Poincare map itself is four dimensional. In general, the trends we

see in the planar case apply to the non-planar case, such as the periapsis passage

regions becoming larger in position space as J is increased as well as for subsequent

periapsis passages. One way to visualize the Poincare map for the non-planar case

is to plot each periapsis passage’s three-dimensional position vector, with an arrow

indicating the direction of its velocity (representing the inclination at the periapsis

passage). An example of this is shown in Figure 4.4. Another way to visualize

the non-planar case is by considering the inclination of the periapsis passage as

a function of its radius. Figure 4.5(a) shows the inclination as a function of the

radius for the first periapsis passage of trajectories with various values of J . Observe

that as J increases, the periapsis passages reach higher inclinations. Figure 4.5(b)

72

Page 86: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

(a) First periapsis passage regions, various values ofJ

(b) First 4 periapsis passages, J = −2.15

Figure 4.5: Inclination as a function of normalized radius for 3-d capture trajectories.

73

Page 87: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

shows the inclination as a function of the radius for multiple periapsis passages of

trajectories with J=-2.15. In this case, the maximum obtainable inclination does

not increase for subsequent periapsis passages, but the periapsis passages do occur

over larger ranges of radii.

We now introduce a more systematic description of these maps and regions.

Define the set of trajectories that lie in the first periapsis passage region close to

Europa as the set S1,

S1 = x|x ∈ Initial Region . (4.13)

Then, under the flow dynamics of the Hill problem the Poincare map can be

represented as Φ, where

Sn+1 = Φ(Sn) = Φ2(Sn−1) = · · · = Φn(S1) . (4.14)

Thus, it is clear that any point in Sn+1 has a unique image in S1. If we use the

symmetry operator defined by Eq.(4.10), Gµ, we can also associate the ‘mirror image’

trajectories as S ′n = Gµ(Sn) defined by reflecting all coordinates and velocities about

the origin.

4.3 Symmetry Between Escape and Capture

Trajectories

Aside from the symmetry Gµ between L1 and L2 which was previously discussed,

additional symmetries exist in this problem (and in the CR3BP). Two that are

considered relate to the symmetry between escape and capture trajectories. If

(x, y, z, x, y, z, t) is a solution of the equations of motion, then the trajectories

obtained by applying the following transformations are also solutions [42]:

(x, y, z, x, y, z, t)G1−→ (−x, y, z, x,−y,−z,−t) , (4.15)

(x, y, z, x, y, z, t)G2−→ (x,−y, z,−x, y,−z,−t) . (4.16)

74

Page 88: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

(a) First periapsis passage regions for capture trajec-tories and symmetric escape regions

(b) First 4 periapsis passage regions with symmetricescape regions, black - 1st, blue - 2nd, red - 3rd, green- 4th

Figure 4.6: Poincare maps showing symmetric capture and escape regions for J = −2.15.

Thus, applying either of the two above symmetries to a capture trajectory will

yield an escaping trajectory. An escaping trajectory is a trajectory that crosses the

initial surface of sphere with an outward velocity and exits the Hill region. Figure

4.6 shows the symmetry between the capture and escape regions on the Poincare

periapsis map for the first periapsis passage in the planar case. This figure shows the

first periapsis passage region for the capture trajectories, and hence for the escaping

trajectories this corresponds to the periapsis passage immediately preceding escape

(or transfer to the region in the vicinity of L1,2).

This symmetry between capture and escape trajectories can be extended past

75

Page 89: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

−0.5

−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

0.4

0.5

x

y L2 L

1

Forbidden Region

Forbidden Region

Figure 4.7: A capture trajectory that escapes after 4 periapsis passages. The periapsis passages areindicated by ‘x’.

the first periapsis passage. This allows us to determine when capture trajectories

can escape from the Hill region. A capture trajectory escapes the Hill region when

it passes through the initial surface of sphere with an outward velocity. However,

a captured trajectory can not escape arbitrarily. We found that it will only escape

when it passes through the ‘first periapsis passage’ escape region, which is the

escaping region symmetric to the first periapsis passage capture region as shown on

Figure 4.6(a). For this example (J=-2.15), we found that the earliest that capture

trajectories pass through this region is during their third periapsis passage, and so

we conclude that once a capture trajectory with J=-2.15 is found, it is guaranteed

to have at least 3 periapsis passages prior to escape. Such a trajectory must also

lie in the pre-image of the escape region. Due to this, these transfer regions can

be easily found by identifying regions where these symmetric regions intersect with

each other. Figure 4.6(b) is a periapsis Poincare map which shows the first four

periapsis passages of capture trajectories as well as one set of their symmetric

regions corresponding to escaping trajectories. We see that there is an overlap

of these regions, demonstrating the mechanism by which capture trajectories can

subsequently escape. An example of this is shown in Figure 4.7. This trajectory has

its third periapsis passage in the portion of the red region that overlaps with the blue

symmetric region. Therefore, its fourth periapsis passage will be in the portion of the

76

Page 90: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

green region that overlaps with the black symmetric region, as seen in Figure 4.6(b).

Any capture trajectory which has a periapsis passage in the black symmetric region

will not have any further periapsis passages and will escape. Thus, from the fact

that this trajectory had its third periapsis passage in a region that overlaps with the

blue symmetric escaping region, it will have a total of four periapsis passages and

then escape. As seen in Figure 4.7, this trajectory does indeed escape from the Hill

region after the fourth periapsis passage.

The symbolic notation defined earlier is used to give a clear description of this

result. First, let the time reversal symmetry be the operator Gt. From the property

of this symmetry:

S−(n+1) = Gt(Sn+1) = Gt(Φn(S1)) = Φ−nGt(S1) = Φ−n(S−1) . (4.17)

The condition for a trajectory x1 ∈ S1 to escape is then that it eventually lies in S−1

or Φn(x1) ∈ S−1 for some n. If this occurs Φn(S1) ∩ S−1 6= 0, or is not empty for

some n, and these trajectories escape in n − (−1) = n + 1 periapsis passages. In

Figure 4.6(b) this region is represented by the overlap of the green points and the

black points on the right side of the plot. Φ3(S1) ∩ S−1 is not empty and thus they

escape in 4 periapsis passages from their initial location in S1. Due to the uniqueness

of solutions, we also infer that the image of this set is invariant:

Φ(Φn(S1) ∩ Φ−m(S−1)) = Φn+1(S1) ∩ Φ−(m+1)(S−1) . (4.18)

In Figure 4.6(a), we immediately note that for low values of J , S1 ∩ S−1 =

S1 ∩Gt(S1) = 0 and no immediate escapes occur. We also note, however, that when

J is large enough, the set S1 crosses the y = 0 axis, meaning that S1∩Gµ(Gt(S1)) 6= 0,

where Gµ is the symmetry operator defined by Eq.(4.10), and thus direct escapes

can occur after one periapsis passage for higher J values.

Although all of the figures in this section were computed for the planar case, the

77

Page 91: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

conclusions and the symbolic notation also apply to the non-planar case where the

sets are four dimensional. In the non-planar case, analogous symmetric escaping

regions exist, and we find that capture trajectories must have a minimum of three

periapsis passages before escaping for low values of J . Figures computed for the

planar case are used to visualize the results graphically.

4.4 Safe Zones

We have previously discussed the possibility of capture trajectories eventually

escaping from the Hill region; however, we have not yet considered the trajectories

that impact with the planetary satellite. In the Poincare maps presented thus far all

trajectories have been considered, even those whose periapsis passages clearly fall

beneath the surface of the planetary satellite. To differentiate between trajectories

that impact and trajectories that don’t we introduce the term ‘safe trajectory’

to denote a trajectory that doesn’t impact with the planetary satellite or escape

from the Hill region for some specified period of time. The regions in which these

trajectories lie are denoted as ‘safe zones’. In particular, consider the first periapsis

passage region S1, and denote the set of all safe trajectories in the set the safe zone

in S1. Then a trajectory can be identified as safe based on where its first periapsis

passage occurs. Although safe zones are defined here by the first periapsis passage of

the trajectory, they could be defined by any of the periapsis passages. However, since

the periapsis passage regions become distended for subsequent periapsis passages,

the first periapsis passage region is the smallest, and hence is a logical choice for

defining the safe zones.

We first consider safe zones in the planar case, and will continue to use Europa

as our example. For this discussion, a trajectory is considered to be safe if it persists

for a non-dimensional time period of 4π, which for Europa corresponds to 7.1 days.

Figure 4.8 shows the safe zone projected into the first periapsis passage of capture

78

Page 92: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Figure 4.8: Poincare map showing the safe zone, impact and escape regions. Red - impact, Blue -escape, Black - safe.

−0.17 −0.16 −0.15 −0.14 −0.13 −0.12 −0.11 −0.1 −0.09 −0.08

0.01

0.015

0.02

0.025

0.03

0.035

0.04

0.045

0.05

x

y

surface of Europa

J=−2.15

J=−2.14

J=−2.13

J=−2.12

J=−2.11

J=−2.10

Figure 4.9: Poincare maps of planar safe zones for various values of J .

trajectories (S1) with J=-2.15. Capture trajectories with J = −2.15 and their first

periapsis passage in the black region will not impact or escape for at least one week.

The trajectories that lie in the red region will impact and the trajectories in the blue

region will escape.

Next, consider the safe zones for various values of J . The results are shown in

Figure 4.9 and are presented in a slightly different fashion. Only the safe zone is

plotted, not the entire region of first periapsis passages. Once again, these maps are

computed for 4π time units and so do not impact or escape over that period of time.

79

Page 93: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Observe that as the Jacobi integral value increases, the safe zone regions become

more distended and have a larger y value.

By characterizing the safe zones in this way, we can conclude that if a trajectory

has its first periapsis passage in the safe zone, if will be safe for at least one week. It

is then possible, for a trajectory with a particular Jacobi integral value, to determine

where its first periapsis passage should be. A trajectory that falls in this region is

potentially useful. If the goal is to perform a maneuver to place it in a more stable

trajectory but the maneuver fails, the spacecraft will remain safe for a period of

time, hopefully long enough to perform another maneuver. Another motivation for

characterizing these safe zones arises when considering low-thrust maneuvers. Since

we know where in phase space the safe zones are located, low-thrust maneuvers can

be designed such that the spacecraft travels through safe zones over the course of its

maneuver, thus guaranteeing that the spacecraft will not impact or escape.

The safe zones discussed above were computed for the planar case of the Hill

3-body problem. Analogous regions can be computed for the non-planar case;

however, it is necessary to study them in a different way. In this case, it is not

only where the periapsis passages are located in position space that is important,

but their inclinations. It is also possible to characterize the three dimensional safe

zones in terms of the radius of the trajectory at its periapsis passage along with

its inclination. However, it is not the visualization of the safe zones in the three

dimensional problem that are important, but their existence. The method used to

find safe zones for non-planar trajectories is essentially the same as the method

used in the planar case. First, capture trajectories for a particular value of J are

identified. Then, they are integrated for a period of 4π to determine whether or not

they escape from the region or impact with the planetary satellite over a one week

period for the Europa case. The capture trajectories that do not impact or escape

during that time period are defined to be the safe trajectories.

80

Page 94: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

(a) Inclination as a function of normalized ra-dius

(b) Eccentricity as a function of inclination

Figure 4.10: Characteristics of safe capture trajectories for J = −1.70.

4.5 Transferring to Stable Frozen Orbits

The preceding section defined a method to compute safe trajectories over a given

period of time. For science operations, it is necessary to transfer the spacecraft into

an orbit that is long-term stable. Recall from the previous chapter that for the

tide-only case, stable circular frozen orbits exist for inclination between 0 − 39.23o

and 140.77− 180o and stable elliptic frozen orbits exist over all inclinations. These

are therefore the stable orbits that we target for transfers.

4.5.1 Using a Safe Capture Trajectory to Initiate a Transfer

We first consider the elliptic frozen orbits. In order for this type of transfer to

be affordable in terms of ∆v, the safe trajectories and the frozen orbits must lie

close to each other in phase space. The analysis of capture trajectories so far has

been in terms of their periapsis passages. This is useful here since an ideal location

from which to initiate a transfer to a more stable orbit is at periapsis. Good

science trajectories should have low periapsis altitudes and this can be ensured by

performing the transfer at periapsis. Additionally, periapsis is the optimal place to

make changes to the orbital energy.

81

Page 95: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

When considering periapsis passages of capture trajectories from which to initiate

a maneuver, we look at all of the periapsis passages of these trajectories. Ideally,

it would be better to use a periapsis passage that occurs near the beginning of

the defined safe period, for obvious reasons. However, for simplicity, any periapsis

passage that occurs during the safe period is allowed. First consider the safe zone

for capture trajectories with J=-1.70 to see over what range of radii they occur.

Figure 4.10(a) shows that the periapsis passages occur over a wide range of radii,

all the way up to a normalized radius of almost 0.7, which corresponds to 13,773

km (the radius of Europa is 1561 km). It is therefore necessary to restrict periapsis

passages to those which have altitudes much closer to the surface. In Figure 4.10(b)

observe that the periapses of the safe capture trajectories occur over all eccentricities,

up to about 0.72.

4.5.2 Transfer to Elliptic Frozen Orbits

For transfers to elliptic frozen orbits, we restrict the periapsis passages of capture

trajectories to those that are within 250 km of the surface of Europa. Figure 4.11

shows the eccentricity as a function of inclination for periapsis passages of safe

capture trajectories that satisfy this condition for different values of J . As J

increases, the periapsis passages reach higher inclinations. In addition, restricting

the periapsis passages to fall within 250 km of the surface of Europa causes a

large reduction in the range of eccentricities over which they occur. For example,

comparing the part of Figure 4.11 that corresponds to the capture trajectories for

J=-1.70 to Figure 4.10(b), note that when the radius of periapsis is not restricted,

the periapsis passages had eccentricities in the range of e = 0 → 0.72. On the other

hand, when the radius of periapsis is restricted, the eccentricity only ranges over

about e = 0.68 → 0.72. Thus we conclude that the lower eccentricity periapsis

passages of the safe capture trajectories occur at larger radii and hence are farther

82

Page 96: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

40 50 60 70 80 90 100 110 120 130 140 1500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

inclination (degrees)

ecce

ntric

ity

J=−1.00J=−1.20J=−1.35J=−1.50J=−1.70J=−1.85frozen orbits

Figure 4.11: Periapsis passages of capture trajectories for J = −1.70 with altitudes ≤ 250 km.

from the surface of Europa.

Recall that the elliptic frozen orbits for the tide-only case have orbit elements

ω = ±π/2, and e =√

1− 5/3 cos2 i. Also note that for the larger eccentricity frozen

orbits being considered here, the gravity field of the planetary satellite does not

strongly affect the frozen orbit solutions, justifying our use of the tide-only case.

Since the goal is to transfer from a safe capture trajectory to a frozen orbit, it is

necessary to determine what characteristics the periapsis passages of the capture

trajectories must have in order to perform this maneuver. The first step is to only

consider periapsis passages of capture trajectories that have an argument of periapsis

close to the frozen orbit value. Since the frozen orbits oscillate in the unaveraged

system, the condition ω = ±90o can be relaxed somewhat to consider periapsis

passages of safe capture trajectories that fall within 5 degrees of ω = ±90o. Then, if

the periapsis passage of the safe capture trajectory has an inclination for which an

elliptic frozen orbit exists, a maneuver can be performed to change the eccentricity

so it corresponds with the inclination based on the frozen orbit relation, placing the

spacecraft in an elliptic frozen orbit.

Figure 4.12 shows the set of periapsis passages of safe capture trajectories with

83

Page 97: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

J=-1.60 that lie within 250 km of the surface of Europa and have an argument of

periapsis within 5 degrees of ±90o. A line denoting where the elliptic frozen orbits

lie is shown, which makes apparent the change in eccentricity required to transfer

from a capture trajectory to a frozen orbit. Note that the frozen orbit line actually

passes through the safe zone. This means that capture trajectories that lie on the

frozen orbit line have orbital elements corresponding to a frozen orbit at that instant.

However, a true elliptic frozen orbit is stable and bounded and would not arrive from

outside of the Hill sphere. This implies that the frozen orbit assumptions are not

valid at this point. Frozen orbits at higher eccentricities would be subject to larger

perturbations and so would not be valid either. We do not know precisely where the

averaging assumptions fail, however, we have evidence from numerical integrations

that points to the left of the frozen orbit line are still governed by the averaged

equations of motion. Therefore, when considering safe capture trajectories from

which to initiate a transfer, we do not consider periapsis passages that fall to the

right of the frozen orbit line, since this is beyond the point at which the averaging

assumptions break down.

Figure 4.12 also shows the cost, in meters per second, to transfer from the safe

capture trajectory to an elliptic frozen orbit. The periapsis passages of safe capture

trajectories used to initiate the transfer are those shown in Figure 4.12. The region

in the figure denoted ‘not valid’ refers to the region where the frozen orbits have

broken down. The region denoted ‘valid’ refers to the fact that these transfers are

practical since the frozen orbits exist. The gray region is present since we do not

have an absolute boundary on where the frozen orbits cease to exist, and so some

transfers in the gray region may be valid while others may not.

To prove that our scheme to transfer to an elliptic frozen orbit works, we

computed the transfer for one of the periapsis passages of the safe capture

trajectories and then continued the integration to show that the resulting trajectory

84

Page 98: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

48 50 52 54 56 58 60 62 64

0.6

0.65

0.7

0.75

0.8

0.85

inclination (degrees)

ecce

ntric

ity

frozen orbits

∆v

(a) (b)

Figure 4.12: Potential periapsis passages with J = −1.60 for transfers to elliptic frozen orbits andcosts. (a) Periapsis passages with altitudes ≤ 250 km ω = ±90o ± 5o. (b) Cost to transfer to acorresponding elliptic frozen orbit.

is an elliptic frozen orbit. Starting from the safe capture trajectories with J = −1.60,

as shown in Figure 4.12, we choose the periapsis passage that has the following

parameters: e = 0.718, ω = −93.66o, i = 51.14o, and a periapsis altitude of 166 km.

The eccentricity that corresponds to a frozen orbit at this inclination is 0.586, and

so the cost of performing this eccentricity change is 69 m/s. To show that the new

orbit is an elliptic frozen orbit, we plot the trajectory on contour plots of eccentricity

as a function of argument of periapsis and inclination as a function of argument of

periapsis. Figure 4.13 shows that the new orbit does stay inside the libration region

of the contour plot,verifying the successful transfer from a safe capture trajectory to

a bounded orbit close to frozen orbit conditions.

4.5.3 Transfers to Circular Frozen Orbits

Two approaches to transferring to a circular frozen orbit are considered: two-impulse

and one-impulse maneuvers. The two-impulse scheme involves performing a

maneuver at a periapsis passage of the capture trajectory that allows us to use

the dynamics of elliptic frozen orbits to achieve a lower eccentricity (which in turn

increases the inclination). Following that, a second maneuver is performed at the

85

Page 99: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−180 −160 −140 −120 −100 −80 −60 −40 −20 00

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

ω (degrees)

e

−180 −160 −140 −120 −100 −80 −60 −40 −20 00

10

20

30

40

50

60

70

80

90

ω (degrees)

i (de

gree

s)

Figure 4.13: Contour plots and the numerical results for the orbit obtained after the transfermaneuver.

much lower eccentricity to circularize the orbit. The one-impulse maneuver involves

circularizing directly from the periapsis passage of the capture trajectory. Note that

this analysis assumes that the circular frozen orbits in the tide-only case exist. For

low-altitude orbits, we showed in Chapter 3 that the planetary satellite’s gravity field

has a large effect on the motion. When the effect of the planetary satellite’s gravity

field is included, the previously circular frozen orbits become near-circular, and have

small eccentricities. However, this does not impact our analysis very strongly since

we compute costs to circularize, which can be considered as a worst-case assumption

that is more costly than transferring to a near-circular orbit.

Consider first the two-impulse approach. Previously, in order to transfer to an

elliptic frozen orbit, a maneuver was chosen to achieve the frozen orbit eccentricity.

By examining the left plot in Figure 4.13 we see that if a slightly higher eccentricity

is chosen (i.e. an eccentricity near the top of the libration region), much larger

variations in eccentricity are achieved, to the point where the eccentricity occasionally

gets very close to 0. Figure 4.14 shows this phenomenon, where instead of performing

a maneuver to change the eccentricity to 0.586 as above, the eccentricity is changed

to 0.687. The same periapsis passage as in the elliptic frozen orbit transfer example

is used (i.e, e = 0.718, i = 51.14o, ω = −93.66o, altitude=166km). The cost of this

86

Page 100: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

maneuver is 38.6 m/s. The next step in this approach is to circularize the orbit.

Depending on the altitude of the orbit after it is circularized, some control maneuvers

may be necessary to keep the orbit from drifting too much. The frequency and size

of these maneuvers will depend on how much of a drift we can tolerate and how

quickly the orbit is drifting (depending on its altitude).

The circularization maneuver is performed at a periapsis passage of the orbit

shown in Figure 4.14. In order to understand how the radius, eccentricity and

inclination of the orbit vary and where the periapsis passages occur, refer to

Figure 4.15 which shows the radius, eccentricity and inclination as functions of

time. The periapsis passages are indicated with a red ‘x’ and the solid blue line in

the top plot represents the surface of Europa. Observe that a lower eccentricity is

correlated to a higher radius and higher inclination. Higher inclinations are more

desirable; however, the higher radius associated with the higher inclination will lead

to an initially large circular orbit. To compare these issues, two examples will be

provided: one where the circularization maneuver is performed from the minimum

eccentricity of the orbit shown in Figure 4.14 and one where the circularization is

performed from a higher eccentricity. In both cases, the radius of the orbit will be

relatively large and some control maneuvers will be necessary to maintain this orbit.

Table 4.1 shows the cost of achieving both of these circular orbits as well as their

average radius and initial inclination. Figure 4.16 shows their characteristics.

We now consider the one-impulse approach for transferring to a circular orbit.

This approach involves directly circularizing the orbit at the periapsis passage of

the capture trajectory. The cost of this maneuver is 444.5 m/s and the resulting

orbit has a radius of 1731 km which corresponds to an altitude of 166 km and an

inclination of 51.14o.

All of the transfers to circular frozen orbits computed thus far result in orbits

that have different radii and inclinations. To compare them, it is necessary to

87

Page 101: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 4.1: Comparison of costs to transfer from a capture trajectory to a circular frozen orbit.

Example 1a Example 2b

From capture trajectory to near-frozen elliptic orbit 38.6 m/s 38.6 m/s

Circularization 10.3 m/s 230.3 m/s

Total Cost 48.9 m/s 268.9 m/s

Average Radius 5300 km 2875 km

Initial Inclination 61.6o 57.9o

a Circularizing at the minimum eccentricity of the near-frozen elliptic orbitb Circularizing at higher eccentricity

−180 −160 −140 −120 −100 −80 −60 −40 −20 00

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

ω (degrees)

e

−180 −160 −140 −120 −100 −80 −60 −40 −20 00

10

20

30

40

50

60

70

80

90

ω (degrees)

i (de

gree

s)

Figure 4.14: Contour plots and numerical results for a transfer orbit to a circular frozen orbit.

88

Page 102: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 5 10 15 20 250

0.2

0.4

0.6

0.8

e

0 5 10 15 20 2545

50

55

60

65

i (de

gree

s)

0 5 10 15 20 250

5000

10000

r (k

m)

time (days)

Figure 4.15: Characteristics of the near-frozen elliptic transfer orbit as functions of time. Eachperiapsis passage is represented by an ‘x’ and the solid line in the top plot represents the surface ofEuropa.

0 1 2 3 4 5 6 70

0.05

0.1

0.15

0.2

e

0 1 2 3 4 5 6 758

60

62

64

i (de

gree

s)

time (days)

0 1 2 3 4 5 6 74500

5000

5500

6000

r (k

m)

(a)

0 1 2 3 4 5 6 70

0.01

0.02

0.03

e

0 1 2 3 4 5 6 755

56

57

58

i (de

gree

s)

time (days)

0 1 2 3 4 5 6 72800

2850

2900

2950

r (k

m)

(b)

Figure 4.16: Characteristics of both possible circularized orbits over one week. (a) Transfer fromthe minimum eccentricity of the near-frozen elliptic orbit. (b) Transfer from a higher eccentricityof that same orbit.

89

Page 103: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 4.2: Comparison of costs to transfer from capture trajectory to tightly bound circularorbit

Method

Using Dynamics to Circularize Direct Circularization

Example 1 (i = 61.6o) 591.5 m/s 692.2 m/s

Example 2 (i = 57.9o) 586.6 m/s 604.7 m/s

a Circularization takes place at the minimum eccentricity of near-frozen elliptic orbitb Circularization takes place at a higher eccentricity

consider additional transfers so that a common orbit is achieved for both approaches.

Two common orbits are considered, with inclinations of 61.6o and 57.9o and a radius

of 1731 km. The two inclinations correspond to the inclinations achieved by the

two examples of the two-impulse approach and the radius corresponds to the radius

achieved by the one-impulse approach. The orbit obtained by the one-impulse

approach has an inclination of 51.14o and so to transfer to inclinations of 61.6o and

57.9o costs 247.7 km/s and 160.2 km/s, respectively. The radii of the orbits achieved

by the examples for the two-impulse approach are about 5300 km and 2875 km and

the costs of the Hohmann transfers to reduce those radii to 166 km are 542.6 m/s

and 299.7 m/s respectively. Table 4.2 shows a comparison of the costs of the two

approaches. We see that the method which uses the dynamics of the system in the

vicinity of the elliptic frozen orbits is more efficient, and the difference between the

two methods is larger when the final inclination is larger.

If the total costs of only the two-impulse approach are compared, we see that

although the second example is slightly cheaper, an inclination of almost 4o lower

is achieved. If the Hohmann transfer component of this approach is neglected, we

see that while the second example achieves a slightly smaller circular frozen orbit, it

has a cost of 268.9 km/s which is 5.5 times greater than the transfer via the first

example (see Table 4.1). Both circular orbits are stable there so is a clear advantage

90

Page 104: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

to following the first example. Since the cost to achieve the stable circular orbit is

quite small (less than 50 m/s) it is possible that a low-thrust vehicle could achieve

this orbit. Following that, the low-thrust vehicle could transition down to a tighter

circular orbit with a spiral maneuver.

As another approach to transferring from a safe capture trajectory to a circular

frozen orbit, we examine the possibility of performing a simultaneous circularization

and plane change maneuver. It is not possible to do this maneuver with the current

example, since for this maneuver to be feasible the argument of periapsis at the

periapsis passage must be either 0o or 180o and for our current example the argument

of periapsis is approximately −90o. Therefore, we look back to our data of all

the periapsis passages of capture trajectories at this energy level, J = −1.60, and

examine their arguments of periapsis. Figure 4.17(a) shows argument of periapsis

plotted as a function of eccentricity for all of the periapsis passages of the capture

trajectories with J = −1.60. This plot includes periapsis passages with any radius

of periapsis. However, as before, our goal is a final orbit with a low altitude.

Therefore, restrict the periapsis passages to those with an altitude less than 250 km.

Figure 4.17(b) also shows the argument of periapsis as a function of eccentricity

for the periapsis passages that satisfy this condition. We see that the argument of

periapsis values lie between about 50 to 150o and −150 to −50o. Therefore, there

are no periapsis passages with altitudes close to the surface for which a simultaneous

circularization and plane change maneuver can be performed.

We conclude, from comparing the two plots in Figure 4.17, that the periapsis

passages with lower arguments of periapsis must have a large radius. We also see

that these periapsis passages have lower eccentricities. This is not surprising with the

doubly averaged dynamics in mind. In the doubly averaged system, the semi-major

axis is constant on average. Therefore, since the capture trajectories originate far

from the planetary satellite, for them to have a periapsis passage with a low altitude,

91

Page 105: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

(a) (b)

Figure 4.17: ω as a function of eccentricity for safe capture trajectories with J = −1.60. (a) Allperiapsis passages. (b) Periapsis passages with altitude ≤ 250 km.

the eccentricity must be large. The numerical results support this since when all

periapsis passages are considered, they range in eccentricity from ∼ 0.32 → 0.75, but

when only those with altitudes less than 250 km are considered, the eccentricity only

ranges between about 0.705 → 0.75. Therefore, by limiting the radius of periapsis,

only periapsis passages with large eccentricities remain. Considering the averaged

dynamics once again, note that for a trajectory to reach a large eccentricity it must

be circulating about a libration region and libration regions are centered about

±90o. Then, examining Figure 4.17(b) it is clear this is the case. The arguments

of periapsis in this plot extend about ±50o on either sides of ±90o. We therefore

conclude that it is not dynamically possible to have a periapsis passage of a capture

trajectory that will allow a simultaneous circularization and plane change maneuver

to place the trajectory in a tightly bound circular orbit.

92

Page 106: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

CHAPTER 5

THE DESIGN OF LONG LIFETIME SCIENCE

ORBITS

The requirements for a science orbit about a planetary satellite are generally a

near-polar, low-altitude orbit so that the science requirements of the mission, such

as imaging the surface and measuring various features of the planetary satellite,

are satisfied. Since the environment in the vicinity of many planetary satellites,

such as Europa, is highly perturbed, it is important for a science orbit to have a

long lifetime. For example, the nominal length of the science phase of a mission to

Europa could be one month, but it is essential to ensure that the spacecraft will not

impact with the planetary satellite if a particular control maneuver fails. Therefore,

the goal of long lifetime orbit design for Europa is lifetimes of around 100 days. This

will ensure that the spacecraft will be safe while awaiting the next possible maneuver

opportunity.

From Chapter 3, we have seen that stable frozen orbits exist at small inclinations

for low-altitude circular orbits (in the tide + J2 + C22 + J3 model) and at larger

altitudes (since the orbits are elliptic) for near-polar orbits (in tide-only and non

tide-only models). However, neither of these types of stable frozen orbits satisfy the

science requirements. Thus, we look towards unstable frozen orbits as a means for

designing long lifetime orbits that satisfy the science requirements. If a spacecraft is

arbitrarily placed in a near-polar, low-altitude orbit its lifetime is somewhat random

and can range from very short to very long. See, for example, Figure 5.1 which

93

Page 107: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 20 40 60 80 100 120 140 160 1800

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

time (days)ec

cent

ricity

impact

Figure 5.1: Eccentricity as a function of time for trajectories with an initial altitude of 100 km,initial inclination of 70o and different initial argument of periapsis values.

shows the eccentricity as a function of time for a few trajectories in the Europa

system. These trajectories have an initial altitude of 100 km, an inclination of 70o

and different initial argument of periapsis values. Their lifetimes vary from 40-165

days, demonstrating that arbitrarily placing a spacecraft in a low-altitude near-polar

orbit can result in impact after a relatively short period of time.

It will be shown later in this chapter that integrating a near-circular, near-polar

unstable frozen orbit in the 3-DOF system (including J2, C22 and J3) does not yield

a very long lifetime orbit. However, the structure of the dynamics in the vicinity

of these unstable frozen orbits helps in the design of long lifetime trajectories.

Specifically, since they have non-zero eccentricities, their stable and unstable

manifolds in the 1-DOF system are shifted and this causes variations in the rate of

change of the eccentricity along their paths. These variations can be exploited by

choosing the path that remains in the vicinity of a frozen orbit for the longest period

of time. If a trajectory in the full 3-DOF system follows this path, it can have a

relatively long lifetime on the order of at least 100 days, as compared to lifetimes as

short as 40 days for randomly chosen initial conditions.

After identifying the desired trajectory in the 1-DOF system, we develop a

94

Page 108: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

method to systematically compute initial conditions in the 3-DOF system such that

the trajectory in that system will follow the 1-DOF system on average. To do this,

we use all three systems identified thus far, namely, the 1, 2 and 3-DOF systems.

We start with a point on the stable manifold of a frozen orbit in the 1-DOF system.

We then linearize the 1-DOF system about that point on the manifold and obtain

corrections such that the initial conditions lead to a trajectory in the 2-DOF system

that has motion in the 1-DOF system as its average. This procedure is repeated on

the 3-DOF system, by linearizing about the corrected point in the 2-DOF system.

The result is a set of initial conditions in the 3-DOF system that, when integrated,

follow the 1-DOF system on average.

A toolbox to compute long lifetime science orbits using the methods described in

this chapter was written in Matlab and delivered to the Jet Propulsion Laboratory

so they could use it to design these orbits. It is described at the end of this chapter.

5.1 Dynamics in the 1-DOF System

The frozen orbits that will be the basis of our analysis in this chapter are the

near-polar, near-circular unstable frozen orbits that exist in the tide + J2 + C22 +

J3 model. They are indicated by the heavy dashed line in Figure 5.2. Although

these frozen orbits have constant orbital elements on average in the 1-DOF system,

they are unstable, and so initializing at one in the 3-DOF system results in the

trajectory diverging from the frozen orbit location. Recall Figure 3.16, where we see

the trajectory in the (e,ω) plot diverge from its initial frozen orbit location. The

trajectory will impact with the planetary satellite after a short period of time since,

as seen in the plot, the eccentricity increases very rapidly leading to a decrease in

the radius of periapsis. Thus, the unstable frozen orbits can’t be used directly as

long lifetime orbits.

Since the 1-DOF system is Hamiltonian and the near-circular, near-polar frozen

95

Page 109: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0

20

40

60

80

100

120

140

160

180

esinω

incl

inat

ion

(deg

rees

)

Figure 5.2: Frozen orbit solutions: unstable near-circular, near-polar orbits indicated by heavydashed line.

orbits are unstable, they have stable and unstable manifolds. The stable manifolds

approach the frozen orbit and the unstable manifolds depart from the frozen orbit.

If a spacecraft is initialized on the stable manifold of a frozen orbit, we expect

that it will drift towards the frozen orbit location on the stable manifold, and then

follow the unstable manifold towards impact with the planetary satellite. A good

understanding of the manifolds can help us determine where exactly to initialize

a spacecraft in order to obtain the maximum possible lifetime. Note that these

manifolds exist only in the 1-DOF system, but their existence here impacts dynamics

at higher degrees of freedom.

The manifolds of the frozen orbits are determined by computing the contour

of constant potential, R, that passes through the frozen orbit. This represents the

motion in the 1-DOF system corresponding to a trajectory that passes through the

frozen orbit. The manifolds are plotted in the three different regimes, (h, k), (G, ω)

and (e, ω). Figure 5.3 shows the stable and unstable manifolds for a frozen orbit

with a 70 degree inclination. In all three plots, the unstable manifolds are numbered

I and III and the stable manifolds are numbered II and IV. We can see most clearly

from Figure 5.3(b) that there are four possible paths along the manifolds that pass

96

Page 110: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

through the frozen orbit, identified with an ‘x’. By examining the possible paths

along the manifolds in Figure 5.3(c) the number of degrees of ω covered along each

path can be determined. For this computation, only the portion of the trajectory

that is near-circular is considered. Once the trajectory departs from its near-circular

value, the radius of periapsis will decrease to below the surface of the planetary

satellite since the semi-major axis remains constant. The four possible paths and

approximate number of degrees of ω covered along each path are as follows:

1. II→ I: 270 degrees

2. II→ III: 180 degrees

3. IV→ I: 180 degrees

4. IV→ III: 90 degrees

Therefore, by beginning on the stable manifold labelled II, passing through the

frozen orbit and then continuing on the unstable manifold labelled I, 270 degrees of

ω are covered while the orbit is near-circular. This is the path that will produce

the longest lifetime orbit in the 3-DOF system. Different ranges of ω are covered

along different paths because the frozen orbits are not circular. Figure 5.4 shows

the manifolds in (h,k) space with a circle denoting where the radius of periapsis is

at the surface of Europa. Therefore, values of (h,k) outside of the circle represent

radii of periapsis below the surface of Europa and hence impact. Different distances

along the manifold are covered inside of the circle along the different paths since the

frozen orbit located at the center of the ‘x’ is not at (0,0). A longer distance inside

the circle means a slower rate of change of the eccentricity which leads to a longer

lifetime. In the 3-DOF system, the trajectory will not follow the manifolds exactly.

It is possible, however, to determine initial conditions such that this path along the

manifold is followed on average and very long lifetime orbits are attained.

97

Page 111: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 250 300 3500

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

ω (degrees)

ecce

ntric

ityII I

IV III

(a) (e,ω)-space

−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

k=ecosω

h=es

inω

I II

III IV

(b) (h,k)-space

−2000 −1500 −1000 −500 0 500 1000 1500 2000

−2000

−1500

−1000

−500

0

500

1000

1500

2000

Gcosω

Gsi

II I

IV III

(c) (G,ω)-space

Figure 5.3: Stable and unstable manifolds for a frozen orbit, identified with an ’x’, withe∗ = 0.0129, i∗ = 70o, ω∗ = −90o and a∗ = 1682.5 km.

98

Page 112: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.25 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2 0.25−0.25

−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

0.2

0.25

k=ecosω

h=es

inω

I II

III IV

impact

Figure 5.4: Expansion of the central region in 5.3(b) with Europa impact circle

In the tide-only model, the manifolds of the circular unstable frozen orbits have

the same structure, with the only difference being that the frozen orbit is located at

the origin. Therefore, there is no advantage to choosing the II→I path as opposed to

the II→III path. When this design method is applied to the Saturn moons Enceladus

and Dione later in this chapter, the tide-only model is used since the gravity fields of

those moons are not known. This means that the advantage in terms of lifetime of

a non-circular frozen orbit doesn’t apply. Therefore, the lifetimes obtained for these

moons are conservative estimates of what would be possible for a real mission where

the gravity field coefficients would probably be known by the time the science phase

of the mission takes place.

These manifolds explain why when a trajectory with initial conditions

corresponding to a frozen orbit is integrated in the unaveraged system, it does not

have a particularly long lifetime. As Figure 5.5 shows, when the trajectory that

is initialized at the frozen orbit location is integrated, it immediately follows the

unstable manifold path III since the equilibrium does not exist in the 3-DOF system.

Additional lifetime is not gained since the trajectory is not initialized on the stable

manifold. This trajectory has a lifetime of 44 days.

99

Page 113: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.1 −0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08 0.1−0.1

−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

0.1

k=ecosω

h=es

inω

Figure 5.5: 3-DOF system integration of a frozen orbit with e∗ = 0.0129, i∗ = 70o, ω∗ = −90o anda∗ = 1682.5 km with its stable and unstable manifolds.

5.2 Computing Initial Conditions in the 3-DOF

System

To this point, we have discussed reducing the 3-DOF system to a 1-DOF system

in order to have a better qualitative understanding of the overall dynamics. From

this system, equilibrium solutions are identified and motion in their vicinity is

studied. However, as seen in the previous section, using the frozen orbits as initial

conditions in the full 3-DOF unaveraged system does not produce very long lifetime

trajectories. We have already shown that the longest lifetime orbits begin on the

stable manifold of the frozen orbits, move towards the frozen orbits and then move

off on the unstable manifold. The question then becomes how to determine initial

conditions in the 3-DOF system such that the trajectory will follow the manifolds in

the 1-DOF system. The easiest first guess would be to initialize the trajectory with

initial conditions that lie exactly on the 1-DOF manifold.

Computing the point on the manifold is fairly straightforward. First a frozen

orbit is chosen and the integrals of motions in the 1-DOF system, θ = cos2 i∗(1−e∗2),

R∗ = R(e∗, i∗, ω∗) are computed. To compute a point on the manifold, start with

a desired initial eccentricity em. Then, having already computed θ from the frozen

100

Page 114: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

orbit conditions, the inclination on the manifold, im is obtained from:

im = cos−1

√θ

1− e2m

. (5.1)

Then, substituting em and im into

R∗ = R(em, im, ωm) , (5.2)

we obtain the following quadratic equation in sin ωm:

0 =− 15N2a2

8e2

m sin2 im sin2 ωm +3µJ3

2a4(1− e2m)5/2

em sin im (1

−5

4sin2 im

)sin ωm +

N2a2

4

[(1− 3

2sin2 im

)(1 +

3

2e2

m

)(5.3)

+15

4e2

m sin2 im

]− R∗ ,

which can be easily solved for ωm. Continuing with the frozen orbit considered

thus far with e∗ = 0.0129, i∗ = 70o and ω∗ = 270o, we compute a point on the

1-DOF manifold with the following characteristics: em = 0.02, im = 69.998o and

ωm = 170.97o.

As Figure 5.6 shows, although the trajectory initialized with the above conditions

and integrated in the 3-DOF system does follow the manifolds, it is not centered on

them and its lifetime is 87 days. This occurs because there are oscillations due to

both the motion of Europa about Jupiter and the spacecraft about Europa that are

not accounted for in the 1-DOF system. As well, the effect of C22 is not accounted

for in the 1-DOF system since averaging cancels it. The ideal trajectory would be

centered on the manifold since such a trajectory would spend the longest time in

the vicinity of the frozen orbit, giving it the longest possible lifetime. Therefore, we

now discuss a method that starts with a point in the 1-DOF system, pulls the point

back to a point in the 2-DOF system and concludes by pulling back that point to

initial conditions to be used in the 3-DOF system. These initial conditions take into

101

Page 115: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

(a)

−0.021 −0.02 −0.019 −0.018 −0.017 −0.016 −0.015 −0.014 −0.013

−4

−2

0

2

4

6

8

x 10−3

k=ecosω

h=es

inω

(b)

Figure 5.6: (a) Integration of trajectory in 3-DOF system, initialized on 1-DOF stable manifold.(b) Expansion of the initial part of the trajectory.

account all of the oscillations and produce a trajectory that follows the manifolds in

the 1-DOF system very closely.

The method used to relate both the 2-DOF system to the 1-DOF system and the

3-DOF system to the 2-DOF system is a linearization method. It is performed two

times, first on the 2-DOF system and then on the 3-DOF system. The method will

first be described in general terms, and then some specifics for each system will be

discussed.

Let x be a vector containing the orbital elements under consideration. Then,

the time derivatives of the orbital elements, obtained from the Lagrange Planetary

Equations, can be expressed as follows:

x′ = g0(x) + g1(x, t) , (5.4)

where ′ denotes the time derivative, g0(x) is a vector containing secular (averaged)

components of the equations, which are time-invariant and g1(x, t) is a vector

containing the time-periodic components of the equations, with period T. We assume

that |g1| |g0|. Eq.(5.4) is linearized about a point x0 in the time-invariant system

102

Page 116: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

such that x′0 = g0(x0). Then, let x = x0 + δx where δx is a small perturbation from

the point. Substituting this into Eq.(5.4), we obtain:

(x0 + δx)′ = g0(x0 + δx) + g1(x0 + δx, t) . (5.5)

We perform a Taylor series expansion on the right hand side of Eq.(5.5), to first

order in the time-invariant terms and to zeroth order in the time-periodic terms and

obtain:

x′0 + δx′ = g0(x0) +∂g0(x0)

∂x

∣∣∣∣x0

δx + g1(x0, t) + · · · . (5.6)

The expansion is carried to only the zeroth order in the time-periodic terms since

the first order terms are much smaller than the zeroth order terms. This allows for a

straightforward analysis of the system. The linearized system can be reduced to the

following since x′0 = g0(x0):

δx′ = Aδx + B(t) , (5.7)

where higher order terms are ignored and

A =∂g0(x0)

∂x

∣∣∣∣x0

, (5.8)

B(t) = g1(x0, t) . (5.9)

Note that the motion of x0 is ignored since we are only interested in evaluating the

system over a short time span. Therefore, A is a constant matrix and B(t) is a

time-periodic vector. We can solve for δx as follows:

δx = eA(t−t0)δx0 +

∫ t

t0

eA(t−τ)B(τ)dτ . (5.10)

Our goal for the linearization is to obtain initial conditions δx0 such that the

motion in the original system has the motion in the time-invariant system as its

average. To do this, we take the average of Eq.(5.10) over one period T and set it

103

Page 117: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

equal to zero.

0 =1

T

∫ T

0

eA(t−t0)dtδx0 +1

T

∫ T

0

∫ t

t0

eA(t−τ)B(τ)dτdt . (5.11)

Let

D =1

T

∫ T

0

eA(t−t0)dt , (5.12)

E =1

T

∫ T

0

∫ t

t0

eA(t−τ)B(τ)dτdt . (5.13)

Then, we solve for δx0 from Eq.(5.11) to obtain

δx0 = −D−1E . (5.14)

Therefore, if the original system is integrated with initial conditions x0 + δx0, the

motion should be, on average, the time-invariant system initialized at x0.

5.2.1 From 1-DOF System to 2-DOF System

In this section the algorithm is used to compute initial conditions δx0 in the

2-DOF system that will produce motion that follows the 1-DOF system on average.

Since the semi-major axis is constant, consider only x =

[e i ω

]T

. The time

derivatives of the orbital elements in the 2-DOF system are obtained by using R in

Eqs.(3.4)-(3.6):

de

dt=

1− e2

na2e

∂R

∂M−√

1− e2

na2e

∂R

∂ω, (5.15)

di

dt=

1√1− e2na2

(cot i

∂R

∂ω− csc i

∂R

∂Ω

), (5.16)

dt=

√1− e2

na2e

∂R

∂e− cot i√

1− e2na2e

∂R

∂i. (5.17)

These equations can be broken up into two components where the g0 components

correspond to the LPE applied to R and the g1 components correspond to the

LPE applied to R − R. Then, the g0 components are time-invariant and the g1

104

Page 118: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

components are time-periodic. The time periodicity appears as periodicity in Ω

with a period of π. Therefore, to simplify the system, rewrite Eqs. (5.15)-(5.17)

as derivatives with respect to Ω rather than derivatives with respect to time by

multiplying the right-hand sides of the equations by −(1/N).

Since our goal is to obtain long lifetime orbits, the 2-DOF system is linearized

about a point on the stable manifold of a frozen orbit in the 1-DOF system. Let this

point be xm0 =

[e0 i0 ω0

]T

, with Ω0. Then, A and B(Ω) as defined in Eqs.(5.8)

and (5.9) are computed fairly easily analytically. Following that, D is computed

using the Romberg quadrature algorithm [10] and E is computed using the Gaussian

quadrature for double integrals algorithm [10] to determine the corrections to xm0 ,

δxm0 . The Romberg quadrature algorithm is described in Burden and Faires[10] for a

scalar. In this case however, the integral that needs to be computed is the integral

of each term in the matrix D. Therefore, the algorithm is repeated for each entry in

D. The specific algorithms used to to compute each term in D and each term in E

are given in Appendix A.

We illustrate this method with an example. First, choose the frozen orbit with

i∗ = 70o, e∗ = 0.0129, ω∗ = 270o and Ω0 = 0. Then, compute the point on the stable

manifold of this frozen orbit with the following characteristics:

xm0 =

0.02

69.99756766o

170.97339256o

Then, by following the linearization algorithm, we obtain the following corrections

to the initial conditions:

δxm0 =

0.00027136

0.86310567o

0.21413018o

To verify that the 2-DOF trajectory, initialized at xm

0 + δxm0 , has as its average the

105

Page 119: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.80.017

0.0175

0.018

0.0185

0.019

0.0195

0.02

0.0205

time (days)

ecce

ntric

ity

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.869

69.5

70

70.5

71

time (days)in

clin

atio

n (d

eg)

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8170

171

172

173

174

175

176

177

time (days)

ω (

deg)

Figure 5.7: Eccentricity, inclination and argument of periapsis as functions of time. Solid lines -2-DOF trajectory, initialized at xm

0 + δxm0 , dashed lines - 1-DOF trajectory, initialized at xm

0 .

1-DOF trajectory initialized at xm0 , both the 2-DOF and 1-DOF trajectories are

plotted. As Figure 5.7 shows, the dashed lines representing the 1-DOF trajectory

pass through the center of the solid curves which represent the 2-DOF trajectory.

5.2.2 From 2-DOF System to 3-DOF System

In this section, initial conditions in the 3-DOF system that will produce motion

that follows the 2-DOF system on average are computed. Since the results from the

previous section are incorporated, the resulting motion also follows the 1-DOF system

on average. In this case, there are six equations and x =

[a e i ω Ω σ

]T

where σ = M − nt and n is the mean motion of the spacecraft about the planetary

satellite. The time derivatives of these orbital elements are those stated in

106

Page 120: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Eqs.(3.4)-(3.7) plus the equation for σ taken from [8]:

da

dt=

2

na

∂R

∂M, (5.18)

de

dt=

1− e2

na2e

∂R

∂M−√

1− e2

na2e

∂R

∂ω, (5.19)

di

dt=

1√1− e2na2

(cot i

∂R

∂ω− csc i

∂R

∂Ω

), (5.20)

dt=

√1− e2

na2e

∂R

∂e− cot i√

1− e2na2e

∂R

∂i, (5.21)

dt=

1√1− e2na2 sin i

∂R

∂i, (5.22)

dt= − 2

na

∂R

∂a− 1− e2

na2e

∂R

∂e. (5.23)

For this system, the g0 components correspond to the LPE applied to R and the g1

components correspond to the LPE applied to R− R. Therefore, the g0 components

are time-invariant and the g1 components are time-periodic. In this case, the time

periodicity appears in terms of the true anomaly, ν, with a period of 2π which

corresponds to one period of the spacecraft about Europa. Since the true anomaly

is a function of the mean anomaly, the equations are expressed as derivatives with

respect to mean anomaly rather than with respect to time to simplify the analysis.

To transform from time derivatives to derivatives with respect to mean anomaly, the

equations must be multiplied by 1/n.

Since our goal is to determine initial conditions such that the 3-DOF trajectory

has the 2-DOF trajectory as its average, we linearize about a point in the 2-DOF

system. In particular, we linearize about the point in the 2-DOF system that

we obtained in the previous section and let this point be x0 = xm0 + δxm

0 . The

constant matrix A is still straightforward to compute analytically even though it is

a 6x6 matrix, and D is once again computed using the Romberg algorithm given

in Appendix A. However, the vector B, which in this case is B(ν(M)), cannot be

computed analytically since Kepler’s equation must be solved to determine the true

107

Page 121: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

anomaly, ν, as a function of the mean anomaly, M . Therefore, we turn to the

computation of E. The computation of E where

E =1

∫ 2π

0

∫ M

0

eA(M−τ)B(ν(τ))dτdM (5.24)

can still be accomplished using the double integral version of Gaussian quadrature

again, but the algorithm must be modified to solve Kepler’s equation, and hence

compute B at each step. This modified algorithm to compute E term-by-term is

also given in Appendix A. The corrections to x0, δx0, are then computed.

We now continue with the example from the previous section to show that when

integrated with the initial conditions obtained from the linearized system, the 3-DOF

system follows the 2-DOF system on average. The point that is linearized about,

which is also the initial condition for the 2-DOF system, is

x0 = xm0 + δxm

0 =

1682.53057218 km

0.02027136

70.860673330

171.18752273o

0o

0o

,

The corrections to these initial conditions, δx0 are found to be:

δx0 =

2.98938278 km

0.00219834

0.00239633o

0.17321674o

−0.00679549o

−0.07757915o

.

Figure 5.8 shows plots of the integrated 3-DOF trajectory, initialized at x0 +δx0, and

108

Page 122: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 250 300 350 4001679

1680

1681

1682

1683

1684

1685

1686

sem

i−m

ajor

axi

s (k

m)

mean anomaly (deg)

0 50 100 150 200 250 300 350 4000.017

0.018

0.019

0.02

0.021

0.022

0.023

ecce

ntric

itymean anomaly (deg)

0 50 100 150 200 250 300 350 40070.81

70.82

70.83

70.84

70.85

70.86

70.87

70.88in

clin

atio

n (d

eg)

mean anomaly (deg)

0 50 100 150 200 250 300 350 400164

166

168

170

172

174

176

178

ω (

deg)

mean anomaly (deg)

Figure 5.8: Semi-major axis, eccentricity, inclination, and argument of periapsis as functions ofmean anomaly. Solid lines - 3-DOF trajectory, initialized at x0 + δx0. Dashed lines - 2-DOFtrajectory, initialized at x0.

the integrated 2-DOF trajectory, initialized at x0. In all four orbital element plots,

the 2-DOF trajectory cuts through the center of the 3-DOF trajectory. Therefore,

the goal of the 3-DOF motion following the 2-DOF motion on average has been

accomplished.

5.3 Long Lifetime Orbits in the 3-DOF System

Now that we have generated initial conditions in the 3-DOF system that follow

the manifold in the 1-DOF system on average, it is necessary to verify that these

initial conditions actually generate long lifetime orbits. Consider, for example,

Figure 5.9 which shows the integration, in the 3-DOF system, of the initial conditions

109

Page 123: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08

−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

impact

Figure 5.9: Integration in the 3-DOF of a trajectory initialized at x0 with the 1-DOF manifolds.

generated in the previous section. Although it was shown in the previous section

that the 3-DOF motion did follow the 1-DOF on average, the trajectory will not, in

general, follow the 1-DOF manifold exactly since the linearization used is only an

approximation to the true motion. When several examples of potential long lifetime

orbits are computed, we find that sometimes the correct manifold path is followed

and sometimes it is not. In all of the cases, the trajectory does not have the desired

manifold path as its exact average. For the case in Figure 5.9, the trajectory follows

the wrong manifold path and has a lifetime of 84 days. This is not unexpected since

there should about a 50/50 chance of the trajectory lying on one side of the manifold

or the other.

To ensure that the trajectory always follows the desired manifold path a small

correction in the argument of periapsis, called the bias, is introduced. We show that

it is possible to find a particular argument of periapsis correction that gives the

maximum long lifetime orbit for a specific set of initial conditions. For the example

in Figure 5.9, the bias in the argument of periapsis that produces the longest lifetime

orbit is -0.0067 radians (−0.38o). This maximum lifetime orbit follows the 1-DOF

manifold exactly. However, from a practical mission standpoint, it is not that useful

110

Page 124: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 20 40 60 80 100 120 140 160 1800

0.02

0.04

0.06

0.08

time (days)

ecce

ntric

ity

Figure 5.10: Time history of the eccentricity for a long lifetime orbit.

to find the maximum lifetime orbit. This is because the uncertainty in the orbit’s

position is, in general, greater than the accuracy with which the maximum lifetime

orbit is determined. In addition, since the maximum lifetime orbit is one that follows

the 1-DOF manifold very closely, a very small error in the initial position could

push the trajectory to the wrong manifold path. The details regarding the orbit

uncertainty will be discussed in the next chapter. For now, we focus on finding long

lifetime orbits that follow the desired manifold path, but do not necessarily have

maximum lifetimes.

5.3.1 Minimum Eccentricity on the Manifold

In the example that has been discussed thus far, the initial eccentricity on the

manifold was chosen arbitrarily to be 0.02 since the only requirement for an initial

eccentricity is that it be on the stable manifold of the long lifetime orbit. There

is however, a more systematic way to choose the initial eccentricity. Consider

Figure 5.10 which shows the time history of the eccentricity for a long lifetime orbit

corresponding to a 70o inclination frozen orbit. Note in this plot that the eccentricity

decreases rapidly before it settles around the frozen orbit value where it stays almost

constant for the majority of the orbit. Observe further that just before the constant

value is reached, the eccentricity goes through a minimum. It is this minimum

eccentricity value that will be used from now on to initialize long lifetime orbits.

One advantage of using this minimum eccentricity value is that the period of time

111

Page 125: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.03 −0.02 −0.01 0 0.01 0.02 0.03−0.03

−0.02

−0.01

0

0.01

0.02

0.03

k=ecosω

h=es

inω

emin

frozenorbit

manifoldpoint

Figure 5.11: 1-DOF manifolds with minimum eccentricity circle (dashed curve).

before the trajectory reaches this value is quite short compared the amount of time

it spends oscillating about the frozen orbit. Thus, not much orbital lifetime is gained

by starting at a higher eccentricity value. In addition, always initializing trajectories

at the minimum eccentricity for the frozen orbit under consideration provides a good

way to compare the lifetimes of the orbits. Finally, in terms of controlling the orbit,

initializing at the minimum eccentricity provides advantages that will be shown in

the following chapter.

The minimum eccentricity on the stable manifold can be computed geometrically.

First, consider Figure 5.11. The line connecting the frozen orbit point to the

manifold point is the linear approximation of the manifold in that region. Although

the manifold in the 1-DOF system is governed by a very complicated function, it can

be approximated by a line in the vicinity of the frozen orbit, as seen in Figure 5.3(b).

The manifold point is a point that has the same eccentricity as the frozen orbit and

an argument of periapsis consistent with the 1-DOF manifold (computed using the

method in section 5.2). The minimum eccentricity then lies on this line and is the

point with the minimum distance from the origin, as shown in Figure 5.11. Let the

frozen orbit point be (k∗,h∗) and the manifold point be (km,hm). Then, the equation

112

Page 126: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

for the line connecting the two points is:

h =

(hm − h∗

km − k∗

)(k + km)− hm , (5.25)

and the minimum distance from the origin to this line is[43]:

d =

∣∣∣∣∣∣∣det

km − k∗ k∗

hm − h∗ h∗

∣∣∣∣∣∣∣√

(hm − h∗)2 + (km − k∗)2. (5.26)

Since, in the (k,h) plot, the eccentricity is equivalent to the radial distance from the

origin to a point, the minimum eccentricity is given by the distance d in Eq.(5.26).

Then, the corresponding argument of periapsis on the manifold can be computed

by the method in section 3.2. Note that since the line used to the approximate the

manifold is not an exact solution of the manifold, it is only used to compute the

minimum eccentricity (distance to the line) and not the argument of periapsis (angle

between the origin and the line). Thus, the (e,ω) pair actually sits on the manifold.

If the approximation was used to compute both values, the resulting pair would not

lie on the manifold, but on the approximation to the manifold.

5.3.2 Examples of Long Lifetime Orbits

The general algorithm used to compute long lifetime orbits is given in Table 5.1.

As previously noted, it is not always necessary to find the maximum lifetime orbit

corresponding to a particular frozen orbit. This maximum lifetime orbit will be the

trajectory that follows the desired manifold path exactly (i.e. it has the manifold

as its average). The goal here is to show that it is possible to design long lifetime

trajectories that follow in general the II→I path as in Figure 5.4. They do not need

to have the manifold has their exact average, but must lie relatively close to the

manifold. The first examples presented are for the Europa system, with the tide, J2,

C22 and J3 components included in the model. Following that, we also show that it

113

Page 127: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 5.1: Algorithm for computing a long lifetime orbit

1. Choose frozen orbit in the 1-DOF system based on its inclination.2. Compute the minimum eccentricity point on the stable manifold of the

frozen orbit.3. Compute the corrected initial conditions to the 2-DOF system by using the

linearization of the 2-DOF system about the point on the manifold.4. Compute the corrected initial conditions to the 3-DOF system by using the

linearization of the 3-DOF system about the corrected 2-DOF initial conditions.5. Find an argument of periapsis value that produces a long lifetime orbit by

numerical search, if necessary.

is possible to design orbits for other planetary satellites such as the Saturnian moons

Enceladus and Dione. Since, as demonstrated in Chapter 3, the environments in the

vicinity of those moons are much more unstable than the Europa environment, the

long lifetime orbits designed for those systems are not nearly as long-lived as those

in the Europa system.

Starting with the Europa system, trajectories are computed using the minimum

eccentricity point on the manifold in step 2 of the algorithm in Table 5.1. We find

that the trajectory follows the desired manifold path without a bias correction in all

the cases, which cover the entire range of unstable, near-polar, near-circular frozen

orbits. The lifetimes of these trajectories are:

Frozen Orbit Lifetime

Inclination (days)

70o 105

75o 119

85o 110

95o 96

105o 95

110o 99

Note that the lifetimes of the direct orbits are all larger than the lifetimes of

114

Page 128: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

the retrograde orbits. It has been shown that direct and retrograde orbits have

different properties in the 3-DOF system and that this difference disappears during

the averaging technique used here [22]. Therefore, it is possible that, due to the

averaging used for this analysis, the initial condition computation algorithm provides

better results for direct than for retrograde orbits. These trajectories fulfill the goal

of designing a long lifetime orbit that follows the desired manifold path, without

having it as an exact average. It is an interesting exercise to demonstrate that the

algorithm brings us very close to a trajectory that follows the manifold exactly (i.e.

the maximum lifetime orbit). To find the maximum lifetime orbit for each case, a

numerical search over argument of periapsis values is performed to find the necessary

bias (step 5 of the algorithm in Table 5.1). Also, note that although all of the

trajectories initialized at the minimum eccentricity point follow the correct manifold

path in the Europa case, as shown above, that will not necessarily be true for all

initial points on the manifold. For those cases, a numerical search over ω values can

be performed to either find the maximum lifetime orbit or any long lifetime orbit

that follows the correct manifold path.

Table 5.2 shows the bias in the argument of periapsis that produces the longest

lifetime orbit to an accuracy of 0.0002 radians (0.011o) for each frozen orbit

inclination case discussed above. Note that all of the bias values are less than about

one and a half degrees and the lifetimes increase by between 45 and 90 percent. This

large variation in improvement in lifetime occurs since the maximum lifetime orbit is

only being computed to an accuracy of 0.0002 radians in the argument of periapsis

bias. If a smaller step size in argument of periapsis was used, the lifetimes would all

increase by percentages in the higher range. Also note that the bias values are larger

for the retrograde orbits than for the direct orbits. This is probably because the

original lifetimes of the retrograde orbits, using the unbiased value of ω, are smaller

than the original lifetimes for the direct orbits. Although the existence of these

115

Page 129: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 5.2: Long Lifetime Orbits for the Europa System

Frozen Orbit Unbiased Lifetime Maximum Lifetime Bias (rad)Inclination (days) (days)

70o 105 182 0.009575o 119 174 0.002685o 110 171 0.004695o 96 182 0.0169105o 95 172 0.0254110o 99 188 0.0242

maximum lifetime orbits is interesting in a theoretical sense, from a mission planning

point of view, actually obtaining the maximum lifetime orbit is not very practical.

This is because, as we will show in the next chapter, the accuracy with which the

position and velocity of an orbit can be obtained is larger than the accuracy needed

for the maximum lifetime orbit. Figure 5.12 shows the unbiased long lifetime orbit

and the maximum long lifetime orbit for the 75o inclination case. Observe that the

maximum lifetime orbit has the manifold as its average, while the unbiased long

lifetime orbit sits slightly above the manifold.

Now, proceeding with the Saturnian moons, first note that the models for these

systems only include the tidal force since gravity field information is not currently

available. Therefore, the frozen orbits under consideration are the near-polar,

circular unstable frozen orbits. The goal here is simply to demonstrate that although

Europa has been the primary focus of this dissertation, the method for designing

long lifetime orbits described here is applicable to other planetary satellite systems.

For a real mission, the gravity fields of these planetary satellites would likely be

measured during the mission, prior to the science phase, and our design technique

including gravity field coefficients could be used. This would improve the lifetimes

of the orbits computed due to the advantage of having a non-circular frozen orbit.

The minimum eccentricity point on the manifold is actually the frozen orbit

in this case, since the frozen orbit is circular. Therefore, an arbitrary starting

116

Page 130: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08

−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

impact

(a) General long lifetime orbit

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08

−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

impact

(b) Maximum lifetime orbit

Figure 5.12: Long lifetime orbits about Europa (70o frozen orbit inclination)

117

Page 131: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 5.3: Long Lifetime Orbits about Enceladus

Frozen Orbit Lifetime BiasInclination (days) (rad)

70o 10.09 085o 9.79 0.0315105o 11.18 0.0715110o 11.18 0.0890

Table 5.4: Long Lifetime Orbits about Dione

Frozen Orbit Lifetime BiasInclination (days) (rad)

70o 53.79 0.012685o 62.73 0.00725105o 53.46 0.02688110o 62.13 0.03418

eccentricity on the manifold is chosen in each case. The long lifetime orbits computed

are trajectories that follow the desired manifold path, without necessarily being the

maximum lifetime orbit. Bias corrections are included where necessary to produce

a trajectory that follows the desired path. Since Enceladus has a radius of only 252

km, it was necessary to set the frozen orbit altitude to 50 km instead of 100 km as

in the Europa case. The perturbations on the orbits about Enceladus with 100 km

altitudes were too large to make meaningful design possible. Dione has a radius of

560 km, and so frozen orbit altitudes of 100 km were feasible for the design.

Tables 5.3 and 5.4 show the results of the the long lifetime trajectory design for

Enceladus and Dione. Figures 5.13 and 5.14 show the trajectories in (h,k) space for

both planetary satellites for the 70o frozen orbit inclination case. First, observe the

the lifetimes of the orbits for both planetary satellites are significantly less than the

lifetimes obtained for the Europa system. The orbits about Enceladus have lifetimes

in the vicinity of 10 days, and the orbits about Dione have lifetimes in the vicinity

of 50-60 days. This is not surprising since the environments in the vicinity of both

118

Page 132: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2

−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

0.2

k=ecosω

h=es

inω

impact

Figure 5.13: Long lifetime orbit about Enceladus (70o frozen orbit inclination)

planetary satellites are more highly perturbed than the environment about Europa.

These shorter lifetimes follow from the characteristic time results in Chapter 3,

Figure 3.9.

The main result of this section is that the long lifetime orbit design technique is

a feasible method for designing long duration orbits about planetary satellites. The

maximum possible length of the designed orbit depends on the planetary satellite

system under consideration, where the more highly perturbed a system is, the

shorter the lifetimes will be. In addition, it is possible to do an initial design using

a tide-only model to get an estimate of the lifetimes of trajectories possible for a

particular system. These orbits can then be refined in a real mission situation once

the gravity field has been measured. For Europa in particular, which is the main

motivation of this dissertation, it is possible to design orbits with lifetimes of at least

100 days using this method.

119

Page 133: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.15 −0.1 −0.05 0 0.05 0.1 0.15

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

k=ecosω

h=es

inω

impact

Figure 5.14: Long lifetime orbit about Dione (70o frozen orbit inclination)

5.4 A Toolbox for Computing Long Lifetime

Orbits

The main goal of the toolbox is to provide a set of functions, with inputs modifiable

by the user, to perform analysis of the frozen orbit solutions and compute initial

conditions of the long lifetime science orbits. It can be used for science orbit design

for any planetary satellite that can be modeled by the Hill 3-body problem. This

toolbox was written for the Jupiter Icy Moons Orbiter Mission and delivered to the

Jet Propulsion Laboratory.

The toolbox, written in Matlab, integrates all of the results in this chapter and

provides three main functions:

1. Plot the frozen orbit solutions in the 1-DOF system. This plot gives an overall

picture of the underlying dynamics of the full 3-DOF system.

2. Compute initial conditions in the 3-DOF system that yield a trajectory that

follows the manifolds in the 1-DOF system on average.

3. Integrate a set of initial conditions in the 3-DOF system and produce a series

of plots.

120

Page 134: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

This toolbox not only provides initial conditions for long lifetime science orbits, but

also allows for analysis. Since one of the input groups is the physical parameters of

the system, it can be used to compare frozen orbit solutions for different planetary

satellites, or to understand how the frozen orbit solutions depend on the physical

parameters (i.e., the gravity field terms) for a particular planetary satellite. The

user can also choose the approximate initial eccentricity of the science orbit through

the choice of the eccentricity on the 1-DOF manifold, and the desired frozen orbit

inclination. The toolbox produces initial conditions in the 3-DOF system that have

not been adjusted with a bias in the argument of periapsis. Therefore, the user

can also study how changing the initial argument of periapsis affects the science

orbit duration. This is especially important when considering the effect of orbit

uncertainty. This analysis is done with the third part of the toolbox, which also has

the capability to integrate, in the 3-DOF system, of any set of initial conditions.

121

Page 135: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

CHAPTER 6

CONTROL OF LONG LIFETIME ORBITS

Thus far, we have simplified the original 3-DOF system to analyze the underlying

motion of the system and used this information to design long lifetime orbits in the

full 3-DOF system. However, as shown in the previous chapter, long lifetime orbits

have finite lifetimes. It is not only important to discuss the control of the orbits

in the 3-DOF system used here, but also the control of orbits in a real-life system

where it is impossible to attain the desired orbit exactly due to uncertainty errors in

the position and velocity. The specific control issues that will be discussed in this

chapter are:

• Low thrust control to keep the spacecraft at a specific location

• Impulsive maneuvers to reset the long lifetime orbit once it reaches the unstable

manifold

• Impulsive maneuvers to relocate the spacecraft to the desired position on the

manifold after an initial error

The goal of our low-thrust control technique is to continually thrust to keep

the spacecraft at a location that satisfies the requirements of a science orbit. In

theory, this seems reasonable and has the advantage of keeping the orbital elements

constant. This allows for an indefinite lifetime as long as the thrust does not stop.

Thrust laws are developed for purely radial and and purely transverse thrust. The

results show that attempting to control the spacecraft in this way requires very

122

Page 136: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

large thrust values that are not practical. Although the transverse thrust technique

requires an order of magnitude thrust levels less than the radial thrust technique,

neither of the methods are feasible in the practical sense. Recalling the manifolds of

the unstable frozen orbits, it is not surprising that attempting to keep a spacecraft

at one of those locations is very costly. As we saw, an uncontrolled spacecraft placed

at a frozen orbit location will diverge from the frozen orbit quite rapidly along

the unstable manifold. By attempting to keep the spacecraft at the frozen orbit

location, we do not take advantage of the dynamics of the system. It is therefore

more practical to allow the spacecraft to follow the stable and unstable manifolds

and make corrections when the spacecraft gets to a certain point on the unstable

manifold past the frozen orbit.

Therefore, the next technique considered is to allow the orbit to evolve from

its initial point at the minimum eccentricity location on the stable manifold to the

minimum eccentricity point on the unstable manifold. Then, two maneuvers are

performed to reset the orbit back to the stable manifold. Instead of requiring that the

spacecraft return to the exact same initial point, some of its current characteristics

such as the radius of periapsis and inclination are used to design a new long lifetime

orbit. Finally, the two maneuvers are optimized for the lowest cost transfer to the

new target. The ideal transfer is a circularization maneuver followed by a maneuver

to increase the eccentricity at a location on the transfer orbit that would set the

argument of periapsis appropriately. That sequence is used as a starting guess for the

optimization routine that optimizes over the total cost of both transfer maneuvers.

We find that the cost to do both maneuvers is fairly low, and is practical for a real

mission.

The final control issue discussed in this chapter is correcting the spacecraft’s

position on the designed long lifetime trajectory if it drifts away due to initial errors.

Monte Carlo analysis of trajectories with initial errors show that errors within the

123

Page 137: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

range of uncertainty for a mission to Europa can cause the trajectory to follow the

wrong manifold path. The uncertainty limits for a mission to a planetary satellite

such as Europa were taken from [40]. Since initial errors in the spacecraft’s position

can destroy the long-lifetime properties of the trajectory, it is important to determine

ways to get the spacecraft back on the correct orbit with a relatively low cost. We

show that if the erroneous trajectory is integrated along with the correct trajectory

for several days and then a sequence of maneuvers is designed to place the spacecraft

at the correct location corresponding to the time of the correction, the cost of the

correction is very low. The transfer is a two-maneuver sequence, optimized for the

total cost. The ideal transfer is a Hohmann-type transfer between two elliptic orbits

and this is used as the initial guess of the optimization. This technique results in

very low costs and is a realistic technique for a real mission.

6.1 Low-Thrust Control

Since the unstable near-polar, near-circular unstable frozen orbits have characteristics

that make them well-suited for a science orbit about a planetary satellite, it is logical

to attempt to use low-thrust control to keep the spacecraft at a frozen orbit location.

In terms of a real mission, it is easier to consider thrust along only one axis. The

radial thrust law is developed by analyzing the averaged Gauss equations and the

doubly averaged Lagrange Planetary Equations in tandem. Although the thrust law

is developed to control the orbit in the 1-DOF system, it also controls the orbit in

the 3-DOF system by preventing impact with the planetary satellite for very long

periods of time. Thrusting in the radial direction was the preferred orientation for

the proposed Jupiter Icy Moons Orbiter spacecraft as it combined gravity gradient

stability of its attitude with thrusting along its long axis. Unfortunately, the thrust

required to control the spacecraft using this thrust law is much too large to be

practical for a real mission.

124

Page 138: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

The transverse thrust law is developed by determining the thrust required to

control the eccentricity of the orbit in the transverse direction. For this case, a

circular orbit rather than a near-circular frozen orbit is used since circular orbits are

also very useful as science orbits for mission to planetary satellites. The thrust law

developed controls the eccentricity of the orbit while maintaining the semi-major

axis of the orbit. The thrust required, while being an order of magnitude smaller

than that required for the radial thrust law, is still too large for a mission. We

demonstrate these thrust laws here to motivate the next section in which control

schemes with more reasonable thrust requirements are developed for the long lifetime

orbits developed in Chapter 5 that take advantage of the natural dynamics of the

system.

6.1.1 Thrust in the Radial Direction

The 1-DOF system is the simplest system under consideration and gives a good

description of the underlying dynamics. The thrust law in the radial direction is

therefore developed using this system and then tested in the full 3-DOF system.

The Gauss Equations describe the motion of a spacecraft subject to perturbing

accelerations in terms of the orbital elements. The important orbital elements in the

1-DOF are the semi-major axis, eccentricity, inclination and argument of periapsis.

Although the semi-major axis is constant in the 1-DOF system, it does affect the

other orbital elements and so its motion must be included in the analysis when

125

Page 139: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

thrusting is added. The Gauss equations for these orbital elements are [11]:

daG

dt=

2e sin ν

n√

1− e2Fr +

2a√

1− e2

nrFs , (6.1)

deG

dt=

√1− e2 sin ν

naFr +

√1− e2

na2eFs , (6.2)

diGdt

=r cos (ω + ν)

na2√

1− e2Fω , (6.3)

dωG

dt= −

√1− e2 cos ν

naeFr +

√1− e2

nae

[sin ν

(1 +

1

1 + e cos ν

)]Fs

− r cot i sin (ω + ν)

na2√

1− e2Fω . (6.4)

The three thrust components are Fr, in the radial direction; Fs, 90o from r in the

velocity-increasing direction; and Fω, normal to the orbit plane.

In the 1-DOF system, an increasing argument of periapsis value causes the

eccentricity to increase. Therefore, for the spacecraft to remain at the frozen orbit

location on average, the argument of periapsis must be constant on average. This

can be accomplished with the proper choice of a thrust law. To ensure that the

thrust does not cause changes in the other orbital elements, it must be chosen such

that it has no contribution to the motion of the semi-major axis, eccentricity and

inclination. Since it has already been shown that averaging over the mean anomaly

of the orbit of the spacecraft is a reasonable approximation, the Gauss equations are

also averaged in that way, where constant thrust is assumed. Let

dxG

dt= f(a, e, i, ω, ν) (6.5)

be one of Eqs.(6.1)-(6.4). Then, the average of that orbital element over the mean

anomaly is

dxG

dt=

1

∫ 2π

0

f(a, e, i, ω, ν)dM . (6.6)

Since Eq.(6.1)-(6.4) are functions of the true anomaly, the following transformation

126

Page 140: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

is used to integrate over ν:

dM =(1− e2)3/2

(1 + e cos ν)2dν . (6.7)

Then, the averaged Gauss equations are:

daG

dt=

2√

1− e2

nFs , (6.8)

deG

dt=

3e(1− e2)

2naFs , (6.9)

diGdt

= −e√

1− e2 cos ω

naFω , (6.10)

dωG

dt=

1

naFr +

4− e2

naeFs +

3ae cot i sin ω

2√

1− e2Fω . (6.11)

The requirement that ˙ag = ˙eG = ˙iG = ˙ωG = 0 entails letting Fs = Fω = 0. To

ensure that the argument of periapsis is constant on average, use Eq.(3.25) which

describes the motion of the argument of periapsis in the 1-DOF system to determine

the thrust component Fr. The motion of the argument of periapsis in the controlled

1-DOF system is given by the addition of Eq.(3.25) and Eq.(6.11):

dt=

3N2

8n

1√1− e2

[5 cos2 i− 1 + 5 sin2 i cos 2ω + e2(1− 5 cos 2ω)]

+3nJ2

4a2(1− e2)2

(1− 5

4sin2 i

)+

3J3n

2a3(1− e2)3

sin ω sin i

e[(1

−5

4sin2 i

)(1 + 4e2

)− e2

sin2 i

(1− 19

4sin2 i +

15

4sin4 i

)]+

1

naFr . (6.12)

Then, the radial thrust law necessary for ω = 0 is:

Fr = −nadω

dt,

where dω/dt is given in Eq.(6.12).

The first step in testing this thrust law is make sure that it stabilizes the

spacecraft in the 1-DOF system. The doubly averaged Lagrange Equations plus the

unaveraged Gauss equations, with Fr as defined in Eq.(6.13) and Fs = Fω = 0 are

127

Page 141: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

integrated:

da

dt=

2e sin ν

n√

1− e2Fr , (6.13)

de

dt=

15N2

8ne√

1− e2 sin2 i sin 2ω − 3J3n

2a3(1− e2)3sin i

(1− 5

4sin2 i

)cos ω

+

√1− e2 sin ν

naFr , (6.14)

di

dt= −15N2

16n

e2

√1− e2

sin 2i sin 2ω +3J3n

2a3(1− e2)3e cos i

(1− 5

4sin2 i

)cos ω , (6.15)

dt=

3N2

8n

1√1− e2

[5 cos2 i− 1 + 5 sin2 i cos 2ω + e2(1− 5 cos 2ω)]

+3nJ2

4a2(1− e2)2

(1− 5

4sin2 i

)+

3J3n

2a3(1− e2)3

sin ω sin i

e

[(1− 5

4sin2 i

)·(1 + 4e2

)− e2

sin2 i

(1− 19

4sin2 i +

15

4sin4 i

)]−√

1− e2 cos ν

naeFr , (6.16)

dt= −3N2

8n

cos i√1− e2

(2 + 3e2 − 5e2 cos 2ω)− 3nJ2

2a2(1− e2)2cos i

+3nJ3

2a3(1− e2)3

e sin ω cos i

sin i

(1− 15

4sin2 i

). (6.17)

Figure 6.1 shows the time histories of the controlled orbital elements over a 6 hour

period. Observe that the argument of periapsis oscillates and has a slow drift.

Compare this to Figure 6.2 which is the integration of uncontrolled doubly averaged

Lagrange Equations. The radial thrust slows the drift of the argument of periapsis

which in turn slows the drift of the eccentricity which will delay the impact of the

spacecraft with the planetary satellite.

Now that we have verified that the radial thrust law controls the spacecraft in the

1-DOF system, it is tested in the 3-DOF system. Figure 6.3 shows the time histories

of the orbital elements integrated in the 3-DOF system with the radial thrust law

Fr as defined in Eq.(6.13), over a 250 day period. It is clear that this thrust law

adequately controls the spacecraft since it has not impacted with the planetary

satellite after 250 days. Figure 6.4 shows the manifold in the 1-DOF system and the

integrated trajectory. Observe that the trajectory remains close to the frozen orbit.

128

Page 142: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 1 2 3 4 5 61694.69

1694.7

1694.71

1694.72

a (k

m)

0 1 2 3 4 5 60.02

0.0202

0.0204

ecce

ntric

ity

0 1 2 3 4 5 669.999

69.9995

70

70.0005

70.001

i (de

g)

time (hours)

0 1 2 3 4 5 6−91

−90.5

−90

−89.5

ω (

deg)

0 1 2 3 4 5 6−0.8

−0.6

−0.4

−0.2

0

Ω (

deg)

time (hours)

Figure 6.1: Time histories of orbital elements integrated in the controlled 1-DOF system.

0 1 2 3 4 5 61693

1694

1695

1696

a (k

m)

0 1 2 3 4 5 60.0199

0.02

0.0201

ecce

ntric

ity

0 1 2 3 4 5 669.999

69.9995

70

70.0005

70.001

i (de

g)

time (hours)

0 1 2 3 4 5 6−91

−90.5

−90

ω (

deg)

0 1 2 3 4 5 6−0.8

−0.6

−0.4

−0.2

0

Ω (

deg)

time (hours)

Figure 6.2: Time histories of orbital elements integrated in the uncontrolled 1-DOF system.

0 50 100 150 200 2501680

1685

1690

1695

1700

a (k

m)

0 50 100 150 200 2500

0.005

0.01

0.015

0.02

ecce

ntric

ity

0 50 100 150 200 25068

69

70

71

i (de

g)

time (days)

0 50 100 150 200 2500

100

200

300

400

ω (

deg)

0 50 100 150 200 250−200

−100

0

100

200

Ω (

deg)

time (days)

Figure 6.3: Integration of the 3-DOF system controlled with Fr.

129

Page 143: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

Figure 6.4: Manifold in the 1-DOF system and integration of the 3-DOF system controlled with Fr.

Although the radial thrust law, Fr adequately controls the spacecraft it requires

instantaneous values of the orbital elements which results in constant computation.

The total thrust magnitude and the x, y and z components of the thrust are shown

in Figure 6.5. Observe that the thrust is of the order 10−6km/s2. The total ∆v

can also be computed by integrating the thrust magnitude (see Figure 6.6). The

total ∆v required over a 150 day period is 12 km/s which is an average of 80 meters

per second per day. Then, instead of using the thrust law Fr in Eq.(6.13), we try

controlling the spacecraft using a constant thrust in the radial direction equal to the

average of 80 m/s/day which is −9.26× 10−7km/s2. The results of this integration

are shown in Figure 6.7. We see that the thrust used is not enough to prevent the

spacecraft from impacting. It impacts with Europa after about 50 days. It is possible

however, to control the spacecraft using a slightly larger constant radial thrust. We

find that the minimum constant thrust required is −1.045× 10−6km/s2 which is 90.2

meters per second per day. Figure 6.8 shows the results of this integration where

the spacecraft has not impacted with Europa after 150 days. The problem with the

radial thrust technique is that although in theory it is possible to keep the spacecraft

at the frozen orbit location by either constant or time-varying radial thrust, the

thrust required is much too large to be practical for a mission.

130

Page 144: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 150 200 2500

0.5

1

1.5

2

2.5

3

3.5

4

4.5x 10

−6

time (days)

mag

nitu

de o

f thr

ust (

km/s

2 )

50 100 150 200 250

−2

−1

0

1

2

x 10−6

Fx (

km/s

2 )

50 100 150 200 250

−2

0

2x 10

−6

Fy (

km/s

2 )

50 100 150 200−4

−2

0

2

4x 10

−6

Fz (

km/s

2 )

time (days)

Figure 6.5: Time histories of total thrust magnitude and components of thrust.

0 50 100 1500

2

4

6

8

10

12

14

∆ v

(km

/s)

time (days)

Figure 6.6: Total ∆v used to control the spacecraft using the thrust law given in Eq.(6.13).

131

Page 145: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 5 10 15 20 25 30 35 40 45 501680

1685

1690

a (k

m)

0 5 10 15 20 25 30 35 40 45 500

0.02

0.04

0.06

0.08

e

0 5 10 15 20 25 30 35 40 45 5068

69

70

71

i (de

g)

time (days)

0 5 10 15 20 25 30 35 40 45 50150

200

250

300

ω (

deg)

0 5 10 15 20 25 30 35 40 45 50−100

−50

0

50

Ω (

deg)

time (days)

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

Figure 6.7: Integration using constant thrust in the radial direction of −9.26× 10−7km/s2. Thespacecraft impacts with Europa after about 50 days.

132

Page 146: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 50 100 1501680

1685

1690

a (k

m)

0 50 100 1500

0.02

0.04

0.06

0.08

e

0 50 100 15068

69

70

71

i (de

g)

time (days)

0 50 100 150150

200

250

300

ω (

deg)

0 50 100 150−200

−100

0

100

200

Ω (

deg)

time (days)

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

Figure 6.8: Integration using constant thrust in the radial direction of −1.045× 10−6km/s2. Thespacecraft has not impacted with Europa after 150 days.

133

Page 147: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.090

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5x 10

−3

time (days)

ecce

ntric

ity

Figure 6.9: Eccentricity as a function of time over one orbital period.

0 50 100 1501680

1685

1690

1695

1700

a (k

m)

0 50 100 1500

2

4

6x 10

−3

e

0 50 100 15068

69

70

71

i (de

g)

time (days)

0 50 100 1500

100

200

300

400

ω (

deg)

0 50 100 150−200

−100

0

100

200

Ω (

deg)

time (days)

Figure 6.10: Time histories of the orbital elements when the transverse thrust is applied.

6.1.2 Thrust in the Transverse Direction

Since the radial thrust technique resulted in very large thrust levels, we try

developing a transverse thrust law to control the spacecraft. For this case, the goal

is to control the spacecraft such that it stays in a circular, highly inclined orbit. The

control is therefore in place to prevent the eccentricity of the orbit from increasing.

The first step is to initialize a trajectory in a circular orbit with an inclination of 70o

and integrate it for one orbital period. Figure 6.9 shows the result of this integration.

The net change in the eccentricity is very small at 2.32× 10−4.

134

Page 148: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

The next step is the computation of the ∆v required to circularize the orbit after

one orbital period. In this example, the ∆v required is 0.8395 m/s. Therefore, to

prevent the eccentricity of the orbit from increasing a first guess at the required

transverse thrust is the ∆v divided by the orbital period. The orbital period

is 7.454 × 103 (2.07 hours) and the transverse thrust is 1.126 × 10−7 km/s2

(9.73 m/s/day). In order to prevent the semi-major axis from increasing, the thrust

is applied in the velocity direction for half of the orbit and against the velocity

direction for the other half, symmetrically about periapsis. Figure 6.10 show the

time histories of the orbital elements for the integration with the transverse thrust

law used to control the orbit. During the 150 day period of the integration, the

eccentricity and semi-major axis stay relatively constant and the spacecraft does not

impact with Europa.

Although the transverse thrust required to control the spacecraft is about an

order of magnitude less than the radial thrust required, it is still much too large in

the practical sense. These exercises in developing purely radial and purely transverse

low-thrust laws to keep some of the orbital elements of the spacecraft constant show

that due to the strong perturbations present in the system, this is not an efficient

way to control a spacecraft. The following sections detail control methods that take

advantage of the natural dynamics of the system to reduce the thrust required to

prevent the spacecraft from impacting with the planetary satellite.

6.2 Resetting a Long Lifetime Orbit

Recall the long lifetime orbits developed in Chapter 5. They were designed by

computing initial conditions such that the spacecraft follows the stable and unstable

manifolds of the 1-DOF system. Although these orbits have long lifetimes, they

are finite and the spacecraft will eventually follow the unstable manifold to impact

with the planetary satellite. It is important to develop a control strategy to delay

135

Page 149: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

impact with the planetary satellite. Since an algorithm to design long lifetime orbits

already exists, it is natural to use this algorithm during the development of the

control scheme. The technique used to control the long lifetime orbit is to allow the

trajectory to reach the unstable manifold and then, at a certain point, reset it back

to the stable manifold so another long lifetime orbit can evolve.

6.2.1 Designing the Target

The initial conditions of a long lifetime orbit are computed by starting at the

minimum eccentricity point on the manifold and are a specific 6-state. The minimum

eccentricity point on the manifold corresponds to a particular frozen orbit which in

turn corresponds to a specific radius of periapsis. As the orbit evolves, the orbital

elements oscillate, and in particular, the radius of periapsis oscillates. Therefore,

when the orbit is stopped on the unstable manifold, resetting it back to the original

long lifetime orbit initial conditions is very difficult since the modification of all of

the parameters of the orbit would be necessary. Since the goal is simply to prevent

the impact of the spacecraft with the planetary satellite by resetting the trajectory

to the stable manifold, it is not necessary to reset it to the exact same long lifetime

orbit. Therefore, at the point that the orbit is stopped on the unstable manifold a

new long lifetime orbit is computed, with parameters consistent with the current

orbital parameters.

Denote the initial conditions of the new long lifetime orbit to be designed the

target. The first step in designing the target is to specify its corresponding frozen

orbit. Since the algorithm used to design long lifetime orbits places the spacecraft

at the minimum eccentricity point on the stable manifold, it is logical to stop the

orbit at the minimum eccentricity point on the unstable manifold. This minimizes

the eccentricity changes that will be required during the transfer. For simplicity in

the design, the stopping point used is actually the first periapsis passage after the

136

Page 150: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

180 200 220 240 260 280 300 320 340

0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

ω (degrees)

ecce

ntric

ity

target

∆ v1 ∆ v

2

frozen orbit

stopping point

transfer orbit

Figure 6.11: Diagram of the reset scheme in the 1-DOF system in (e,ω)-space.

minimum eccentricity point on the unstable manifold. In the ideal 1-DOF system,

there are no oscillations while the trajectory moves from its initial point on the

stable manifold to the stopping point on the unstable manifold. Thus, to reset the

orbit to the stable manifold, only the argument of periapsis needs to be changed.

The easiest way to do this is a two maneuver sequence consisting of a circularization

at the stopping point and an a maneuver to increase the eccentricity at ω + ν = ωll,

where ωll is the initial argument of periapsis of the original long lifetime orbit.

Since the ideal transfer orbit is circular, this resets the orbit to its original point

on the stable manifold. See Figure 6.11 for a diagram of this reset scheme in the

1-DOF system. Obviously this system is not the ideal 1-DOF system and there are

perturbations present. However, the ideal transfer can be used to generate a target

in the 3-DOF system.

The first step is to circularize the orbit at the stopping point and let the orbit

evolve in the 3-DOF system until ω + ν = ωll. Denote this point with the subscript

‘IT’ (ideal target). Then, the frozen orbit to be used in the actual target design has

a radius of periapsis equal to rIT and an inclination equal to iIT . The frozen orbit

eccentricity can then be computed along with the minimum eccentricity on the new

137

Page 151: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

manifold. Using all of these values, the initial conditions of a new long lifetime orbit

can be computed using the algorithm in Chapter 5. Note that ΩIT = ΩIT − NtIT .

Since the target is always designed for time equal to zero, Ω0 = ΩIT . The orbital

elements of the target are denoted with the subscript ‘T’. A summary of the

procedure for computing the target is as follows:

1. Stop the original long lifetime trajectory at the first periapsis after the

minimum eccentricity point on the unstable manifold.

2. Perform the ‘ideal transfer’ which consists of a circularization at the stopping

point and integrate until ω + ν = ωoriginal.

3. Let the frozen orbit corresponding to the target have a radius of periapsis and

inclination of rIT and iIT respectively and Ω0 = ΩIT .

4. Compute the target position using the algorithm for computing long lifetime

orbits in Chapter 5.

6.2.2 Transferring to the Target

The next step is determining how to transfer to the target, knowing that there are

perturbations present which prevent the use of the ideal transfer described above.

Note however, that the ideal transfer is a good first guess. Therefore, we start with

the circularization burn, and integrate until ω + ν = ωT . Denote the transfer time

tf . Aside from reaching the target, our goal is also to minimize the total cost of the

transfer, which consists of two burns – one at periapsis on the unstable manifold and

one at the target. First, we focus on reaching the target. Let rT denote the target

position and vT denote the target velocity. Then,

rT = r(tf , r0,v0 + δv) , (6.18)

vT = v(tf , r0,v0 + δv) , (6.19)

138

Page 152: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

where r0 and v0 correspond to the stopping point (at periapsis) on the unstable

manifold and δv is the first burn. A Taylor expansion of Eq.(6.18) yields:

rT = r(tf ) +∂r

∂v0

∣∣∣∣tf

δv + · · · (6.20)

Then, keeping only first order terms and rearranging for δv:

δv =

(∂r

∂v0

)−1∣∣∣∣∣tf

(rT − r0) . (6.21)

Denote the state transition matrix as

Φ(0, tf ) =

φrr φrv

φvr φvv

. (6.22)

Then, ∂r∂v0

∣∣∣tf

= φrv. Also, let δr = (rT − r(tf )). We can converge on the δv necessary

for the transfer orbit to end at the target rT using the following iteration until

δr = 0:

δv = δv0 + δv1 + δv2 + · · · , (6.23)

where

δvi+1 = φ−1rv δri . (6.24)

The above method ensures that the spacecraft reaches the target position by the

end of the transfer orbit. At the target position a second burn is executed such that

the spacecraft velocity is equal to the target velocity. The first burn is δv and the

second burn is vT − v(tf ). However, this method does not optimize the total cost.

Therefore, the next step is refining the transfer orbit and hence the maneuvers to

reduce the cost of both burns. Let ∆v∗1 = δv and ∆v∗2 = vT − v(tf ). Then, consider

a correction to the transfer time tf , denoted as δt, needed to optimize the total cost.

This necessitates the introduction of corrections to both burns δv1 and δv2. The

139

Page 153: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

target position and velocity are now:

rT = r(tf + δt, r0,v0 + ∆v∗1 + δv1) , (6.25)

vT = v(tf + δt, r0,v0 + ∆v∗1 + δv1) + ∆v∗2 + δv2 . (6.26)

A Taylor series expansion of Eqs.(6.25) and (6.26) yields

rT = r(tf ) +∂r

∂t

∣∣∣∣tf

δt +∂r

∂v0

∣∣∣∣tf

δv1 + · · · (6.27)

vT = v(tf ) +∂v

∂t

∣∣∣∣tf

δt +∂v

∂v0

∣∣∣∣tf

δv1 + · · ·+ ∆v∗2 + δv2 . (6.28)

Then, truncating at first order in the Taylor expansions and noting that rT = r(tf )

and vT = v(tf ) + ∆v∗2, we obtain the following expressions for δv1 and δv2:

δv1 = −(

∂r

∂v0

)−1∣∣∣∣∣tf

v(tf )δt , (6.29)

δv2 = −

(∂v

∂t

∣∣∣∣tf

+∂v

∂v0

∣∣∣∣tf

δv1

)δt . (6.30)

Then, using the definition of the state transition in Eq.(6.22), φrv = ∂r∂v0

∣∣∣tf

and

φvv = ∂v∂v0

∣∣∣tf

. The expressions for δv1 and δv2 are then:

δv1 = −φ−1rv v(tf )δt , (6.31)

δv2 = −[a(tf )− φvvφ

−1rv v(tf )

]δt , (6.32)

where a(tf ) is the acceleration at tf .

The goal of this scheme is to minimize the total cost. Let the cost function be

J = (∆v∗1 + δv1)2 + (∆v∗2 + δv2)

2 . (6.33)

Since δv1 and δv2 are functions of δt, denote them as δv1 = α1δt and δv2 = α2δt.

140

Page 154: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Then, J can be expressed as

J = |∆v∗1|2 + |∆v∗2|2 + 2[(∆v∗1)T α1 + (∆v∗2)

T α2]δt + (|α1|2 + |α2|2)δt2 (6.34)

= Aδt2 + 2Bδt + C . (6.35)

Since the cost J is a function of δt, take the derivative of J with respect to δt

and set it equal to zero to minimize the cost:

∂J

∂t= 2Aδt + 2B = 0 . (6.36)

Therefore, the δt required to minimize the total cost is:

δt = −B

A, (6.37)

where

A = v(tf )T φ−T

rv φ−1rv v(tf ) + [a(tf )− φvvφ

−1rv v(tf )]

T [a(tf )− φvvφ−1rv v(tf )] , (6.38)

B = −(∆v∗1)−1φ−1

rv v(tf )− (∆v∗2)T [a(tf )− φvvφ

−1rv v(tf )] . (6.39)

The overall procedure to compute the maneuvers necessary to transfer the

spacecraft from a periapsis location on the unstable manifold to a new long lifetime

orbit initialized on its stable manifold is as follows:

1. Compute the ideal transfer consisting of a circularization at periapsis on the

unstable manifold and an integration until ω + ν = ωT . Let the time required

for this transfer be tf .

2. Use the iteration scheme

δv = δv0 + δv1 + δv2 + · · · (6.40)

where

δvi+1 = φ−1rv δri (6.41)

141

Page 155: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

to compute the necessary correction to the first burn such that δr = 0.

3. Let ∆v∗1 = δv and ∆v∗2 = vT − v(tf ) and compute new corrections to the

burns, δv1 and δv2, to minimize the cost using δt = −BA.

4. Repeat steps 2 and 3 until δti+1 − δti < 10−9.

The optimization scheme described above doesn’t actually minimize the total

cost, but minimizes the the sum of the squares of the two burns. To minimize the

total cost would require J = |∆v∗1 + δv1| + |∆v∗2 + δv2|. However, testing of both

methods results in approximately the same total cost for both cases and so the sum

of the squares method is used since it is computationally simpler.

6.2.3 Examples

The examples detailed in this section show that the technique described above

provides a practical way to reset a long lifetime orbit with a relatively low cost. As

previously mentioned, it is not necessary to find the maximum lifetime orbit on a

particular manifold since it would not be achievable in a real mission. However it is

important to ensure that the long lifetime orbit follows the correct manifold. The

long lifetime orbits used in these examples are those in Table 5.2 (the unbiased long

lifetime orbits, not the maxima). These unbiased long lifetime orbits all follow the

correct manifold path. However, when the target long lifetime orbit is computed, we

must verify that it too follows the correct long lifetime path. We found that of the

six examples computed, the 85o and 95o orbits follow the correct manifold path and

the others do not. For the four that do not follow the correct manifold path, a bias in

the argument of periapsis is included. This bias does not correspond to the longest

lifetime orbit, but to a long lifetime orbit that follows the correct manifold path and

has a lifetime in the 100-110 day vicinity. This range was chosen so that all of the

target orbits have approximately the same lifetimes. This eases the comparison of

the costs associated with the transfers.

142

Page 156: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 6.1: Examples of Long Lifetime Orbit Reset

Frozen Orbit ∆v1 ∆v2 Target Trajectory Target Trajectory Cost/dayInclination (m/s) (m/s) Lifetime (days) Bias (rad) (cm/s)

70o 16.30 17.94 110 0.06 31.175o 19.07 21.33 106 0.06 38.185o 11.20 15.67 99 – 27.195o 15.45 19.76 103 – 34.2105o 12.32 16.19 106 -0.04 26.9110o 14.36 18.62 107 -0.095 30.8

Table 6.1 shows the details of the examples, identified by the inclination of the

associated frozen orbit of the original long lifetime trajectory. For each orbit, the

magnitude of each of the two burns, the lifetime of the target trajectory and the

bias used for the target trajectory (if necessary) are given. The point of comparison

for each example is the average cost per day. This is computed by dividing the total

cost of the two maneuvers by the lifetime of the target trajectory. It is clear that

this method has a much lower cost than both the radial and transverse low-thrust

methods. Specifically, the average cost per day here is an order of magnitude less

than the transverse thrust law and two orders of magnitude less than the radial

thrust law. Figure 6.12 shows the 1-DOF manifold of an original long lifetime

trajectory, the integration of an original long lifetime trajectory and the integration

of a target trajectory. Observe that the target trajectory and original trajectory are

very close to each other. Figure 6.13 shows the time histories of the orbital elements

for the original long lifetime trajectory and the target trajectory. Observe that the

orbital elements of both trajectories are very close to each other. This is important

since resetting the orbit is only useful if the extension of the orbit still satisfies the

science requirement of the mission, which it should if this scheme is used.

143

Page 157: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08

−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

k=ecosω

h=es

inω

Figure 6.12: Original (black) and target (red) long lifetime trajectories.

0 20 40 60 80 100 1201695

1700

1705

1710

a (k

m)

0 20 40 60 80 100 1200

0.05

0.1

e

0 20 40 60 80 100 12084

85

86

87

i (de

g)

time (days)

0 20 40 60 80 100 1201500

1600

1700

1800

1900

radi

us (

km)

0 20 40 60 80 100 1200

100

200

300

400

ω (

deg)

0 20 40 60 80 100 1200

100

200

300

400

Ω (

deg)

time (days)

Figure 6.13: Time histories of the orbital elements of the original (black) and target (red) longlifetime trajectories.

144

Page 158: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

6.3 Orbit Uncertainty

Up to this point, we have assumed that it is possible to attain the desired long

lifetime orbits. However, in a real mission, it is important to take the uncertainty

of the orbiter’s position and velocity into account. In [40], Thompson, et al studied

orbit determination for a 30-day mission to Europa. They performed simulations

using a science orbit with similar characteristics to our long lifetime orbits. In

particular, their orbit is near-polar, near circular and has an average 100 km altitude.

One of the biggest drivers of the uncertainty of orbits about Europa is the lack of

knowledge of the Europa gravity field. Only the degree two terms of the Europa

gravity field are currently known, and they were determined from Galileo flybys of

Europa [4].

The first study performed in [40] uses a worst-case gravity error, which assumes

that no additional knowledge of Europa’s gravity field is known during the science

phase of the mission. This results in errors in the spacecraft’s radial, crosstrack

and alongtrack position components on the order of kilometers. These results agree

with those in [15]. These errors are judged to be unacceptable for a Europa orbiter

mission, and so a further assumption is made. It is assumed that prior to the science

orbit phase of the mission, two weeks of tracking data during orbit conditions similar

to the science orbit are available. This data reduces the error in the gravity field.

The best-case scenario, which is a lower bound on the gravity error, is computed to

provide a gravity field estimate up to degree 20. This results in predicted position

errors of 5.1m, 7.3m and 5.7m (3-sigma) for radial, crosstrack and alongtrack position

components respectively. The radial component points from the planetary satellite

to the spacecraft, the alongtrack component points in the direction of the velocity

and the crosstrack component completes the triad. However, they assume that a

more realistic assumption for the gravity error is in between the best case and worst

case (lower and upper) bounds. Using this gravity error results in predicted position

145

Page 159: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

errors 24 hours after a data cutoff of 48m, 60m and 35m for radial, crosstrack and

alongtrack positions respectively (also 3-sigma values).

Using the results in [40] as a guide, we study how errors in the spacecraft’s

position affects the designed long lifetime orbits. To study the orbit uncertainty

of long lifetime orbits, Monte Carlo simulations are performed. A large number of

trajectories are integrated, each with a random initial position error. The initial

position error is randomly determining for radial, crosstrack and alongtrack errors

using a Gaussian distribution with a particular standard deviation. The first set

of results are for 3-sigma values corresponding to the realistic gravity error noted

above. Each of the six long lifetime orbits with different inclinations from section

5.3.2 are considered, with 200 trajectories integrated for each. Recall that the six

long lifetime orbits are initialized at the minimum eccentricity point and follow the

correct unstable manifold but are not the maximum lifetime orbits for each case.

Figure 6.14 shows the Monte Carlo simulation results for the minimum

eccentricity long lifetime orbits. Each cluster contains several points, each of which

represents the stopping point of a trajectory integrated to determine which manifold

path it follows. Observe that in four of the six cases, all of the 200 trajectories follow

the correct manifold path. For the 75o inclination case, 44 of 200 trajectories (22%)

follow the wrong manifold path and for the 85o inclination case, 15 of 200 trajectories

(7.5%) follow the wrong manifold path. These results are explained by taking into

account the lifetimes of the six original long lifetime trajectories (Table 5.2). The

two trajectories with the longest lifetimes are the 75o and 85o inclination cases. A

longer lifetime means that the trajectory follows the 1-DOF manifold more closely.

The shorter lifetimes that follow the correct manifold path therefore correspond

the trajectories that lie slightly above the manifold. Then, the more closely the

correct manifold path is followed, the more likely it is that an initial position error

will cause the trajectory to follow the wrong manifold path. Since the 75o and 85o

146

Page 160: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 6.2: Monte Carlo Results for Long Lifetime Orbits

Frozen Orbit Original Orbit Maximum OrbitInclination Lifetime Correct % Lifetime Correct %

70o 105 days 100 182 days 47.275o 119 days 78.0 174 days 50.585o 110 days 92.5 171 days 52.595o 96 days 100 182 days 56.5105o 95 days 100 172 days 51.0110o 99 days 100 188 days 48.0

inclination orbits have the longest lifetimes, they follow the 1-DOF manifold more

closely and therefore it is not surprising that in some cases an initial position error

results in a trajectory that follows the wrong manifold path. Further, note that the

75o inclination orbit has a longer lifetime than the 85o inclination orbit. It follows

then that a larger percentage of the trajectories in the 75o case follow the wrong

path than the 85o case.

The theory that the closer a long lifetime trajectory is to the 1-DOF manifold,

the more susceptible it is to errors in the initial position can be further verified by

performing Monte Carlo analysis of some maximum long lifetime trajectories. The

same maximum lifetime trajectories as those in Table 5.2 are used, and 200 Monte

Carlo simulations are run for each case. The initial position errors are once again

determined randomly via a Gaussian distribution with the same 3-sigma values as

above. Table 6.2 shows the results, along with the lifetimes of the maximum lifetime

orbits. Observe that the number of trajectories that follow the correct manifold is

approximately 50% in each case. This is exactly what is expected since the maximum

lifetime trajectories follow the 1-DOF manifolds very closely. Therefore, a small

initial position error can bump the trajectory either just above the manifold or just

below the manifold. If it is just above the manifold, it will follow the correct manifold

path and if it is just below the manifold, it will follow the incorrect manifold path.

It is important when planning a mission to a planetary satellite such as Europa

147

Page 161: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.02

−0.025

−0.02

−0.015

−0.01

−0.005

0

k=ecosω

h=es

inω

(a) 70o

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.02−0.03

−0.025

−0.02

−0.015

−0.01

−0.005

k=ecosω

h=es

inω

(b) 75o

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.02−0.035

−0.03

−0.025

−0.02

−0.015

−0.01

k=ecosω

h=es

inω

(c) 85o

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.02−0.035

−0.03

−0.025

−0.02

−0.015

−0.01

k=ecosω

h=es

inω

(d) 95o

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.02−0.03

−0.025

−0.02

−0.015

−0.01

−0.005

k=ecosω

h=es

inω

(e) 105o

−0.02 −0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.02

−0.025

−0.02

−0.015

−0.01

−0.005

0

k=ecosω

h=es

inω

(f) 110o

Figure 6.14: Monte Carlo simulation results for various long lifetime trajectories (200 trajectoriesfor each case). Each point in the plot represents the stopping point of a trajectory once it reachesthe desired (top right) or not desired (bottom left) manifold.

148

Page 162: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

to ensure that the uncertainty error in the position of the spacecraft will not lead

to impact with the planetary satellite. Therefore, using the maximum lifetime orbit

as the nominal science orbit is not the best course of action since there is about

a 50% chance that the spacecraft will deviate too far from the nominal trajectory.

The further above the manifold the long lifetime trajectory is, the more likely it

is to follow the correct manifold path. However, the farther above the manifold

the trajectory is, the shorter lifetime it has. Choosing the nominal trajectory is

therefore a trade off between guaranteeing that it will follow the correct manifold

path and having a long lifetime. In addition, before the mission starts, the orbit

accuracy possible will not be known exactly. The values used here are based on the

assumption that prior to the science portion of the mission the gravity field will be

determined with some level of certainty. Even if a trajectory is chosen such that it

will always follow the correct manifold path for the error levels used here, it is not

certain that this level of orbit accuracy is attainable. Therefore, we discuss in the

next section how to return the spacecraft to its nominal orbit once it has drifted

away due to an initial error.

6.4 Long Lifetime Orbits with Initial Errors

It is inevitable that during the science phase of a mission to a planetary satellite that

the spacecraft will drift away from the nominal trajectory. This will most likely be

due to initial errors in the spacecraft’s position and velocity. Once it is determined

that the spacecraft is no longer on the nominal trajectory, it is important to devise

a method for returning it to the nominal trajectory so that the mission can proceed

as planned. For example, recall in section 6.1 the method to reset the long lifetime

orbit once the spacecraft reaches the unstable manifold. As shown in section 6.3, due

to initial position errors, it is possible that a long lifetime trajectory will follow the

wrong manifold path. If it does so, it is not possible to use the algorithm in section

149

Page 163: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

6.1 to reset the trajectory. Therefore, it is important to return to spacecraft to its

nominal trajectory before it has the opportunity to follow the wrong manifold path.

6.4.1 Design of the Correction Maneuvers

The original long lifetime orbit under consideration is determined as usual from the

minimum eccentricity point. Initial errors are introduced by randomly determining

position errors as in the previous section. A time (in days) is chosen to represent the

amount of time until the correction to the trajectory will be made. The amount of

time (TE) is long enough for the spacecraft to drift away from the nominal trajectory.

The maneuvers to be used for the correction are two burns corresponding to a

Hohmann-type transfer. Therefore, the initial transfer point should be at periapsis

and the target at apoapsis. The initial transfer point is determined by integrating

the erroneous trajectory until the closest periapsis passage to TE (either before or

after). Its initial position and velocity are denoted as r0 and v0, respectively. In

addition, denote its initial longitude of the ascending node as Ω0.

The nominal trajectory is the long lifetime orbit with no initial errors. The goal

of this correction scheme is to transfer the spacecraft to the nominal trajectory. The

difficulty lies in determining how to designate the target on the nominal trajectory.

Unfortunately, it is not as simple as integrating the nominal trajectory for the

same amount of time as the erroneous trajectory and setting the target accordingly.

Using this method creates inconsistencies between the initial and target longitude of

ascending node values which creates difficulties in converging to the transfer orbit.

The method that produces the best results is to integrate the nominal trajectory

until its longitude of ascending node value is equal to Ω0 and then finding the

closest periapsis passage to that point (before or after). This periapsis passage on

the nominal trajectory therefore corresponds to point where the spacecraft on the

erroneous trajectory should be. Then, the target is set to be the first apoapsis

150

Page 164: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Figure 6.15: Schematic of Hohmann transfer between two elliptic orbits.

passage following that periapsis passage on the nominal trajectory. Denote the

target position and velocity as rT and vT , respectively.

As in section 6.2, the goal of the transfer is to reach the target and minimize

the sum of the squares of the costs of the two burns (Eq.(6.33)). The exact same

algorithm is used, with the only change being the ideal, first guess transfer orbit.

The first guess for the first maneuver is an eccentricity change at periapsis such

that the transfer orbit arrives at the target at periapsis. Since, as before, there

are perturbations in the system that prevent the target from being reached by this

simple method, it is used as a first guess and corrections are made. Figure 6.15 is a

diagram of a Hohmann transfer between two elliptic orbits. Let the semi-major axis

of the transfer orbit, atr be given by

atr =r0 + rt

2. (6.42)

Then, the velocity of the first guess transfer orbit is:

vg =

õ

(2

r0

− 1

atr

)v0 , (6.43)

where v0 is the direction of the initial velocity v0.

151

Page 165: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

The overall procedure to compute the maneuvers necessary to return the

spacecraft to its nominal trajectory mirrors the procedure in Section 6.2 and is as

follows:

1. Compute the ideal transfer consisting of an eccentricity change at periapsis

where the velocity is vg and integrate to apoapsis. Let the time required for

this transfer be tf .

2. Use the iteration scheme

δv = δv0 + δv1 + δv2 + · · · , (6.44)

where

δvi+1 = φ−1rv δri (6.45)

to compute the necessary correction to the first burn such that δr = 0.

3. Let ∆v∗1 = δv and ∆v∗2 = vt − v(tf ) and compute new corrections to the

burns, δv1 (Eq.(6.31)) and δv2 (Eq.(6.32)), to minimize the cost using δt = −BA

(Eqs.(6.38),(6.39)).

4. Repeat steps 2 and 3 until δti+1 − δti < ε where ε = 10−9 for our simulations.

6.4.2 Results

The algorithm described above is designed to correct the position and velocity of

a spacecraft after it has drifted away from its nominal trajectory due to an initial

error. The question is, for how long after the initial error is it possible to correct

the position? After a certain amount of time, the correction algorithm will not

converge to a transfer orbit with maneuvers that have acceptable costs. The longer

a spacecraft follows an erroneous trajectory, the more difficult is it to correct.

Therefore, after a certain point, as long as the spacecraft follows the correct manifold

152

Page 166: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

path, it makes more sense to reset the orbit once it reaches the unstable manifold,

as in section 6.1, instead of using the technique described in this section.

The first set of examples presented have initial errors that follow the realistic

model of gravity field uncertainty described in [40]. For each long lifetime orbit case,

random initial position errors in the radial, crosstrack and alongtrack directions

are introduced, and the trajectory is integrated for 1, 5, 10 and 20 days before the

correction is made. For each case, 40 simulations are performed, and the average

cost for each maneuver is computed. Corrections are not made after periods of time

longer than 20 days since after that length of time, the trajectory has drifted far

enough away from the nominal trajectory such that this method is not feasible. This

is likely because the method is based on an initial guess which is the ideal transfer

occurring in the absence of all perturbations. It also assumes that the initial and final

orbits are co-planar which is, in general, not the case. However for corrections done

within 20 days of the initial error, these assumptions are valid and the algorithm

converges to a transfer orbit with low-cost maneuvers. After the trajectory has

evolved for longer, the algorithm does not always converge to a transfer orbit, and

when it does, the costs are sometimes very large. Therefore, this method is only

valid when the spacecraft has not drifted too far from the nominal trajectory.

The resulting costs for the two maneuvers with the error model described above

are shown in Table 6.3. The factors to be considered when determining when the

correction should be made are: how quickly after discovering that the trajectory is

not on the nominal trajectory the correction can be performed and when it is most

cost-effective to perform the correction. In examining Table 6.3, we see that the

lowest cost correction occurs at the 1 day mark. However, depending on when it

becomes known that the spacecraft is not on the nominal trajectory, it might not be

possible to perform the maneuver after 1 day. Therefore, the fact that the cost is

still relatively low for corrections after 5 and 10 days is a useful result showing that

153

Page 167: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 6.3: Average cost to correct a long lifetime trajectory after an initial position error

1 day 5 days 10 days 20 daysFrozen Orbit ∆v1 + ∆v2 ∆v1 + ∆v2 ∆v1 + ∆v2 ∆v1 + ∆v2

Inclination m/s m/s m/s m/s70o 0.0989 0.1441 0.3733 0.544985o 0.2190 0.3589 0.8830 2.388595o 0.1642 0.2986 0.6010 3.7737105o 0.1547 0.2949 0.3299 1.8024

it is not necessary to perform the correction very close to the time when the error is

discovered. These results show that performing a correction maneuver within about

a week or two from the time of the initial error is feasible.

The distribution used to compute initial errors for the results in Table 6.3 are

those designated in [40] as the most likely orbit uncertainty values for the science

portion of a mission to Europa. As previously noted, these orbit uncertainty values

were obtained by assuming that prior to the science phase of the mission, the

gravity field of Europa could be determined more accurately than the currently

available values. Therefore, it is important to study the effect of larger levels of

orbit uncertainty on the ability to make corrections to the spacecraft’s position.

Table 6.4 shows results corresponding to initial error values double the size of those

in Table 6.3. In other words, the standard deviations of the error distributions

are two times the values in [40]. It is not surprising that an increase in the initial

position error leads to a decrease in the length of time that can elapse for a correction

to be possible. For the results in Table 6.4, attempts to do a correction after 20

days caused errors in the convergence of the transfer orbit, and large costs when a

transfer orbit was found. Therefore, the results shown are for corrections after 1 day,

5 days, 10 days and 15 days. First, note that the total cost to return the spacecraft

to its nominal trajectory is larger in this case than for the results in Table 6.3.

This is not surprising since the initial errors are larger in this case, and so larger

154

Page 168: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

Table 6.4: Average cost to correct a long lifetime trajectory after a larger initial position error

1 day 5 days 10 days 15 daysFrozen Orbit ∆v1 + ∆v2 ∆v1 + ∆v2 ∆v1 + ∆v2 ∆v1 + ∆v2

Inclination m/s m/s m/s m/s70o 0.2256 0.2466 0.8433 0.796685o 0.3746 0.7773 1.4339 1.329495o 0.4414 0.6415 1.3459 2.4245105o 0.2733 0.3783 0.7837 1.8184

maneuvers are needed. However, the costs for the corrections after 1 day, 5 days and

sometimes 10 days are still relatively low, and corrections are feasible once again

even if it is necessary to wait for a few days after the initial error. The results for

both error models are consistent since as time goes on, the deviation between the

erroneous trajectory and the nominal trajectory gets larger, increasing the cost of

the correction in the same way that a larger initial error also increases the cost of

the correction.

155

Page 169: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

CHAPTER 7

CONCLUSIONS AND FUTURE DIRECTIONS

The main focus of this dissertation is to develop an understanding of spacecraft

dynamics in the vicinity of planetary satellites and to use these dynamics for mission

design. The main contributions are:

• The analysis of the 1-DOF system dynamics including the frozen orbits, their

stability and their dependence on the higher order gravity field terms of the

planetary satellite

• The analysis of capture trajectories in safe zones that allow an uncontrolled

spacecraft to remain in orbit and from which transfers to lower altitude orbits

are possible

• The design method for long lifetime science orbits using the stable and unstable

manifolds of the frozen orbits

• The optimal transfer strategy for restarting a long lifetime orbit using the fact

that the manifolds are offset from the origin

• The demonstration that it is possible, with relatively low costs, to return the

spacecraft to its nominal trajectory after initial position errors

156

Page 170: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

7.1 Summary of the Results

The first main result obtained is the derivation of frozen orbit solutions in the

1-DOF system that are relevant for planetary satellites. When the J3 planetary

satellite gravity field term is included in the model, the structure of the frozen

orbit solutions change, qualitatively independent of the size of J3. Therefore, this

analysis is applicable to any planetary satellite system that can be approximated by

the Hill problem, no matter what its J3 value is. Both stable and unstable frozen

orbit solutions exist, and the stable frozen orbit solutions are long-term stable in

the 3-DOF system as long as their altitude is not too large. These stable orbits are

not of interest from a scientific perspective, however, due to their low inclinations.

The unstable frozen orbit solutions are of interest to scientific observations due to

their high (near-polar) inclinations. Thus these orbits are used to motivate science

orbit design. These are especially challenging as an uncontrolled, or poorly designed,

science orbit will impact on the planetary satellite surface in a relatively short time.

The investigation of capture trajectories in the Hill 3-body problem is

accomplished using a Periapsis Poincare map and a specific application is made to

trajectories about Jupiter’s moon Europa. However, the same analysis could be

repeated for other planetary satellites of interest. Trajectories that don’t impact

with Europa or escape for one week time periods are identified and the regions in

which they occur denoted as safe zones. These regions are characterized by the first

periapsis of each trajectory. It is an important result that the determination that

a trajectory will be safe for one week can be made from its first periapsis passage

location. These safe zones are evaluated to find trajectories from which it is possible

to transfer to long-term stable orbits with a relatively low cost. Orbits considered

as possible targets are both elliptic and circular frozen orbits which are either

long-term stable or can be stabilized with small control maneuvers. We found that

the lowest-cost method to transfer to a circular orbit is by using the dynamics of the

157

Page 171: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

system to decrease the eccentricity rather than circularizing the orbit directly. In

particular we describe a low cost sequence that results in a circular orbit reachable

by a low-thrust spacecraft. From this orbit it could then spiral down into a lower

altitude circular orbit.

The design of science orbits with long lifetimes begins with the identification of

the stable and unstable manifolds of the low-altitude, near-polar unstable frozen

orbits in the 1-DOF system. The stable to unstable manifold path that has the

largest change in argument of periapsis while the orbit stays near-circular is identified

as the desired path for a long lifetime orbit. Following that, an algorithm is derived

that allows us to relate initial conditions on the 1-DOF manifold to initial conditions

in the 3-DOF system with an intermediate step in the 2-DOF system. This

algorithm allows for the semi-analytic computation of the desired initial conditions,

and easily incorporates changes in the system parameters. In addition, it utilizes the

main dynamical features of the system to extend the lifetimes as compared to the

unstable frozen orbits. The use of this method for Europa results in science orbits

with lifetimes of at least 100 days. Therefore, the use of this orbit protects the

spacecraft in the case of control maneuver failures. This method is also applied to

Saturn’s moons Enceladus and Dione to show that it is applicable to other planetary

satellites. Viable science orbits are designed in those cases, but with shorter lifetimes

since their environments are more highly perturbed. A toolbox for computing initial

conditions for long lifetime science orbits that was written and delivered to the Jet

Propulsion Laboratory is also described.

The control of long lifetime science orbits has been investigated for a few different

scenarios. The first attempt maintains constant orbital elements using continuous

thrust. However, the thrust levels required for this type of control are found to be

too large to be practical for a mission. Next, scenarios where a long lifetime science

orbit reaches the unstable manifold and needs to be reset to the stable manifold

158

Page 172: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

is studied. For this scenario, an algorithm to redesign a long lifetime orbit based

on the parameters of the trajectory at the unstable manifold is derived. Then, a

two-maneuver sequence is computed to execute a transfer to the new transfer orbit.

The total cost of these maneuvers is optimized, and fuel costs of around 30 cm/s/day

are obtained. Next, the effect of orbit uncertainty on the long lifetime science

orbits is investigated using Monte Carlo simulations. We found that the closer the

trajectory is to the 1-DOF manifold, the more susceptible it is to veering off the

desired manifold path due to an initial position error. Finally, a scheme is developed

to return the spacecraft to its nominal trajectory following an initial position error.

A transfer orbit is optimized by the total cost of the two maneuvers necessary, and

comparisons are made based on how many days after the initial error the correction

is executed. We found that even if it is necessary to wait 5 days after the initial

error, the cost to return the spacecraft to the nominal trajectory is less than 0.4 m/s.

7.2 Future Directions

In this section, a few possible extensions of the research described in this dissertation

are discussed.

7.2.1 Transfers from Capture Trajectories

In our discussion of safe zones for captured trajectories, the safe zones are identified

by their energy level (i.e. their Jacobi integral value). Transfers from safe capture

trajectories to stable frozen orbits are then computed using impulsive maneuvers

whereby the entire set of trajectories remain at the same energy level. An extension

of this work is to derive low-cost transfers that traverse safe zones at different energy

levels rather than staying within one energy level. This is of particular interest since

low-thrust trajectories are becoming more desirable with the increased use of electric

propulsion in spacecraft. In addition, rather than stopping the scheme at a stable

159

Page 173: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

frozen orbit at a higher altitude, the low-thrust trajectory could be continued using

a spiraling-in approach to transfer the spacecraft to a low-altitude science orbit.

7.2.2 Science Orbit Design

The most logical extension for the science orbit design is to incorporate the science

orbits designed in this dissertation into precision navigation and mission design

programs that include detailed models of all of the physics acting on the system.

This is probably the most effective way to design science orbits for a real mission

since the advantage of our orbit design scheme is its semi-analytic results that are

only possible since the system is not too complex. However, since the system we use

is an approximation to the true system, at some point the more complicated true

system must be incorporated into the orbit design process.

7.2.3 Control of Long Lifetime Science Orbits

There are many issues that could be explored regarding the control of long lifetime

science orbits. Perhaps the most important is ensuring that the spacecraft remains

on the nominal trajectory. We have shown that it is possible to return the spacecraft

to the nominal trajectory after a few days following an initial position error. A more

robust technique would be to devise active controls such that the spacecraft always

stays on the nominal trajectory. This is particularly important in cases where the

system is so unstable that even leaving the trajectory uncontrolled for one day would

be disastrous. One such example is Enceladus, where we showed that long lifetime

science orbits have lifetimes of only 10 days.

Another possible extension concerns resetting the long lifetime science orbit

before impact with the planetary satellite. The transfer method described in this

dissertation involves computing the target (initial conditions of a new long lifetime

orbit) and then optimizing the transfer over the total cost. An improvement to

this method would have the optimization take place over both the total cost and

160

Page 174: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

the target itself. The goal of this resetting scheme is to transfer the spacecraft to

another long lifetime trajectory, but the exact location where the spacecraft ends

up on the new trajectory is not important. Therefore an optimization procedure

that allowed for variations of the target on the desired trajectory would most likely

produce lower-cost maneuvers than the current one which sets the target before the

optimization takes place.

161

Page 175: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

APPENDIX

162

Page 176: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

APPENDIX A

NUMERICAL ALGORITHMS

This appendix details the algorithms used for the numerical computations of

initial conditions for science orbits in Chapter 5.

A.1 Romberg Algorithm

The number of iterations performed in this algorithm is defined by a number

n. Larger values of n correspond to a higher accuracy in the result, but also

longer computation times since more computations are being performed. We chose

n = 10 since we found that that value gives us a good estimate while keeping the

computation time not too large.

Purpose: Approximate the integral Ij,k which is the entry in the jth row and kth

column of:

I =

∫ b

a

exp (At)dt

1. h = b− a

2. R1,1 = [exp (Aa)]j,k + [exp (Ab)]j,k/2h

3. for i=2 to n

(a) sum = 0

(b) m = 2(i−2)

(c) for k=1 to m

163

Page 177: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

i. sum = sum + [exp (A(a + (k − 0.5)h))]j,k

(d) R2,1 = (R1,1 + h ∗ sum)/2

(e) for j=2 to i

i. L = 22(j−1))

ii. R2,j = R2,j−1 + (R2,j−1 −R1,j−1)/(L− 1)

(f) h = h/2

(g) for j=1 to i

i. R1,j = R2,j

4. Ij,k = R2,n

A.2 Gaussian Quadrature

For the Gaussian algorithm for double integrals, a set of weights and abscissas

must be specified. The values that we use[1] are denoted rj,k and coj,k, respectively.

The number of iterations performed is given by two numbers n and m. We chose

n = m = 12, which gives us a good approximation of the integral while ensuring not

too long of a computation time.

Purpose: Approximate the integral Ik which is the kth entry of:

I =

∫ b

a

∫ τ

0

exp (A(t− τ))B(τ)dτdt

1. h1 = (b− a)/2

2. h2 = (b + a)/2

3. aj = 0

4. for i=1 to m

(a) t = h1 ∗ rm−1,i + h2

164

Page 178: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

(b) jx = 0

(c) c1 = 0

(d) d1 = t

(e) k1 = (d1− c1)/2

(f) k2 = (d1 + c1)/2

(g) for j=1 to n

i. τ = k1 ∗ rn−1,j + k2

ii. q = [exp (A(t− τ))B]k

iii. jx = jx + con−1,jq

(h) aj = aj + com−1,ik1 ∗ jx

5. aj = aj ∗ h1

6. Ik = aj

A.3 Gaussian Quadrature with Kepler Equation

Modification

For the Gaussian algorithm for double integrals, a set of weights and abscissas

must be specified. The values that we use[1] are denoted rj,k and coj,k, respectively.

The number of iterations performed is given by two numbers n and m. We chose

n = m = 10 in this case, which gives us a good approximation of the integral while

ensuring not too long of a computation time. The number of iterations is smaller

than in Appendix B since this algorithm is used on a much more complex system

and increasing the number of iterations any larger than 10 resulted in prohibitively

long run times.

The function kepler(τ, e) computes the true anomaly ν given a mean anomaly τ

and an eccentricity e. This is done by using the built-in Matlab function ‘solve’ to

165

Page 179: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

find the roots of

τ − E + e sin E = 0 (A.1)

where E is the eccentric anomaly. Then,

ν = cos−1 ((cos(E)− e)/(1− e cos(E))) (A.2)

Purpose: Approximate the integral Ik which is the kth entry of:

I =1

∫ 2π

0

∫ M

0

eA(M−τ)B(ν(τ))dτdM

1. h1 = (b− a)/2

2. h2 = (b + a)/2

3. aj = 0

4. for i=1 to m

(a) M = h1 ∗ rm−1,i + h2

(b) jx = 0

(c) c1 = 0

(d) d1 = M

(e) k1 = (d1− c1)/2

(f) k2 = (d1 + c1)/2

(g) for j=1 to n

i. τ = k1 ∗ rn−1,j + k2

ii. ν = kepler(τ, e)

iii. q = [exp (A(M − τ))B(ν)]k

iv. jx = jx + con−1,jq

(h) aj = aj + com−1,ik1 ∗ jx

166

Page 180: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

5. aj = aj ∗ h1

6. Ik = aj

167

Page 181: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

BIBLIOGRAPHY

[1] Abscissas and weights of gauss-legendre integra-tion. From efunda: engineering fundamentals.http://www.efunda.com/math/num integration/findgausslegendre.cfm, 2005.

[2] Pioneer 3, 4. From JPL Mission and Spacecraft Library.http://leonardo.jpl.nasa.gov/msl/QuickLooks/pioneer3QL.html, 2006.

[3] J. Aiello. Personal Communication, August 2004.

[4] J.D. Anderson, G. Schubert, R.A. Jacobson, E.L. Lau, W.B. Moore, and W.L.Sjogren. Europa’s differentiated internal structure: Inferences from four Galileoencounters. Science, 281:2019–2022, 1998.

[5] S.A. Astakhov, A.D. Burbanks, S. Wiggins, and D. Farrelly. Chaos-assistedcapture of irregular moons. Nature, 423:264–267, 2003.

[6] et al. Belton, M.J.S. Galileo’s first images of Jupiter and the Galilean satellites.Science, 274(5286):377–385, 1996.

[7] R.A. Broucke. Long-term third-body effects via double averaging. Journal ofGuidance, Control and Dynamics, 26(1):27–32, 2003.

[8] D. Brouwer and G.M. Clemence. Methods of Celestial Mechanics. AcademicPress Inc., New York, NY, 1961.

[9] L. Buck, C.F. Chyba, M. Goulet, A. Smith, and P. Thomas. Persistenceof thin ice regions in Europa’s ice crust. Geophysical Research Letters,29(22):2055–2058, 2002.

[10] R.L. Burden and J.D. Faires. Numerical Analysis, 6th ed. Brooks/ColePublishing Company, Pacific Grove, CA, 1997.

[11] V.A. Chobotov, editor. Orbital Mechanics, Second Edition. American Instituteof Aeronautics and Astronautics, Inc., Reston, VA, 1996.

[12] J.M.A. Danby. Fundamentals of Celestial Mechanics, 2nd Ed. Willmann-Bell,Richmond, VA, 1992.

168

Page 182: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

[13] R. Dvorak. Progress in the elliptic restricted 3-body problem: Asteroids in the3/1, 2/1 and 1/1 resonance. Celestial Mechanics and Dynamical Astronomy,54(1-3):195–205, 1992.

[14] D.T. Greenwood. Principles of Dynamics, Second Edition. Prentice-Hall, Inc.,Upper Saddle River, New Jersey, 1988.

[15] M.D. Guman, D.C. Roth, and B.G Williams. Navigation feasibility studies forthe Europa orbiter mission. Part I, Advances in the Astronautical Sciences,99:161–174, 1998.

[16] R.S. Harrington. The stellar three-body problem. Celestial Mechanics,1(1):200–209, 1969.

[17] M. Henon. New families of periodic orbits in Hill’s problem of three bodies.Celestial Mechanics and Dynamical Astronomy, 85(3):223–246, 2003.

[18] M. Henon and J.-M. Petit. Series expansions for Hill’s problem. CelestialMechanics, 38:67–100, 1986.

[19] G.W. Hill. Researches in the lunar theory. American Journal of Mathematics,1:5–26, 1878.

[20] W.S. Koon, M.W. Lo, J.E. Marsden, and S.D. Ross. Heteroclinic connectionsbetween periodic orbits and resonance transitions and celestial mechanics.Chaos, 10(2):427–469, 2000.

[21] Y. Kozai. Secular perturbations of asteroids with high inclination andeccentricity. Astronomical Journal, 67(9):591–598, 1962.

[22] M. Lara, J.F. San-Juan, and S. Ferrer. Secular motion around tri-axial,sychronously rotating planetary satellites: Application to Europa. Part II,Advances in the Astronautical Sciences, 120:1147–1166, 2005.

[23] M.L. Lidov and S.L. Ziglin. Non restricted double-averaged three body problemin Hill’s case. Celestial Mechanics, 13(4):471–489, 1976.

[24] O. Montenbruck and E. Gill. Satellite Orbits: Models, Methods, andApplications. Springer-Verlag, Germany, 2001.

[25] M.E. Paskowitz and D.J. Scheeres. Identifying safe zones for planetary satelliteorbiters. In Proceedings of the AIAA/AAS Astrodynamics Specialist Conferenceat Providence, RI, paper AIAA 2004-4862, 2004.

[26] M.E. Paskowitz and D.J. Scheeres. Orbit mechanics about planetary satellites.In Proceedings of the AAS/AIAA Space Flight Mechanics Meeting at Maui, HI,paper AAS 04-244, 2004.

169

Page 183: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

[27] M.E. Paskowitz and D.J. Scheeres. Orbit mechanics about planetary satellitesincluding higher order gravity fields. In Proceedings of the AAS/AIAA SpaceFlight Mechanics Meeting at Copper Mountain, CO, paper AAS 05-190, 2005.

[28] M.E. Paskowitz and D.J. Scheeres. Transient behavior of planetary satelliteorbits. In Proceedings of the AAS/AIAA Astrodynamics Specialist Conferenceat South Lake Tahoe, CA, paper AAS 05-358, 2005.

[29] M.E. Paskowitz and D.J. Scheeres. Design of science orbits about planetarysatellites: Application to Europa. Journal of Guidance, Control, and Dynamics,29(5):1147–1158, 2006.

[30] M.E. Paskowitz and D.J. Scheeres. Robust capture and transfer trajectoriesfor planetary satellite orbiters. Journal of Guidance, Control, and Dynamics,29(2):342–353, 2006.

[31] M.E. Paskowitz and D.J. Scheeres. A toolbox for designing long lifetime orbitsabout planetary satellites: Application to JIMO at Europa. In Proceedings ofthe AAS/AIAA Space Flight Mechanics Meeting at Tampa, FL, paper AAS06-191, 2006.

[32] M. Paskowitz Possner and D.J. Scheeres. Control of science orbits aboutplanetary satellites. in preparation, 2006.

[33] L. Rowan. Cassini drops in. Science, 307(5713):1222, 2005.

[34] J.F. San-Juan, M. Lara, and S. Ferrer. Phase space structure around oblateplanetary satellites. Journal of Guidance, Control and Dynamics, 29(1):113–120,2006.

[35] D.J. Scheeres. The restricted hill four-body problem with applications tothe earth-moon-sun system. Celestial Mechanics and Dynamical Astronomy,70:75–98, 1998.

[36] D.J. Scheeres and J. Bellerose. The restricted hill full 4-body problem:application to spacecraft motion about binary asteroids. Dynamical Systems:An International Journal, 20(1):23–44, 2005.

[37] D.J. Scheeres, M.D. Guman, and B.F. Villac. Stability analysis of planetarysatellite orbiters: Application to the Europa orbiter. Journal of Guidance,Control and Dynamics, 24(4):778–787, 2001.

[38] D.J. Scheeres and F. Marzari. Spacecraft dynamics in the vicinity of a comet.Journal of Astronautical Sciences, 50(1):35–52, 2002.

[39] D. Stevenson. Europa’s ocean – the case strengthens. Science, 289(5483):1305–1307, 2000.

170

Page 184: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

[40] P.F. Thompson, S. Nandi, and M.C. Wong. Orbit determination studiesfor a low-altitude Europa orbiter. paper presented at the 2006 Space FlightMechanics Meeting, Tampa, Florida, AAS 06-192, 2006.

[41] B.F. Villac. Dynamics in the Hill Problem with Application to SpacecraftManeuvers. PhD thesis, The University of Michigan, 2003.

[42] B.F. Villac and D.J. Scheeres. Escaping trajectories in the Hill 3-body problemand applications. Journal of Guidance, Control and Dynamics, 26(2):224–232,2003.

[43] Eric Weisstein. Point-line distance–two-dimensional. From MathWorld–AWolfram Web Resource. http://mathworld.wolfram.com/Point-LineDistance2-Dimensional.html, 2003.

171

Page 185: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

ABSTRACT

ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE

ORBITERS IN THE HILL 3-BODY PROBLEM

by

Marci Paskowitz Possner

Chair: Daniel J. Scheeres

The exploration of planetary satellites by robotic spacecraft is currently of strong

scientific interest. However, sending a spacecraft to a planetary satellite can be

challenging due to strong perturbations from the central planet. The primary goal of

this dissertation is to identify and utilize the main dynamical features of the system

in the orbit design process.

The system is modeled using a modified form of Hill’s 3-body problem, where

the effect of the planetary satellite’s gravity field is included in the low-altitude

analysis. A thorough study of the dynamics of the system is performed by applying

averaging theory to reduce the complexity and degrees of freedom of the system.

The reduced system has one degree of freedom (DOF) and has equilibrium solutions

called frozen orbits. These frozen orbits are first used as targets for transfers from

capture trajectories in ‘safe zones’. The ‘safe zones’ in phase space are numerically

determined; they contain trajectories that enter the Hill region and allow an

uncontrolled spacecraft to remain in orbit without impact or escape for specified

time periods. Transfers from safe trajectories to frozen orbits are identified and

criteria on their costs evaluated.

Unstable low-altitude, near-polar frozen orbits are the basis for the design of long

Page 186: ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE … · 2017-09-21 · ORBIT DESIGN AND CONTROL OF PLANETARY SATELLITE ORBITERS IN THE ... this thesis would never

lifetime science orbits. The stable and unstable manifolds of these frozen orbits in

the 1-DOF system are investigated and the desired path for long lifetime orbits is

identified. An algorithm is developed to systematically compute initial conditions in

the full system such that the orbits follow the desired path and have sufficiently long

lifetimes to be practical as science orbits about planetary satellites. The analysis of

the control of a planetary satellite orbiter begins with the evaluation of the effect

of orbit uncertainty on the science orbits and the identification of criteria to ensure

that the orbits have the desired behavior. Then, two control schemes are developed:

a) given the terminal conditions of a science orbit, redesign a new science orbit and

execute a low-cost transfer to it, b) return the spacecraft to its nominal trajectory

via a two-sequence set of maneuvers.

2